Citation: Bertrand R. Caré, Pierre-Emmanuel Emeriau, Ruggero Cortini, Jean-Marc Victor. Chromatin epigenomic domain folding: size matters[J]. AIMS Biophysics, 2015, 2(4): 517-530. doi: 10.3934/biophy.2015.4.517
[1] | Shengxiang Wang, Xiaohui Zhang, Shuangjian Guo . The Hom-Long dimodule category and nonlinear equations. Electronic Research Archive, 2022, 30(1): 362-381. doi: 10.3934/era.2022019 |
[2] | Yongjie Wang, Nan Gao . Some properties for almost cellular algebras. Electronic Research Archive, 2021, 29(1): 1681-1689. doi: 10.3934/era.2020086 |
[3] | Xing Zhang, Xiaoyu Jiang, Zhaolin Jiang, Heejung Byun . Algorithms for solving a class of real quasi-symmetric Toeplitz linear systems and its applications. Electronic Research Archive, 2023, 31(4): 1966-1981. doi: 10.3934/era.2023101 |
[4] | Juxiang Sun, Guoqiang Zhao . Gorenstein invariants under right Quasi-Frobenius extensions. Electronic Research Archive, 2025, 33(6): 3561-3570. doi: 10.3934/era.2025158 |
[5] | Shuguan Ji, Yanshuo Li . Quasi-periodic solutions for the incompressible Navier-Stokes equations with nonlocal diffusion. Electronic Research Archive, 2023, 31(12): 7182-7194. doi: 10.3934/era.2023363 |
[6] | Natália Bebiano, João da Providência, Wei-Ru Xu . Approximations for the von Neumann and Rényi entropies of graphs with circulant type Laplacians. Electronic Research Archive, 2022, 30(5): 1864-1880. doi: 10.3934/era.2022094 |
[7] | Jincheng Shi, Shuman Li, Cuntao Xiao, Yan Liu . Spatial behavior for the quasi-static heat conduction within the second gradient of type Ⅲ. Electronic Research Archive, 2024, 32(11): 6235-6257. doi: 10.3934/era.2024290 |
[8] | Wenjie Zuo, Mingguang Shao . Stationary distribution, extinction and density function for a stochastic HIV model with a Hill-type infection rate and distributed delay. Electronic Research Archive, 2022, 30(11): 4066-4085. doi: 10.3934/era.2022206 |
[9] | Jianxing Du, Xifeng Su . On the existence of solutions for the Frenkel-Kontorova models on quasi-crystals. Electronic Research Archive, 2021, 29(6): 4177-4198. doi: 10.3934/era.2021078 |
[10] | Xuerong Hu, Yuxiang Han, Junyan Lu, Linxiang Wang . Modeling and experimental investigation of quasi-zero stiffness vibration isolator using shape memory alloy springs. Electronic Research Archive, 2025, 33(2): 768-790. doi: 10.3934/era.2025035 |
The theory of (co)monads can be used as a tool in various fields of mathematics such as algebra, logic or operational semantics, and theoretical computer science. Note that in algebra theory, there are two different "bimonads". On the one hand, bimonads and Hopf monads without monoidal structures were introduced in [1], and developed in [2,3,4]. On the other hand, bimonads on monoidal categories were introduced in [5]. In 2002, Moerdijk used an opmonoidal monad to define a bimonad. This bimonad $ F $ is both a monad and an opmonoidal functor satisfying the multiplication and the unit of $ F $ are all monoidal natural transformations (see [5] for details). Although Moerdijk called his bimonad "Hopf monad", the antipode was not involved in his definition. In 2007, A. Bruguières and A. Virelizier introduced the notion of Hopf monad with antipode in the rigid categories in [6], and then put it in the non-dual monoidal categories [7]. We refer to [7,8,9,10,11] for the recent research on A. Bruguières and A. Virelizier's bimonads.
Quasi-bialgebras were introduced by V. G. Drinfel'd in [12]. The dual definition, a $ k $-coquasi-bialgebra $ H $ (or a Majid algebra), was introduced by S. Majid in [13]. The associativity of the multiplication are replaced by a weaker property, called coquasi-associativity. The multiplication is associative up to conjugation by a convolution invertible linear form $ \omega \in (H {\otimes} H {\otimes} H)^\ast $, called the coassociator. Note that the definition of a coquasi-bialgebra is not selfdual, and the category of (left or right) comodules over a coquasi-bialgebra is a monoidal category with nontrivial associativity constraint and nontrivial unit constraints. Coquasi-bialgebras in a braided monoidal category also have been studied in [14].
Taking into account the results proved A. Bruguières and A. Virelizier in [6], it is now very natural to ask how to extend coquasi-bialgebras to the non-braided setting. This is the main motivation of the present paper.
In this paper, we present a dual version of the second author's results about quasi-bimonads which appeared in [15]. We mainly provide a generalization of coquasi-bialgebras by introducing the notion of quasi-monoidal comonad. Actually, a quasi-monoidal comonad $ F $ is both a comonad and a quasi-monoidal functor such that its corepresentations is a non-strict monoidal category. The notion of quasi-monoidal comonad is very general. For example, the tensor functor of a (Hom-type) coquasi-bialgebras and bicomonads are all special cases of quasi-monoidal comonads.
The paper is organized as follows. In Section 2 we recall some notions of comonads, quasi-monoidal functors, $ \pi $-categories and so on. In Section 3, we introduce the definition of quasi-monoidal comonads and discuss their corepresentations. In Section 4, we mainly investigate the coquasitriangular structures of a quasi-monoidal comonad. At last, we introduce the gauge equivalent relation on quasi-monoidal comonads.
Throughout the paper, we let $ k $ be a fixed field and $ char(k) = 0 $ and $ Vec_k $ be the category of finite dimensional $ k $-spaces. All the algebras and coalgebras, modules and comodules are supposed to be in $ Vec_k $. For the comultiplication $ {\Delta} $ of a $ k $-space $ C $, we use the Sweedler-Heyneman's notation: $ \Delta(c) = \sum c_{1}{\otimes} c_{2} $ for any $ c\in C $.
Let $ (\mathcal {C}, {\otimes}, I, a, l, r) $ and $ (\mathcal {C}', {\otimes}', I', a', l', r') $ be two monoidal categories. Recall that a quasi-monoidal functor from $ \mathcal {C} $ to $ \mathcal {C}' $ is a triple $ (F, F_2, F_0) $, where $ F:\mathcal {C}\rightarrow \mathcal {C}' $ is a functor, $ F_2: F\otimes' F \rightarrow F\otimes $ is a natural transformation, and $ F_0:I'\rightarrow FI $ is a morphism in $ \mathcal {C}' $.
Furthermore, if the following equations hold for any $ X, Y, Z \in \mathcal {C} $:
$ F2(X,Y⊗Z)∘(idFX⊗′F2(Y,Z))∘a′FX,FY,FZ=F(aX,Y,Z)∘F2(X⊗Y,Z)∘(F2(X,Y)⊗′idFZ), $
|
(2.1) |
$ F(lX)∘F2(I,X)∘(F0⊗′idFX)=l′FX, $
|
(2.2) |
$ F(rX)∘F2(X,I)∘(idFX⊗′F0)=r′FX, $
|
(2.3) |
then $ F = (F, F_2, F_0) $ is called a monoidal functor.
Let $ \mathcal {C} $ be a category, $ F $: $ \mathcal {C} →
$ F\delta \circ \delta = \delta F\circ \delta,\; \; \mbox{and}\; \; id_F = F\varepsilon\circ \delta = \varepsilon F\circ \delta, $ |
then we call the triple $ (F, \delta, \varepsilon) $ a comonad on $ \mathcal {C} $.
Let $ X \in \mathcal {C} $, and $ (F, \delta, \varepsilon) $ a comonad on $ \mathcal {C} $. If there exists a morphism $ \rho^X $: $ X →
$ F{\rho^X}\circ \rho^X = \delta_X\circ \rho^X ,\; \; \mbox{and}\; \; \varepsilon_X\circ \rho^X = id_X, $ |
then we call the couple $ (X, \rho^X) $ an F-comodule.
A morphism between $ F $-comodules $ g $: $ X \rightarrow X' $ is called $ F $-colinear, if $ g $ satisfies: $ Fg \circ \rho^X = \rho^{X'} \circ g $. The category of $ F $-comodules is denoted by $ \mathcal {C}^F $.
Let $ (\mathcal {C}, {\otimes}, I, a, l, r) $ be a monoidal category, $ (F, \delta, {\varepsilon}) $ be a comonad on $ \mathcal {C} $, and $ (F, F_2, F_0): \mathcal {C}\rightarrow \mathcal {C} $ be a monoidal functor. Then recall from [18] or [19] that $ F $ is called a monoidal comonad (or a bicomonad) on $ \mathcal {C} $ if $ \delta $ and $ {\varepsilon} $ are both monoidal natural transformations, i.e. the following compatibility conditions hold for any $ X, Y \in \mathcal {C} $:
$ \left\{(C1)F(F2(X,Y))∘F2(FX,FY)∘(δX⊗δY)=δX⊗Y∘F2(X,Y),(C2)εX⊗Y∘F2(X,Y)=εX⊗εY,(C3)F(F0)∘F0=δI∘F0,(C4)εI∘F0=idI.
Given a category $ \mathcal {C} $ and a positive integer $ n $, we denote $ \mathcal {C}^n = \mathcal {C} \times\mathcal {C} \times \cdots \times \mathcal {C} $ the $ n $-tuple cartesian product of $ \mathcal {C} $. If $ F $ is a comonad on $ \mathcal {C} $, then $ F^{\times n} $ (the $ n $-tuple cartesian product of $ F $) is a comonad on $ \mathcal {C}^n $, and we have $ {\mathcal {C}^n}^{F^{\times n}} = (\mathcal {C}^F)^n $.
Assume that $ U: \mathcal {C}^F\rightarrow \mathcal {C} $ is the forgetful functor and $ P, Q: \mathcal {C}^n\rightarrow \mathcal {D} $ are functors. Then from [[9], Proposition 4.1], we have the following results.
Lemma 2.1. There is a canonical bijection:
$ Nat(PU^{\times n},QU^{\times n})\cong Nat(PF^{\times n},Q). $ |
Proof. Define $?^\flat:Nat(PU^{\times n}, QU^{\times n})\rightarrow Nat(PF^{\times n}, Q) $, $ f\mapsto f^\flat $, by
$ f♭(X1,⋯,Xn):P(FX1×⋯×FXn)f(FX1,⋯,FXn)→Q(FX1×⋯×FXn)Q(εX1,⋯,εXn)→Q(X1×⋯×Xn), $
|
and $?^\sharp:Nat(PF^{\times n}, Q)\rightarrow Nat(PU^{\times n}, QU^{\times n}) $, $ {\alpha}\mapsto {\alpha}^\sharp $, by
$ α♯(M1,⋯,Mn):P(M1×⋯×Mn)P(ρM1,⋯,ρMn)→P(FM1×⋯×FMn)α(M1,⋯,Mn)→Q(M1×⋯×Mn), $
|
for any $ f \in Nat(PU^{\times n}, QU^{\times n}) $, $ \alpha \in Nat(PF^{\times n}, Q) $ and $ X_i \in \mathcal {C} $, $ (M_i, \rho^{M_i}) \in \mathcal {C}^F $. It is easy to check that $?^\flat $ and $?^\sharp $ are well defined and are inverse with each other.
Let $ P, Q, R:\mathcal {C}^n\rightarrow \mathcal {D} $ be functors. For any $ \alpha \in Nat(PF^{\times n}, Q) $ and $ \beta \in Nat(QF^{\times n}, R) $, define their convolution product $ \beta\ast{\alpha} \in Nat(PF^{\times n}, R) $ by setting, for any objects $ X_1, \cdots, X_n $ in $ \mathcal {C} $,
$ β∗αX1,⋯,Xn=βX1,⋯,Xn∘αFX1,⋯,FXn∘P(δX1,⋯,δXn). $
|
We say that $ \alpha \in Nat(PF^{\times n}, Q) $ is $ \ast $-invertible if there exists $ \beta \in Nat(QF^{\times n}, P) $ such that $ {\beta} \ast {\alpha} = P(\varepsilon^{\times n}) \in Nat(PF^{\times n}, P) $ and $ {\alpha} \ast {\beta} = Q(\varepsilon^{\times n}) \in Nat(QF^{\times n}, Q) $. We denote $ {\beta} $ by $ {\alpha}^{\ast-1} $.
Proposition 2.2. The $ \ast $-invertible elements in $ Nat(PF^{\times n}, Q) $ are in corresponding with the natural isomorphisms in $ Nat(PU^{\times n}, QU^{\times n}) $.
Proof. Suppose that $ f \in Nat(PU^{\times n}, QU^{\times n}) $ is a natural isomorphism. Then we immediately get that $ (f^\flat)^{\ast-1} = (f^{-1})^\flat $.
Conversely, if $ {\alpha} \in Nat(PF^{\times n}, Q) $ is $ \ast $-invertible, then $ {{\alpha}^\sharp}^{-1} = ({\alpha}^{\ast-1})^\sharp $.
Suppose that $ (\mathcal {C}, {\otimes}, I, a, l, r) $ is a monoidal category, $ F:\mathcal {C}\rightarrow \mathcal {C} $ is a functor, $ (F, \delta, {\varepsilon}) $ is a comonad and $ (F, F_2, F_0) $ is a quasi-monoidal functor.
Lemma 3.1. If we define the $ F $-coaction on $ I $ by $ F_0 $, anddefine the $ F $-coaction on $ M {\otimes} N $ (as the tensor product in $ \mathcal {C} $) for any $ (M, \rho^M), (N, \rho^N) \in \mathcal {C}^F $ by:
$ \rho^{M {\otimes} N}: M {\otimes} N \mathop \to \limits^{{\rho^M {\otimes} \rho^N}} FM {\otimes} FN \mathop \to \limits^{{F_2(M,N)}} F(M {\otimes} N), $ |
then $ (I, F_0) $ and $ (M {\otimes} N, \rho^{M {\otimes} N}) $ are all objects in $ \mathcal {C}^F $ if and only ifthe compatibility conditions Eqs (C1)–(C4) hold.
Proof. It is straightforward to check that Eqs (C1) and (C2) hold if and only if $ (M {\otimes} N, \rho^{M {\otimes} N}) \in \mathcal {C}^F $, Eqs (C3) and (C4) hold if and only if $ (I, F_0)\in \mathcal {C}^F $.
From now on, we always assume that the compatibility conditions Eqs (C1)–(C4) hold.
We suppose that there are natural transformations $ \vartheta: (\_ {\otimes} \_) {\otimes} \_ {\circ} F^{\times 3} \Rightarrow \_ {\otimes} (\_ {\otimes} \_): \mathcal {C}^{\times 3} \rightarrow \mathcal {C} $, and $ \iota: I {\otimes} F\_ \Rightarrow \_: \mathcal {C} \rightarrow \mathcal {C} $, $ \kappa:F\_ {\otimes} I \Rightarrow \_: \mathcal {C} \rightarrow \mathcal {C} $. From Lemma 2.1, for any objects $ (M, \rho^M), (N, \rho^N), (P, \rho^P) \in \mathcal {C}^F $, $ {\vartheta}, {\iota}, {\kappa} $ can induce the following natural transformations
$ AM,N,P=ϑ♯M,N,P,LM=ι♯M,RM=κ♯M. $
|
Conversely, if there are natural transformations $ A : (\_ {\otimes} \_) {\otimes} \_ \Rightarrow \_ {\otimes} (\_ {\otimes} \_): \mathcal {C}^{\times 3} \rightarrow \mathcal {C} $ and $ L: I {\otimes} \_ \Rightarrow id: \mathcal {C} \rightarrow \mathcal {C} $, $ R:\_ {\otimes} I \Rightarrow id: \mathcal {C} \rightarrow \mathcal {C} $, then from Lemma 2.1, for any $ X, Y, Z \in \mathcal {C} $, they can induce natural transformations
$ ϑX,Y,Z=A♭X,Y,Z,ιX=L♭X,κX=R♭X. $
|
Next, we will discuss when $ A $ is the associativity constraint and $ L, R $ are the unit constraints in $ \mathcal {C}^F $.
Lemma 3.2. $ A $, $ L $ and $ R $ are isomorphisms if and only if $ {\vartheta} $, $ {\iota} $ and $ {\kappa} $ are $ \ast $-invertible.
Proof. Straightforward from Proposition 2.2.
Lemma 3.3. $ A $ is $ F $-colinear if and only if $ {\vartheta} $ satisfies
![]() |
(3.1) |
for any $ X, Y, Z \in \mathcal {C} $.
Proof. $ \Leftarrow) $: Since the following diagram
![]() |
is commutative for any $ M, N, P \in \mathcal {C}^F $, $ A_{M, N, P} $ is $ F $-colinear.
$ \Rightarrow) $: Notice that $ A_{FX, FY, FZ} $ is $ F $-colinear for any $ X, Y, Z \in \mathcal {C} $, then it follows
$ F(εX⊗εY⊗εZ)∘FAFX,FY,FZ∘ρ(FX⊗FY)⊗FZ=F(εX⊗εY⊗εZ)∘ρFX⊗(FY⊗FZ)∘AFX,FY,FZ. $
|
After a direct computation, we obtain (3.1).
Lemma 3.4. $ A $ satisfies the Pentagon Axiom in $ \mathcal {C}^F $ if and only if $ {\vartheta} $ satisfies
$ (id⊗ϑX,Y,Z)∘ϑW,FX⊗FY,FZ∘(id⊗F2⊗id)∘(ϑFW,FFX,FFY⊗id)∘(δW⊗δ2X⊗δ2Y⊗δZ)=ϑW,X,Y⊗Z∘(id⊗id⊗F2)∘ϑFW⊗FX,FY,FZ∘(F2⊗id⊗id)∘(δW⊗δX⊗δY⊗δZ) $
|
(3.2) |
for any $ W, X, Y, Z \in \mathcal {C} $.
Proof. $ \Leftarrow) $: Since we have
$ (id⊗ϑN,P,Q)∘(id⊗ρN⊗ρP⊗ρQ)∘ϑM,N⊗P,Q∘(id⊗F2⊗id)∘(ρM⊗ρN⊗ρP⊗ρQ)∘(ϑM,N,P⊗id)∘(ρM⊗ρN⊗ρP⊗id)=(id⊗ϑN,P,Q)∘ϑM,FN⊗FP,FQ∘(id⊗F(ρN⊗ρP)⊗ρQ)∘(id⊗F2⊗id)∘(ϑFM,FN,FP⊗id)∘(FρM⊗FρN⊗FρP⊗ρQ)∘(ρM⊗ρN⊗ρP⊗id)=(id⊗ϑN,P,Q)∘ϑM,FN⊗FP,FQ∘(id⊗F2⊗id)∘(ϑFM,FFN,FFP⊗id)∘(δM⊗δ2N⊗δ2P⊗δQ)∘(ρM⊗ρN⊗ρP⊗ρQ)=ϑM,N,P⊗Q∘(id⊗id⊗F2)∘ϑFM⊗FN,FP,FQ∘(F2⊗id⊗id)∘(δM⊗δN⊗δP⊗δQ)∘(ρM⊗ρN⊗ρP⊗ρQ)=ϑM,N,P⊗Q∘(id⊗id⊗F2)∘(ρM⊗ρN⊗ρP⊗ρQ)∘ϑM⊗N,P,Q∘(F2⊗id⊗id)∘(ρM⊗ρN⊗ρP⊗ρQ) $
|
for any $ M, N, P, Q \in \mathcal {C}^F $, $ A $ satisfies the Pentagon Axiom.
$ \Rightarrow) $: For any $ W, X, Y, Z \in \mathcal {C} $, we have cofree $ F $-comodules $ FW, FX, FY, FZ $. Consider the following Pentagon Axiom:
$ AFW,FX,FY⊗FZ∘AFW⊗FX,FY,FZ=(id⊗AFX,FY,FZ)∘AFW,FX⊗FY,FZ∘(AFW,FX,FY⊗id). $
|
Applying $ {\varepsilon}_W {\otimes} {\varepsilon}_X {\otimes} {\varepsilon}_Y {\otimes} {\varepsilon}_Z $ to both sides of the above identity, we get Diagram (3.2).
Lemma 3.5. For any $ X \in \mathcal {C} $,
(1) $ L $ is $ F $-colinear if and only if $ {\iota} $ satisfies
![]() |
(3.3) |
(2) $ R $ is $ F $-colinear if and only if $ {\kappa} $ satisfies
![]() |
(3.4) |
Proof. We only prove (1).
$ \Leftarrow) $: From the following commutative diagram
![]() |
for any $ M \in \mathcal {C}^F $, $ L_{M} $ is $ F $-colinear.
$ \Rightarrow) $: Conversely, since $ FX $ is an $ F $-comodule and $ L_{FX} $ is $ F $-colinear for any $ X \in \mathcal {C} $, it is directly to get Diagram (3.3).
Lemma 3.6. $ A $, $ L $ and $ R $ satisfy the Triangle Axiom in $ \mathcal {C}^F $ if and only if $ {\vartheta} $, $ {\iota} $ and $ {\kappa} $ satisfy
![]() |
(3.5) |
for any $ X, Y, Z \in \mathcal {C} $.
Proof. $ \Leftarrow) $: For any $ M, N\in \mathcal {C}^F $, we compute
$ (idM⊗ιN)∘(idM⊗idI⊗ρN)∘ϑM,I,N∘(ρM⊗F0⊗ρN)=(idM⊗ιN)∘ϑM,I,FN∘(idFM⊗F0⊗δN)∘(ρM⊗idI⊗ρN)=(idM⊗εN)∘(κM⊗idFN)∘(ρM⊗idI⊗ρN)=(κM⊗idN)∘(ρM⊗idI⊗idN) $
|
thus the Triangle Axiom in $ \mathcal {C}^F $ holds.
$ \Rightarrow) $: Conversely, for any $ X, Y \in \mathcal {C} $, since we have
![]() |
it is a direct computation to get Diagram (3.5).
Definition 3.7. Let $ (\mathcal {C}, \otimes, I, a, l, r) $ be a monoidal category on which $ (F, {\delta}, {\varepsilon}) $ is a monad and $ (F, F_2, F_0) $ is a quasi-monoidal functor such that the compatible conditions Eqs (C1)–(C4) are satisfied. If there are $ \ast $-invertible natural transformations $ {\vartheta} $, $ {\iota} $ and $ {\kappa} $ satisfying (3.1)–(3.5), then we call $ (F, {\delta}, {\varepsilon}, F_2, F_0, {\vartheta}, {\iota}, {\kappa}) $ a quasi-monoidal comonad on $ \mathcal {C} $,
Then by Lemma 3.1–3.6, one gets the following result.
Theorem 3.8. Let $ (\mathcal {C}, \otimes, I, a, l, r) $ be a monoidal category on which $ (F, {\delta}, {\varepsilon}) $ is a monad and $ (F, F_2, F_0) $ is a quasi-monoidal functor such that the compatible conditions Eqs (C1)–(C4) is satisfied. Then there exist natural transformations $ {\vartheta} $, $ {\iota} $ and $ {\kappa} $ such that $ (F, {\delta}, {\varepsilon}, F_2, F_0, {\vartheta}, {\iota}, {\kappa}) $ is a quasi-monoidal comonad if and only if there are natural transformations $ A $, $ L $ and $ R $ such that $ (\mathcal {C}^F, {\otimes}, I, A, L, R) $ is a monoidal category.
Example 3.9. Let $ (\mathcal {C}, \otimes, I, a, l, r) $ be a monoidal category on which $ (F, {\delta}, {\varepsilon}) $ is a monad and $ (F, F_2, F_0) $ is a quasi-monoidal functor such that the compatible conditions Eqs (C1)–(C4) are satisfied. If we define
$ {\vartheta}_{X,Y,Z} = a^\flat_{X,Y,Z},\; \; {\iota}_{X} = l^\flat_{X},\; \; {\kappa}_{X,Y,Z} = r^\flat_{X,Y,Z} $ |
for any $ X, Y, Z \in \mathcal{C} $, then Eq (3.2) holds because of the Pentagon Axiom of $ a $; Eq (3.5) holds because of the Triangle Axiom of $ a, l, r $; Eqs (3.1), (3.3) and (3.4) hold if and only if $ (F, F_2, F_0) $ is a monoidal functor. That means, the quasi-monoidal comonad $ (F, {\delta}, {\varepsilon}, F_2, F_0, {\vartheta}, {\iota}, {\kappa}) $ is exactly a monoidal comonad.
Example 3.10. Recall from [9] or [10], we consider the following monoidal category $ \overline{\mathcal{H}}^{i, j}(Vec_k) $ for any $ i, j \in \mathbb{Z} $:
$ \bullet $ the objects of $ \overline{\mathcal{H}}^{i, j}(Vec_k) $ are pairs $ (X, {\alpha}_X) $, where $ X\in Vec_k $ and $ {\alpha}_X\in Aut_{k}(X) $;
$ \bullet $ the morphism $ f:(X, {\alpha}_X)\rightarrow (Y, {\alpha}_Y) $ in $ \overline{\mathcal{H}}^{i, j}(Vec_k) $ is a $ k $-linear map from $ X $ to $ Y $ such that $ {\alpha}_Y\circ f = f\circ {\alpha}_X $;
$ \bullet $ the monoidal structure is given by
$ (X,{\alpha}_X)\otimes(Y,{\alpha}_Y) = (X\otimes Y,{\alpha}_X\otimes{\alpha}_Y), $ |
and the unit is $ (k, id_{k}) $;
$ \bullet $ the associativity constraint $ a $, the unit constraints $ l $ and $ r $ are given by
$ aX,Y,Z:(x⊗y)⊗z↦αi+1X(x)⊗(y⊗α−j−1Z(z));lX(1k⊗x)=αj+1X(x),rX(x⊗1k)=αi+1X(x),∀X∈Veck. $
|
Now assume that $ (H, {\alpha}_H) $ is an object in $ \overline{\mathcal{H}}^{i, j}(Vec_k) $, $ m_H: H{\otimes} H \rightarrow H $ (with notation $ m_H(a {\otimes} b) = ab $), $ \eta_H: k\rightarrow H $ (with notation $ \eta_H(1_k) = 1_H $), and $ \Delta_H: H \rightarrow H{\otimes} H $ (with notation $ \Delta_H(h) = h_1 {\otimes} h_2 $), and $ \varepsilon_H: H\rightarrow k $ are all morphisms in $ \overline{\mathcal{H}}^{i, j}(Vec_k) $. Further, we write
$ \ddot{H} = \_{\otimes} H:\overline{\mathcal{H}}^{i,j}(Vec_k)\rightarrow \overline{\mathcal{H}}^{i,j}(Vec_k),\; \; \; \; (X, {\alpha}_X)\mapsto ( X {\otimes} H, {\alpha}_X {\otimes} {\alpha}_H) $ |
for the right tensor functor of $ H $.
If we define the following structures on $ \ddot{H} $:
$ \bullet $ $ \delta:\ddot{H}\rightarrow \ddot{H}\ddot{H} $ and $ \epsilon:\ddot{H}\rightarrow id_{\overline{\mathcal{H}}^{i, j}(Vec_k)} $ are defined by
$ δX:x⊗h↦(αX(x)⊗h1)⊗α−1H(h2),ϵX:x⊗h↦εH(h)α−1X(x); $
|
$ \bullet $ $ \ddot{H}_2:\ddot{H} {\otimes} \ddot{H} \rightarrow \ddot{H} {\otimes} $ and $ \ddot{H}_0: k\rightarrow \ddot{H}(k) $ are given by
$ ¨H2(X,Y):(x⊗a)⊗(y⊗b)↦(x⊗y)⊗αiH(a)αjH(b),¨H0(1k)=1k⊗1H, $
|
for any $ X, Y \in \overline{\mathcal{H}}^{i, j}(Vec_k) $. Then obviously $ \ddot{H} = (\ddot{H}, \delta, \epsilon) $ forms a comonad on $ \overline{\mathcal{H}}^{i, j}(Vec_k) $ if and only if $ (H, {\alpha}_H, \Delta_H, \varepsilon_H) $ is a Hom-coalgebra over $ k $, Eqs (C1)–(C4) hold if and only if $ m_H $ and $ \eta_H $ are all morphisms of Hom-coalgebras.
Suppose that there are $ {\alpha}_H $-invariant convolution invertible linear forms $ \omega\in (H {\otimes} H {\otimes} H)^\ast $ and $ p, q \in H^\ast $, then we can define the following $ \ast $-invertible natural transformations
$ ϑX,Y,Z:((x⊗a)⊗(y⊗b))⊗(z⊗c)↦ω(α2iH(a),αi+jH(b),αj−1H(c))(αiX(x)⊗(α−1Y(y)⊗α−j−2Z(z))),ιX:1k⊗(x⊗a)↦p(a)αjX(x),κX:(x⊗a)⊗1k↦q(a)αiX(x), $
|
where $ a, b, c \in H $, $ x \in X $, $ y \in Y $, $ z \in Z $ and $ X, Y, Z \in Vec_k $. Thus we immediately get that $ {\vartheta} $ satisfies Eq (3.1) if and only if $ \omega $ satisfies
$ ∑αH(a1)(b1c1)ω(a2,b2,c2)=∑ω(a1,b1,c1)(a2b2)αH(c2); $
|
(3.6) |
$ {\vartheta} $ satisfies Eq (3.2) if and only if $ \omega $ satisfies
$ ∑ω(αH(a1),αH(b1),c1d1)ω(a2b2,αH(c2),αH(d2))=∑ω(b1,c1,αH(d1))ω(αH(a1),α−1H(b21)α−1H(c21),αH(d2))ω(αH(a2),b22,c22); $
|
(3.7) |
$ {\iota} $ satisfies Eq (3.3) and $ {\kappa} $ satisfies Eq (3.4) if and only if $ p, q $ satisfy
$ ∑p(a1)1Ha2=αH(a1)p(a2),∑q(a1)a21H=αH(a1)q(a2); $
|
(3.8) |
$ {\vartheta}, {\iota} $ and $ {\kappa} $ satisfy Eq (3.5) if and only if $ \omega $, $ p $ and $ q $ satisfy
$ ω(a,1H,b)=q(a)p∗−1(b). $
|
(3.9) |
This means, $ \ddot{H} = (\ddot{H}, {\delta}, \epsilon, \ddot{H}_2, \ddot{H}_0, {\vartheta}, {\iota}, {\kappa}) $ forms a quasi-monoidal comonad on $ \overline{\mathcal{H}}^{i, j}(Vec_k) $ if and only if $ H = (H, {\alpha}_H, m_H, \eta_H, {\Delta}_H, {\varepsilon}_H, \omega, p, q) $ forms a Hom-coquasi-bialgebra over $ k $ (see [20] for the dual definition). Further, from Theorem 3.10, one get that $ Corep(H) = (\overline{\mathcal{H}}^{i, j}(Vec_k))^{\ddot{H}} $, the category of right $ H $-Hom-comodules, is a monoidal category and its associativity constraint, unit constraints are given as follows:
$ AM,N,P((m⊗n)⊗p)=∑ω(α2i(m1),αi+j(n1),αj−1(p1))αiM(m0)⊗(α−1N(n0)⊗α−j−2P(p0)),LM(1k⊗m)=∑p(m1)αjM(m0),RM(m⊗1k)=∑q(m1)αiM(m0), $
|
where $ m \in M $, $ n \in N $, $ p \in P $, $ M, N, P \in Corep(H) $.
Example 3.11. Under the consideration of Example 3.10, if all the Hom-structure maps $ {\alpha} $ are identity maps, then the Hom-coquasi-bialgebra is exactly the Majid algebra (also called a Majid algebra, see [13] for details) over $ k $.
Example 3.12. Let $ B = (B, \mu, 1_B, {\Delta}, {\varepsilon}) $ be a bialgebra over $ k $, $ {\alpha}_B:B\rightarrow B $ be an endo-isomrophism. Recall that a $ k $-linear form $ g\in B^\ast $ is called
(1) dual central if $ g(x_1)x_2 = x_1g(x_2) $ for any $ x \in B $;
(2) dual group-like if it is convolution invertible and satisfies $ g(xy) = g(x)g(y) $ for any $ x, y \in B $;
(3) $ {\alpha}_B $-invariant if $ g({\alpha}_B(x)) = g(x) $.
Now suppose that $ p, q \in B^\ast $ are all dual central dual group-like and $ {\alpha}_B $-invariant linear forms. Define a $ k $-linear form $ \omega:B {\otimes} B{\otimes} B\rightarrow k $ by
$ \omega(x,y,z) = p(x){\varepsilon}(y)q^{\ast-1}(z),\; \; \; \; {\rm{for\; any\; }}x,y,z \in B, $ |
define the new multiplication $ \mu^{{\alpha}_B} $ and comultiplication $ {\Delta}^{{\alpha}_B} $ by
$ μαB=αB∘μ,ΔαB=Δ∘αB. $
|
Then it is a direct calculation to check that $ {\alpha}_B, \omega, p, q $ satisfy Eqs.(3.6) - (3.9) (under $ \mu^{{\alpha}_B} $ and $ {\Delta}^{{\alpha}_B} $), hence $ B^{p, q}_{{\alpha}_B} = (B, {\alpha}_B, \mu^{{\alpha}_B}, 1_B, {\Delta}^{{\alpha}_B} $, $ {\varepsilon}, \omega, p, q) $ forms a nontrivial Hom-coquasi-bialgebra.
Recall that a braiding in a monoidal category $ (\mathcal {C}, \otimes, I, a, l, r) $ is a natural isomorphism $ {\tau} $: $ \otimes \Rightarrow \otimes^{op}: \mathcal {C} \times \mathcal {C}\rightarrow \mathcal {C} $ such that the following identities hold
$ aY,Z,X∘τX,Y⊗Z∘aX,Y,Z=(idY⊗τX,Z)∘aY,X,Z∘(τX,Y⊗idZ), $
|
( |
$ a−1Z,X,Y∘τX⊗Y,Z∘a−1X,Y,Z=(τX,Z⊗idY)∘a−1X,Z,Y∘(idX⊗τY,Z) $
|
( |
for any $ X, Y, Z \in \mathcal {C} $.
Now let $ F $ be a quasi-monoidal comonad on $ \mathcal {C} $. Suppose that there is a natural transformation $ \sigma $: $ \otimes{\circ}(F\times F) \Rightarrow \otimes^{op}: \mathcal {C}^{\times 2}\rightarrow \mathcal {C} $. From Lemma 2.1, for any objects $ M, N $ in $ \mathcal {C}^F $, $ \sigma $ can induce a natural transformation
$ {\tau}_{M,N} = \sigma^\sharp_{M,N}: M\otimes N \mathop \to \limits^{{ \rho^M {\otimes} \rho^N} } FM\otimes FN \mathop \to \limits^{{\sigma_{M,N}}} N\otimes M. $ |
Conversely, if there exists $ {\tau} : \otimes \Rightarrow \otimes^{op}: \mathcal {C} \times \mathcal {C}\rightarrow \mathcal {C} $, then from Lemma 2.1, for any $ X, Y \in \mathcal {C} $, $ {\tau} $ can induce the following
$ \sigma_{X,Y} = {\tau}^\flat_{X,Y}: FX\otimes FY\mathop \to \limits^{{{\tau}_{FX,FY}}} FY \otimes FX \mathop \to \limits^{{{\varepsilon}_{Y} \otimes {\varepsilon}_{X}}} Y\otimes X. $ |
Next we will discuss when $ {\tau} $ is a braiding in $ \mathcal {C}^F $.
Lemma 4.1. $ {\tau} $ is an isomorphism if and only if $ \sigma $ is $ \ast $-invertible.
Proof. Straightforward from Proposition 2.2.
Lemma 4.2. $ {\tau} $ is $ F $-colinear if and only if $ {\sigma} $ satisfies
![]() |
(4.1) |
for any $ X, Y \in \mathcal {C} $.
Proof. $ \Leftarrow) $: We compute
![]() |
for any $ M, N \in \mathcal {C}^F $. Hence $ {\tau}_{M, N} $ is $ F $-colinear.
$ \Rightarrow) $: Conversely, notice that $ {\tau}_{FX, FY} $ is $ F $-colinear for any $ X, Y \in \mathcal {C} $, we have
$ F(εY⊗εX)∘FτFX,FY∘ρFX⊗FY=F(εY⊗εX)∘ρFY⊗FX∘τFX,FY, $
|
which implies Diagram (4.1) holds.
Lemma 4.3. Diagram (B1) holds in $ \mathcal {C}^F $ if and only if $ \sigma $ satisfies
$ ϑY,Z,X∘σFX,FY⊗FZ∘(id⊗F2)∘ϑFFX,FFY,FFZ∘(δ2X⊗δ2Y⊗δ2Z)=(id⊗σX,Z)∘ϑY,FX,FZ∘(σFFX,FY⊗id)∘(δ2X⊗δY⊗δZ) $
|
(4.2) |
for any $ X, Y, Z \in \mathcal {C} $.
Proof. $ \Leftarrow) $: Take $ X = M $, $ Y = N $, $ Z = P $ for any $ F $-comodules $ M, N, P $. Multiplied by $ \rho^M {\otimes} \rho^N {\otimes} \rho^P $ right on both sides of Eq (4.2), we immediately get Diagram (B1).
$ \Rightarrow) $: Since Diagram (B1) is commutative for any $ FX, FY, FZ \in \mathcal {C} $, multiplied by $ {\varepsilon} {\otimes} {\varepsilon} {\otimes} {\varepsilon} $ left on both sides of the above equation, we get Eq (4.2).
Lemma 4.4. For any $ X, Y, Z \in \mathcal {C} $, Diagram (B2) holds in $ \mathcal {C}^F $ if and only if $ \sigma $ satisfies
$ ϑ∗−1Z,X,Y∘σFX⊗FY,FZ∘(F2⊗id))∘ϑ∗−1FFX,FFY,FFZ∘(δ2X⊗δ2Y⊗δ2Z)=(σX,Z⊗id)∘ϑ∗−1FX,FZ,Y∘(id⊗σFY,FFZ)∘(δX⊗δY⊗δ2Z), $
|
(4.3) |
where $ {\vartheta}^{\ast -1} $ means the $ \ast $-inverse of $ {\vartheta} $.
Proof. The proof is similar to Lemma 4.3.
Definition 4.5. Let $ (F, {\delta}, {\varepsilon}, F_2, F_0, {\vartheta}, {\iota}, {\kappa}) $ be a quasi-monoidal comonad on a monoidal category $ \mathcal {C} $. If there is a $ \ast $-invertible natural transformation $ \sigma \in Nat(F\otimes F, \otimes^{op}) $, satisfying Eqs (4.1)–(4.3) for any $ X, Y, Z \in \mathcal {C} $, then $ {\sigma} $ is called a coquasitriangular structure of $ F $, and $ (F, \sigma) $ is called a coquasitriangular quasi-monoidal comonad.
Combining Lemma 4.1–Definition 4.5, we obtain the following result.
Theorem 4.6. Let $ (F, {\delta}, {\varepsilon}, F_2, F_0, {\vartheta}, {\iota}, {\kappa}) $ be a quasi-monoidal comonad on a monoidal category $ \mathcal {C} $. Then $ \mathcal {C}^F $ is a braided monoidal category if and only if there exists a natural transformation$ \sigma: F{\otimes} F \rightarrow {\otimes}^{op} $ such that $ (F, \sigma) $ is a coquasitriangular quasi-monoidal comonad. Further, the braiding in $ \mathcal {C}^F $is given by $ {\tau} = \sigma^\sharp $.
Corollary 4.7. Let $ (F, {\sigma}) $ be a coquasitriangular quasi-monoidal comonad on a monoidal category $ \mathcal {C} $. Then for any $ X, Y, Z \in \mathcal {C} $, $ {\sigma} $ satisfies the following generalized Yang-Baxter equation:
$ (id⊗σX,Y)∘ϑZ,FX,FY∘(σFFX,FZ⊗id)∘ϑ∗−1F3X,FFZ,FFY∘(id⊗σF3Y,F3Z)∘ϑF4X,F4Y,F4Z∘(δ4X⊗δ4Y⊗δ4Z)=ϑZ,Y,X∘(σFY,FZ⊗id)∘ϑ∗−1FFY,FFZ,FX∘(id⊗σFFX,F3Z)∘ϑF3Y,F3X,F4Z∘(σF4X,F4Y⊗id)∘(δ4X⊗δ4Y⊗δ4Z). $
|
Proof. Straightforward.
Example 4.8. If $ F $ is a monoidal comonad on $ \mathcal{C} $, and $ {\sigma}:{\otimes} {\circ} F^{\times 2} \Rightarrow {\otimes}^{op} $ is a $ \ast $-invertible natural transformation satisfying Eqs (4.1)–(4.3), then $ (F, {\sigma}) $ is exactly a coquasitriangular monoidal comonad (see [9], Definition 4.12).
Example 4.9. With the notations in Example 3.10, if $ Q \in (H {\otimes} H)^\ast $ is $ {\alpha}_H $-invariant and convolution invertible, then we have the following $ \ast $-invertible natural transformation
$ {\sigma}_{X,Y}:\ddot{H}X {\otimes} \ddot{H}Y\rightarrow Y {\otimes} X,\; \; \; \\ (x {\otimes} a){\otimes} (y {\otimes} b)\mapsto Q({\alpha}_H^i(a),{\alpha}_H^j(b)){\alpha}_Y^{j-i-1}(y) {\otimes} {\alpha}_X^{i-j-1}(x), $ |
where $ x \in X $, $ y \in Y $ and $ X, Y \in \overline{\mathcal{H}}^{i, j}(Vec_k) $. Thus we immediately get that $ {\sigma} $ satisfies Eq (4.1) if and only if $ Q $ satisfies
$ \sum Q(a_{1}, b_{1})a_{2}b_{2} = \sum b_{1}a_{1} Q(a_{2}, b_{2}), $ |
$ {\sigma} $ satisfies Eqs (4.2) and (4.3) if and only if $ Q $ satisfies
$ ∑ω(b1,c1,a1)Q(αH(a21),b21c21)ω(a22,b22,c22)=∑Q(a1,c1)ω(b1,α−1H(a21),c2)Q(α−1H(a22),b2),∑ω∗−1(c1,a1,b1)Q(a21b21,αH(c21))ω∗−1(a22,b22,c22)=∑Q(a1,c1)ω∗−1(a2,α−1H(c21),b1)Q(b2,α−1H(c22)), $
|
where $ a, b, c \in H $. That is, $ (\ddot{H}, {\sigma}) $ forms a coquasitriangular quasi-monoidal comonad if and only if $ (H, Q) $ is a coquasitriangular Hom-coquasi-bialgebra. Further, from Theorem 4.6, one get that $ Corep(H) = (\overline{\mathcal{H}}^{i, j}(Vec_k))^{\ddot{H}} $ is a braided monoidal category.
Example 4.10. With the notations in Example 3.12, if $ p\in B^\ast $ is a dual central dual group-like $ {\alpha}_B $-invariant $ k $-linear form on a bialgebra $ B $, then we get a coquasi-bialgebra $ B_{{\alpha}_B}^{p, p} $. Now suppose that $ Q \in (B {\otimes} B)^\ast $ is the coquasitriangular structure over $ B $. If $ Q{\circ} ({\alpha}_B {\otimes} {\alpha}_B) = Q $, then after a straightforward compute we get that $ Q $ is also a coquasitriangular structure over the Hom-coquasi-bialgebra $ B_{{\alpha}_B}^{p, p} $.
Let $ F = (F, {\delta}, {\varepsilon}, F_2, F_0) $ be a quasi-monoidal comonad on a monoidal category $ (\mathcal {C}, \otimes, I, a, l, r) $.
Definition 5.1. A gauge transformation on $ F $ is a $ \ast $-invertible natural transformation $ \xi:F\otimes F \Rightarrow \otimes $.
Using a gauge transformation $ \xi $ on $ F $, we can build a new quasi-monoidal comonad $ F^\xi $ as follows.
Firstly, as a functor, $ F^\xi = F: \mathcal{C}\rightarrow \mathcal{C} $.
Secondly, the comonad structure of $ F^\xi $ is $ F^\xi = F = (F, {\delta}, {\varepsilon}) $.
Thirdly, the quasi-monoidal functor structure of $ F^\xi $ is given by:
$ \bullet $ for any $ X, Y \in \mathcal{C} $, $ F^\xi_2: F {\otimes} F \Rightarrow F {\otimes} $ is defined as follows
$ Fξ2(X,Y):FX⊗FYδ2X⊗δ2Y→F3X⊗F3Yξ→FFX⊗FFYF2→F(FX⊗FY)F(ξ∗−1X,Y)→F(X⊗Y) $
|
(5.1) |
where $ \xi^{\ast -1} $ means the $ \ast $-inverse of $ \xi $;
$ \bullet $ $ F^\xi_0 = F_0:FI \rightarrow I $.
Proposition 5.2. With the above notations, $ \delta $ and $ {\varepsilon} $ are both monoidal natural transformations
Proof. We only need to show the compatible conditions Eqs (C1)–(C4) hold.
To prove Eq (C1), we compute
![]() |
for any $ X, Y \in \mathcal{C} $. The rest are straightforward.
For any $ X, Y \in \mathcal{C} $, define the natural transformation $ {\vartheta}^\xi:(F {\otimes} F) {\otimes} F\Rightarrow \_ {\otimes} (\_{\otimes} \_) $ by
$ ϑξX,Y,Z=(id⊗ξ∗−1Y,Z)∘ξ∗−1X,FY⊗FZ∘(id⊗F2)∘ϑFX,FFY,FFZ∘(ξFFX⊗F3Y,F3Z)∘(F2⊗id)∘(ξF3X,F4Y⊗id)∘(δ3X⊗δ4Y⊗δ3Z), $
|
(5.2) |
and define the followings natural transformations:
$ ιξX:I⊗FXF0⊗δX→FI⊗FFXξ→I⊗FXιX→X, $
|
(5.3) |
and
$ κξX:FX⊗IδX⊗F0→FFX⊗FIξ→FX⊗IκX→X. $
|
(5.4) |
It is easy to get that $ {\vartheta}^\xi $, $ {\iota}^\xi $ and $ {\kappa}^\xi $ are all $ \ast $-invertible. Further, we have the following properties.
Lemma 5.3. With the above notations, $ {\vartheta}^\xi $ satisfies Eqs (3.1) and (3.2).
Proof. We only prove Eq (3.1). For any $ X, Y, Z \in \mathcal {C} $, we compute
$ F(ϑξX,Y,Z)∘Fξ2∘(Fξ2⊗id)∘(δX⊗δY⊗δZ)=F(id⊗ξ∗−1Y,Z)∘F(ξ∗−1X,FY⊗FZ)∘F(id⊗F2)∘F(ϑFX,FFY,FFZ)∘F(ξFFX⊗F3Y,F3Z)∘F(F2⊗id)∘F(ξF4X,F3Y⊗id)∘F(δ3X⊗δ4Y⊗δ3Z)∘F(ξ∗−1FX⊗FY,FZ)∘F2∘(δFX⊗FY⊗δFZ)∘ξF(FX⊗FY),FFZ∘(δFX⊗FY⊗δFZ)∘(F(ξ∗−1FX,FY)⊗id)∘(F2⊗id)∘(δFX⊗δFY⊗id)∘(ξFFX,FFY⊗id)∘(δFX⊗δFY⊗id)∘(δX⊗δY⊗δZ)=F(id⊗ξ∗−1Y,Z)∘F(ξ∗−1X,FY⊗FZ)∘F(id⊗F2)∘F(ϑFX,FFY,FFZ)∘F(ξFFX⊗F3Y,F3Z)∘F(F2⊗id)∘F(δFX⊗δ2FY⊗δ3Z)∘F(ξ∗−1FFX⊗FFY,FZ)∘F2∘(δFFX⊗FFY⊗δFZ)∘ξF(FFX⊗FFY),FFZ∘(δFFX⊗FFY⊗δFZ)∘(F2⊗id)∘(δFX⊗δFY⊗id)∘(ξFFX,FFY⊗id)∘(δ2X⊗δ2Y⊗δZ)=F(id⊗ξ∗−1Y,Z)∘F(ξ∗−1X,FY⊗FZ)∘F(id⊗F2)∘F(ϑFX,FFY,FFZ)∘F2∘(F2⊗id)∘(δ2FX⊗δ2FY⊗δ2FZ)∘ξF3X⊗F3Y,FFZ∘(F2⊗id)∘(δFX⊗δ2FY⊗δFZ)∘(ξFFX,FFY⊗id)∘(δ2X⊗δ2Y⊗δZ)=F(id⊗ξ∗−1Y,Z)∘F(ξ∗−1X,FY⊗FZ)∘F(id⊗F2)∘F2∘(id⊗F2)∘(δX⊗δ2Y⊗δ2Z)∘ϑFX,FY,FZ∘ξFFX⊗FFY,FFZ∘(F2⊗id)∘(ξF3X,F3Y⊗id)∘(δ3X⊗δ3Y⊗δ2Z)=F(ξ∗−1X,Y⊗Z)∘F(id⊗F(ξ∗−1Y,Z))∘F2∘(δX⊗δFY⊗FZ)∘(id⊗F2)∘(id⊗δY⊗δZ)∘(εFX⊗εFY⊗εFZ)∘ϑFFX,FFY,FFZ∘ξF3X⊗F3Y,F3Z∘(F2⊗id)∘(ξF4X,F4Y⊗id)∘(δ4X⊗δ4Y⊗δ3Z)=F(ξ∗−1X,Y⊗Z)∘F2∘(δX⊗δY⊗Z)∘ξFX,F(Y⊗Z)∘ξ∗−1FFX,FF(Y⊗Z)∘(FδX⊗FδY⊗Z)∘(id⊗FF(ξ∗−1Y,Z))∘(id⊗F(F2))∘(id⊗F(δY⊗δZ))∘(id⊗F2)∘ϑFFX,FFY,FFZ∘ξF3X,F3Y⊗F3Z∘(F(id⊗FFεFY)⊗FFFεFZ)∘(F2⊗id)∘(ξF4X,F5Y⊗id)∘(δ4X⊗δ5Y⊗δ4Z)=F(ξ∗−1X,Y⊗Z)∘F2∘(δX⊗δY⊗Z)∘ξFX,F(Y⊗Z)∘(δX⊗δY⊗Z)∘(F(id⊗ξ∗−1Y,Z))∘(id⊗F2(FY,FZ))∘(id⊗δY⊗δZ)∘(id⊗ξFX,FY)∘(id⊗ξ∗−1FFY,FFZ)∘(id⊗δFY⊗δFZ)∘ξ∗−1FX,FFY⊗FFZ∘(id⊗F2(FFY⊗FFZ))∘ϑFFX,F3Y,F3Z∘ξF3X,F4Y⊗F4Z∘(F2⊗id)∘(ξF4X,F5Y⊗id)∘(δ4X⊗δ5Y⊗δ4Z)=Fξ2(FX,FY⊗FZ)∘(id⊗Fξ2(FY,FZ))∘ϑξFX,FY,FZ∘(δX⊗δY⊗δZ). $
|
Thus the conclusion holds.
Lemma 5.4. With the above notations, $ {\iota}^\xi $ satisfies Eq (3.3) and $ {\kappa}^\xi $ satisfies Eq (3.4).
Proof. We only prove Eq (3.3). For any $ X \in \mathcal {C} $, we have
![]() |
which implies Eq (3.3).
Lemma 5.5. With the above notations, $ {\vartheta}^\xi $ and $ {\iota}^\xi $, $ {\kappa}^\xi $ satisfy Eq (3.5).
Proof. For any $ X, Y \in \mathcal {C} $, we obtain
$ (id⊗ιξY)∘(ϑξX,I,FY)∘(id⊗F0⊗δY)=(id⊗ιY)∘(id⊗ξI,FFY)∘(id⊗F0⊗δY)⊗(id⊗ξ∗−1I,FY)∘ξ∗−1X,FI,FFY∘(id⊗F2)∘ϑFX,FFI,F3Y∘ξFFX⊗F3I,F4Y∘(F2⊗id)∘(ξF3X,F4I⊗id)∘(δ3X⊗δ4I⊗δ3FY)⊗(δX⊗F0⊗δY)=ξ∗−1X,Y∘(id⊗FιY)∘(id⊗F2)∘ϑFX,FI,FFY∘ξFFX⊗FFI,F3Y∘(F2⊗id)∘(ξF3X,F3I⊗id)∘(δ3X⊗δ3I⊗δ2FY)∘(δX⊗F0⊗δY)=ξ∗−1X,Y∘(id⊗ιFY)∘ϑFX,I,FFY∘(id⊗F0⊗δY)∘ξFFX⊗I,FFY∘(F2⊗id)∘(ξF3X,FI⊗id)∘(δ3X⊗δI⊗δ2Y)∘(δX⊗F0⊗id)=(id⊗εY)∘ξ∗−1X,FY∘(κFX⊗id)∘ξFFX⊗I,FFY∘(F2⊗id)∘(δFX⊗F0⊗id)∘(ξFFX,I⊗id)∘(δ2X⊗F0⊗δ2Y)=(id⊗εY)∘ξ∗−1X,FY∘ξFX,FFY∘(κFFX⊗id)∘(δFX⊗id⊗id)∘(ξFFX,I⊗id)∘(δ2X⊗F0⊗δ2Y)=(id⊗εY)∘(κX⊗id)∘(ξFX,I⊗id)∘(δX⊗F0⊗id)=(id⊗εY)∘(κξX⊗id) $
|
hence Eq (3.5) holds.
Theorem 5.6. $ F^\xi = (F, {\delta}, {\varepsilon}, F^\xi_2, F_0, {\vartheta}^\xi, {\iota}^\xi, {\kappa}^\xi) $ is a quasi-monoidal comonad.
Remark 5.7. $ (\mathcal{C}^{F^\xi}, {\otimes}, I, A^\xi, L^\xi, R^\xi) $ is a monoidal category, where $ A^\xi = ({\vartheta}^\xi)^\sharp $, $ L^\xi = ({\iota}^\xi)^\sharp $, $ R^\xi = ({\kappa}^\xi)^\sharp $.
Now consider a coquasitriangular quasi-monoidal comonad $ (F, {\sigma}) $. For any gauge transformation $ \xi $ on $ F $, for any $ X, Y \in \mathcal{C} $, define
$ σξX,Y:FX⊗FYδ2⊗δ2→F3X⊗F3Yξ→FFX⊗FFYσ→FY⊗FXξ∗−1→Y⊗X. $
|
(5.5) |
Proposition 5.8. With the above notations, $ \sigma^\xi $ is a coquasitriangular structure of $ F^\xi $. Thus $ F^\xi $ is a coquasitriangular quasi-monoidal comonad. Hence $ \mathcal{C}^{F^\xi} $ is a braided monoidal category with the braiding $ {\tau}^\xi = (\sigma^\xi)^\sharp $.
Proof. Firstly, it is straightforward to get that $ \sigma^\xi $ is $ \ast $-invertible.
Secondly, to prove Eq (4.1), for any $ X, Y \in \mathcal{C} $, we compute
$ Fξ2∘σξFX,FY∘(δX⊗δY)=F(ξ∗−1Y,X)∘F2∘(δY⊗δX)∘ξFY,FX∘ξ∗−1FFY,FFX∘(δ2Y⊗δ2X)∘σFX,FY∘ξFFX,FFY∘(δ2X⊗δ2Y)=F(ξ∗−1Y,X)∘F2∘σFFX,FFY∘ξF3X,F3Y∘(δ3X⊗δ3Y)=F(ξ∗−1Y,X)∘F(σFX,FY)∘F(ξFFX,FFY)∘F(ξ∗−1F3X,F3Y)∘F(δ3X⊗δ3Y)∘F2∘(δX⊗δY)=F(ξ∗−1Y,X)∘F(σFX,FY)∘F(ξFFX,FFY)∘F(δ2X⊗δ2Y)∘F(ξ∗−1FX,FY)∘F2∘(δFX⊗δFY)∘ξFFX,FFY∘(δ2X⊗δ2Y)=F(σξY,X)∘Fξ2(FX,FY)∘(δX⊗δY). $
|
Thirdly, for Eq (4.2), we have
$ ϑξY,Z,X∘σξFX,FY⊗FZ∘(id⊗Fξ2)∘ϑξFFX,FFY,FFZ∘(δ2X⊗δ2Y⊗δ2Z)=(id⊗ξ∗−1Z,X)∘ξ∗−1Y,FZ⊗FX∘(id⊗F2)∘ϑFY,FFZ,FFX∘ξFFY⊗F3Z,F3X∘(F2⊗id)∘(ξF3Y⊗F4Z⊗id)∘(δ3Y⊗δ4Z⊗δ3X)∘ξ∗−1FY⊗FZ,FX∘σFFX,F(FY⊗FZ)∘ξF3X,FF(FY⊗FZ))∘(δ2FX⊗δ2FY⊗FZ)∘(id⊗F(ξ∗−1FY,FZ))∘(id⊗F2)∘(id⊗ξF3Y,F3Z)∘(id⊗δ2FY⊗δ2FZ)∘(id⊗ξ∗−1F2Y,F2Z)∘ξ∗−1FFX,F3Y⊗F3Z∘(id⊗F2)∘ϑF3X,F4Y,F4Z∘ξF4X⊗F5Y,F5Z∘(F2⊗id)∘(ξF5X,F6Y⊗id)∘(δ5X⊗δ6Y⊗δ5Z)=(id⊗ξ∗−1Z,X)∘ξ∗−1Y,FZ⊗FX∘(id⊗F2)∘ϑFY,FFZ,FFX∘σF3X,FFY⊗F3Z∘(id⊗F2)∘ϑF4X,F3Y,F4Z∘(δ2FFX⊗δ2FY⊗δ2FFZ)∘ξF3X⊗FFY,F3Z∘(F2⊗id)∘(ξF4X,F3Y⊗id)∘(δ4X⊗δ3Y⊗δ3Z)=(id⊗ξ∗−1Z,X)∘(id⊗σFX,FZ)∘ξ∗−1Y,FFX⊗FFZ∘(id⊗F2)∘ϑFY,F3X,F3Z∘(σF4X,FFY⊗id)∘(δ2FFX⊗δFY⊗δFFZ)∘ξF3X⊗FFY,F3Z∘(F2⊗id)∘(ξF4X,F3Y⊗id)∘(δ4X⊗δ3Y⊗δ3Z)=(id⊗ξ∗−1Z,X)∘(id⊗σFX,FZ)∘ξ∗−1Y,FFX⊗FFZ∘(id⊗F2)∘ϑFY,F3X,F3Z∘ξF2Y⊗F4X,F4Z∘(F(σF4X,FFY)⊗id)∘(F2⊗id)∘(δ4FX⊗δ3FY⊗δFY)∘(ξFFX,FFY⊗id)∘(id⊗δ2Z⊗δ2Z)=(id⊗σξX,Z)∘ϑξY,FX,FZ∘(σξFFX,FF⊗id)∘(δ2X⊗δY⊗δZ). $
|
At last, we can prove Eq (4.3) in a similar way. Thus the conclusion holds.
Now consider the corepresentations of $ F $ and $ F^\xi $.
Theorem 5.9. $ \mathcal{C}^F $ and $ \mathcal{C}^{F^\xi} $ are isomorphic as monoidal categories.Further, if $ F $ is a coquasitriangular quasi-monoidal comonad, then $ \mathcal{C}^F $ and $ \mathcal{C}^{F^\xi} $are braided isomorphic.
Proof. For any morphism $ f $ and objects $ M, N $ in $ \mathcal{C} $, the monoidal functor is defined as follows
$ \mathbb{E} = (\mathbb{E},\mathbb{E}^\xi_2,\mathbb{E}_0):(\mathcal{C}^F,{\otimes}, I,A,L,R)\rightarrow ( \mathcal{C}^{F^\xi},{\otimes},I,A^\xi,L^\xi,R^\xi ), $ |
where
$ \mathbb{E}(M): = M\; \; {\rm{as\; an}}\; F -{\rm{comodule}},\; \; \mathbb{E}(f): = f,\; \; \mathbb{E}_0 = id_I, $ |
and $ \mathbb{E}^\xi_2(M, N): \mathbb{E}(M) {\otimes} \mathbb{E}(N)\rightarrow \mathbb{E}(M {\otimes} N) $ is given by
$ Eξ2(M,N)=ξ♯:M⊗NρM⊗ρN→FM⊗FNξM,N→M⊗N. $
|
Obviously $ \mathbb{E} $ is well-defined.
Now we will check relation (2.1). Indeed, we have
$ Eξ2(M,N⊗P)∘(id⊗Eξ2(N,P))∘AξM,N,P=ξM,N⊗P∘(id⊗F2)∘(ρM⊗ρN⊗ρP)∘(id⊗ξN,P)∘(id⊗ρN⊗ρP)∘(id⊗ξ∗−1N,P)∘ξ∗−1M,FN⊗FP∘(id⊗F2)∘ϑFM,FFN,FFP∘ξFFM⊗F3N,F3P∘(F2⊗id)∘(ξF3M,F4N⊗id)∘(δ3M⊗δ4N⊗δ3P)∘(ρM⊗ρN⊗ρP)=ξM,N⊗P∘(ρM⊗F2)∘ξ∗−1M,FN⊗FP∘(id⊗F2)∘ϑFM,FFN,FFP∘ξFFM⊗F3N,F3P∘(F2⊗id)∘(ξF3M,F4N⊗id)∘(δ3M⊗δ4N⊗δ3P)∘(ρM⊗ρN⊗ρP)=ξM,N⊗P∘ξ∗−1FM,F(N⊗P)∘(δM⊗δN⊗P)∘(id⊗F2)∘ϑFM,FN,FP∘ξFFM⊗FFN,FFP∘(F2⊗id)∘(ξF3M,F3N⊗id)∘(δ3M⊗δ3N⊗δ2P)∘(ρM⊗ρN⊗ρP)=ϑM,N,P∘ξFM⊗FN,FP∘(F2⊗id)∘(ξF2M,F2N⊗id)∘(δ2M⊗δ2N⊗δP)∘(ρM⊗ρN⊗ρP)=E(AM,N,P)∘Eξ2(M⊗N,P)∘(Eξ2(M,N)⊗id), $
|
which implies Eq (2.1).
Further, we can obtain (2.2) and (2.3) by straightforward computation. Hence the conclusion holds.
Moreover, if $ {\sigma} $ is a coquasitriangular structure of $ F $, then from Theorem 5.6, $ (F^\xi, {\sigma}^\xi) $ is also a coquasitriangular quasi-monoidal comonad. Then we have
$ Eξ2(N,M)∘τξM,N=ξN,M∘(ρN⊗ρM)∘ξ∗−1N,M∘σFM,FN∘ξFFM,FFN∘(δ2M⊗δ2N)∘(ρM⊗ρN)=(εN⊗εM)∘σFM,FN∘ξFFM,FFN∘(δ2M⊗δ2N)∘(ρM⊗ρN)=σFM,FN∘(ρN⊗ρM)∘ξM,N∘(ρM⊗ρN)=E(τM,N)∘Eξ2(M,N), $
|
which implies $ (\mathbb{E}, \mathbb{E}^\xi_2, \mathbb{E}_0) $ is a braided monoidal functor.
Example 5.10. With the notations in Example 3.10, if there is a convolution invertible linear form $ \chi \in (H {\otimes} H)^\ast $ satisfying $ \chi {\circ} ({\alpha}_H {\otimes} {\alpha}_H) = \chi $, then we have the following $ \ast $-invertible natural transformation in $ \overline{\mathcal{H}}^{i, j}(Vec_k) $
$ \xi_{X,Y}: \ddot{H}X {\otimes} \ddot{H}Y\rightarrow X {\otimes} Y,\; \; \; \\ (x {\otimes} a) {\otimes} (y {\otimes} b) \mapsto \chi({\alpha}_H^i(a),{\alpha}_H^j(b)) {\alpha}_X^{-1}(x) {\otimes} {\alpha}_Y^{-1}(y), $ |
where $ a, b \in H $, $ x \in X $, $ y \in Y $ and $ X, Y \in \overline{\mathcal{H}}^{i, j}(Vec_k) $. It is not hard to check that $ \ddot{H}_2^\xi $, $ {\vartheta}^\xi $, $ {\iota}^\xi $ and $ {\kappa}^\xi $ in Eqs (5.1)–(5.4) are deduced from the following
$ mχ(a⊗b)=∑χ∗−1(a1,b1)α−2H(a21)α−2H(b21)χ(a22,b22), $
|
where $ \chi^{\ast-1} $ means the convolution inverse of $ \chi $, and
$ ωχ(a,b,c)=∑χ∗−1(b11,c11)χ∗−1(αH(a11),α−1H(b121)c12)ω(a12,α−1H(b122),c21)χ(a21b21,αH(c22))χ(a22,b22),pχ(a)=∑p(a1)χ(1H,a2),qχ(a)=∑q(a1)χ(a2,1H), $
|
respectively. Thus from Example 3.10 and Theorem 5.6, $ H^\chi = (H, {\alpha}_H, m^\chi, 1_H, {\Delta}, {\varepsilon}, \omega^\chi, p^\chi, q^\chi) $ is also a Hom-coquasi-bialgebra.
Example 5.11. With the notations in Example 3.12, note that the $ B_{{\alpha}_B} = (B, {\alpha}_B, {\alpha}_B {\circ}{\mu}, 1_B, {\Delta} {\circ} {\alpha}_B, {\varepsilon}) $ is a Hom-bialgebra, and it can be seen as a Hom-coquasi-bialgebra $ B_{{\alpha}_B} = (B, {\alpha}_B, {\alpha}_B {\circ}\mu, 1_H, {\Delta} {\circ} {\alpha}_B, {\varepsilon}, {\varepsilon}{\otimes}{\varepsilon}{\otimes}{\varepsilon}, {\varepsilon}, {\varepsilon}) $. If there are $ {\alpha}_B $-invariant and dual central dual group-like $ k $-linear forms $ p, q \in B^\ast $, then we have the following gauge transformation $ \chi\in (B {\otimes} B)^\ast $ by
$ χ(a,b)=q∗−1(a)p(b),where a,b∈B. $
|
Obviously $ B_{{\alpha}_B}^\chi = B_{{\alpha}_B}^{p, q} $.
The work was partially supported by the National Natural Science Foundation of China (No. 11801304, 11871301), and the Taishan Scholar Project of Shandong Province (No. tsqn202103060).
The authors declare there is no conflict of interest.
[1] |
Zentner GE, Henikoff S (2013) Regulation of nucleosome dynamics by histone modifications. Nat Struct Mol Biol 20: 259-266. doi: 10.1038/nsmb.2470
![]() |
[2] |
Ho JWK, Jung YL, Liu T, et al. (2014) Comparative analysis of metazoan chromatin organization. Nature 512: 449-452. doi: 10.1038/nature13415
![]() |
[3] |
Amin V, Harris RA, Onuchic V, et al. (2015) Epigenomic footprints across 111 reference epigenomes reveal tissue-specific epigenetic regulation of lincRNAs. Nat Commun 6: 6370. doi: 10.1038/ncomms7370
![]() |
[4] |
Roadmap Epigenomics Consortium, Kundaje A, Meuleman W, Ernst J, et al. (2015) Integrative analysis of 111 reference human epigenomes. Nature 518: 317-330. Available from: http://www.nature.com/nature/journal/v518/n7539/abs/nature14248.html. doi: 10.1038/nature14248
![]() |
[5] |
Cantone I, Fisher AG (2013) Epigenetic programming and reprogramming during development. Nat Struct Mol Biol 20: 282-289. doi: 10.1038/nsmb.2489
![]() |
[6] |
Zhu J, Adli M, Zou JY, et al. (2013) Genome-wide chromatin state transitions associated with developmental and environmental cues. Cell 152: 642-654. doi: 10.1016/j.cell.2012.12.033
![]() |
[7] | Chen T, Dent SYR (2014) Chromatin modifiers and remodellers: regulators of cellular differentiation. Nat Rev Genet 15: 93-106. |
[8] |
Lieberman-Aiden E, van Berkum NL, Williams L, et al. (2009) Comprehensive mapping of longrange interactions reveals folding principles of the human genome. Science (New York NY) 326: 289-293. doi: 10.1126/science.1181369
![]() |
[9] |
Rao SSP, Huntley MH, Durand NC, et al. (2014) A 3d map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell 159: 1665-1680. doi: 10.1016/j.cell.2014.11.021
![]() |
[10] |
Dixon JR, Selvaraj S, Yue F, et al. (2012) Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature 485:376-380. doi: 10.1038/nature11082
![]() |
[11] |
Sexton T, Yaffe E, Kenigsberg E, et al. (2012) Three-dimensional folding and functional organization principles of the drosophila genome. Cell 148: 458-472. doi: 10.1016/j.cell.2012.01.010
![]() |
[12] |
Nora EP, Lajoie BR, Schulz EG, et al. (2012) Spatial partitioning of the regulatory landscape of the x-inactivation centre. Nature 485: 381-385. doi: 10.1038/nature11049
![]() |
[13] |
Palstra RJ, Tolhuis B, Splinter E, et al. (2003) The beta-globin nuclear compartment in development and erythroid differentiation. Nat Genet 35: 190-194. doi: 10.1038/ng1244
![]() |
[14] |
Zhang Y, Wong CH, Birnbaum RY, et al. (2013) Chromatin connectivity maps reveal dynamic promoter-enhancer long-range associations. Nature 504: 306-310. doi: 10.1038/nature12716
![]() |
[15] |
Le Dily F, BaùD, Pohl A, et al. (2014) Distinct structural transitions of chromatin topological domains correlate with coordinated hormone-induced gene regulation. Genes Dev 28: 2151-2162. doi: 10.1101/gad.241422.114
![]() |
[16] |
Li G, Ruan X, Auerbach RK, et al. (2012) Extensive promoter-centered chromatin interactions provide a topological basis for transcription regulation. Cell 148: 84-98. Available from: http: //www.cell.com/article/S0092867411015170/abstract. doi: 10.1016/j.cell.2011.12.014
![]() |
[17] |
Filion GJ, van Bemmel JG, Braunschweig U, et al. (2010) Systematic protein location mapping reveals five principal chromatin types in drosophila cells. Cell 143: 212-224. doi: 10.1016/j.cell.2010.09.009
![]() |
[18] |
Ernst J, Kheradpour P, Mikkelsen TS, et al. (2011) Mapping and analysis of chromatin state dynamics in nine human cell types. Nature 473: 43-49. doi: 10.1038/nature09906
![]() |
[19] |
Boulé JB, Mozziconacci J, Lavelle C (2015) The polymorphisms of the chromatin fiber. J Phys Condens Matter 27: 033101. Available from: http://iopscience.iop.org/0953-8984/27/3/033101. doi: 10.1088/0953-8984/27/3/033101
![]() |
[20] | Barbieri M, Fraser J, Lavitas LM, et al. (2013) A polymer model explains the complexity of largescale chromatin folding. Nucleus Austin Tex 4: 267-273. |
[21] |
Nicodemi M, Pombo A (2014) Models of chromosome structure. Curr Opin Cell Biol 28: 90-95. doi: 10.1016/j.ceb.2014.04.004
![]() |
[22] |
Giorgetti L, Galupa R, Nora EP, et al. (2014) Predictive polymer modeling reveals coupled fluctuations in chromosome conformation and transcription. Cell 157: 950-963. doi: 10.1016/j.cell.2014.03.025
![]() |
[23] |
Jost D, Carrivain P, Cavalli G, et al. (2014) Modeling epigenome folding: formation and dynamics of topologically associated chromatin domains. Nucleic Acids Res 42: 9553-9561. doi: 10.1093/nar/gku698
![]() |
[24] |
Caré BR, Carrivain P, Forné T, et al. (2014) Finite-size conformational transitions: A unifying concept underlying chromosome dynamics. Commun Theor Phys 62: 607. Available from: http://iopscience.iop.org/0253-6102/62/4/18. doi: 10.1088/0253-6102/62/4/18
![]() |
[25] | Gennes PG (1979) Scaling Concepts in Polymer Physics, Cornell University Press, Ithaca, NY. |
[26] |
Imbert JB, Lesne A, Victor JM (1997) Distribution of the order parameter of the coil-globule transition. Phys Rev E 56: 5630-5647. Available from: http://link.aps.org/doi/10.1103/PhysRevE.56.5630. doi: 10.1103/PhysRevE.56.5630
![]() |
[27] |
Ponmurugan M, Narasimhan SL, Krishna PSR, et al. (2007) Coil-globule transition of a single short polymer chain: an exact enumeration study. J Chem Phys 126: 144906. doi: 10.1063/1.2719195
![]() |
[28] |
Baumg C, Srtner A (1980) Statics and dynamics of the freely jointed polymer chain with lennardb jones interaction. J Chem Phys 72: 871-879. Available from: http://scitation.aip.org/content/aip/journal/jcp/72/2/10.1063/1.439242. doi: 10.1063/1.439242
![]() |
[29] |
Hsu HP (2014) Monte carlo simulations of lattice models for single polymer systems. J Chem Phys 141: 164903. Available from: http://scitation.aip.org/content/aip/journal/jcp/141/16/10.1063/1.4899258. doi: 10.1063/1.4899258
![]() |
[30] |
Yoshikawa K, Matsuzawa Y (1995) Discrete phase transition of giant DNA dynamics of globule formation from a single molecular chain. Physica D: Nonlinear Phenomena 84: 220-227. Available from: http://www.sciencedirect.com/science/article/pii/0167278995000205. doi: 10.1016/0167-2789(95)00020-5
![]() |
[31] | Caré BR (2013) Cgmdode : Coarse-grained macromolecular dynamics with open dynamics engine. Available from: https://bitbucket.org/bcare/cgmdode-hg. |
[32] |
Bussi G, Parrinello M (2008) Stochastic thermostats: comparison of local and global schemes. Comput Phys Commun 179: 26-29. Available from: http://www.sciencedirect.com/ science/article/pii/S0010465508000106. doi: 10.1016/j.cpc.2008.01.006
![]() |
[33] | Carrivain P, Barbi M, Victor JM (2014) In silico single-molecule manipulation of DNA with rigid body dynamics, PLoS Comput Biol 10: e1003456. http://www.ncbi.nlm.nih.gov/pmc/ articles/PMC3930497/. |
[34] |
Cortini R, Caré BR, Victor JM, et al. (2015) Theory and simulations of toroidal and rod-like structures in single-molecule DNA condensation. J Chem Phys 142: 105102. Available from: http://scitation.aip.org/content/aip/journal/jcp/142/10/10.1063/1.4914513. doi: 10.1063/1.4914513
![]() |
[35] | Maeshima K, Imai R, Tamura S, et al. (2014) Chromatin as dynamic 10-nm fibers. Chromosoma 123: (2014), 225-237. |
[36] |
Fierz B (2014) Synthetic chromatin approaches to probe the writing and erasing of histone modifications. Chem Med Chem 9: 495-504. doi: 10.1002/cmdc.201300487
![]() |
[37] |
Dodd IB, Sneppen K (2011) Barriers and silencers: a theoretical toolkit for control and containment of nucleosome-based epigenetic states. J Mol Biol 414: 624-637. doi: 10.1016/j.jmb.2011.10.027
![]() |
[38] |
Dayarian A, Sengupta AM (2013) Titration and hysteresis in epigenetic chromatin silencing. Phys Biol 10: 036005. doi: 10.1088/1478-3975/10/3/036005
![]() |
[39] |
Nagano T, Lubling Y, Stevens TJ, et al. (2013) Single-cell hi-c reveals cell-to-cell variability in chromosome structure. Nature 502: 59-64. Available from: http://www.nature.com/nature/journal/v502/n7469/full/nature12593.html. doi: 10.1038/nature12593
![]() |