This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Dualizing sup-preserving endomaps of a complete lattice

Luigi Santocanale LIS, CNRS UMR 7020, Aix-Marseille Université, France [email protected]
Abstract

It is argued in [6] that the quantale [L,L][L,L]_{\vee} of sup-preserving endomaps of a complete lattice LL is a Girard quantale exactly when LL is completely distributive. We have argued in [17] that this Girard quantale structure arises from the dual quantale of inf-preserving endomaps of LL via Raney’s transforms and extends to a Girard quantaloid structure on the full subcategory of SLatt (the category of complete lattices and sup-preserving maps) whose objects are the completely distributive lattices.

It is the goal of this talk to illustrate further this connection between the quantale structure, Raney’s transforms, and complete distributivity. Raney’s transforms are indeed 𝗆𝗂𝗑\mathsf{mix} maps in the isomix category SLatt and most of the theory can be developed relying on naturality of these maps. We complete then the remarks on cyclic elements of [L,L][L,L]_{\vee} developed in [17] by investigating its dualizing elements. We argue that if [L,L][L,L]_{\vee} has the structure a Frobenius quantale, that is, if it has a dualizing element, not necessarily a cyclic one, then LL is once more completely distributive. It follows then from a general statement on involutive residuated lattices that there is a bijection between dualizing elements of [L,L][L,L]_{\vee} and automorphisms of LL. Finally, we also argue that if LL is finite and [L,L][L,L]_{\vee} is autodual, then LL is distributive.

1 Lattice structure of the homsets in SLatt

The homset [X,Y][X,Y]_{\vee} in SLatt.

The category SLatt of complete lattices and sup-preserving functions is a well-known \ast-autonomous category [2, 10, 9, 6]. For complete lattices X,YX,Y, we denote by [X,Y][X,Y]_{\vee} the homset in this category. The two-element Boolean algebra 𝟐\mathbf{2} is a dualizing element and [X,𝟐][X,\mathbf{2}]_{\vee} is, as a lattice, isomorphic to the dual lattice XopX^{op}. More generally, the functor ()=[,𝟐](\,\cdot\,)^{\ast}=[\,\cdot\,,\mathbf{2}]_{\vee} is naturally isomorphic to the functor ()op(\,\cdot\,)^{op} where, for f:YXf:Y\xrightarrow{\;\;\;\;\;}X, fop:XopYopf^{op}:X^{op}\xrightarrow{\;\;\;\;\;}Y^{op} is the right adjoint of ff, noted here by ρ(f)\rho(f) (the left adjoint of an inf-preserving function g:YXg:Y\xrightarrow{\;\;\;\;\;}X shall be denoted by (g):XY\ell(g):X\xrightarrow{\;\;\;\;\;}Y).

Let us describe the internal structure of the homset [X,Y][X,Y]_{\vee} as a complete lattice. For xXx\in X and yYy\in Y, we define the following elements of [X,Y][X,Y]_{\vee}:

cy(t)\displaystyle c_{y}(t) :={y,t,,t=,\displaystyle:=\begin{cases}y\,,&t\neq\bot\,,\\ \bot\,,&t=\bot\,,\end{cases} ax(t)\displaystyle a_{x}(t) :={,tx,,tx,\displaystyle:=\begin{cases}\top\,,&t\not\leq x\,,\\ \bot\,,&t\leq x\,,\end{cases}
(y¯x)(t)\displaystyle(y\overline{\otimes}x)(t) :={,tx,y,<tx,,t=,\displaystyle:=\begin{cases}\top\,,&t\not\leq x\,,\\ y\,,&\bot<t\leq x\,,\\ \bot\,,&t=\bot\,,\end{cases} ey,x(t)\displaystyle e_{y,x}(t) :={y,tx,,tx.\displaystyle:=\begin{cases}y\,,&t\not\leq x\,,\\ \bot\,,&t\leq x\,.\end{cases}
Lemma 1.1.

For each f[X,Y]f\in[X,Y]_{\vee}, xXx\in X, and yYy\in Y, f(x)yf(x)\leq y if and only if fy¯xf\leq y\overline{\otimes}x. Consequently, for each f[X,Y]f\in[X,Y]_{\vee},

f\displaystyle f ={y¯xf(x)y}=xXf(x)¯x,\displaystyle=\bigwedge\{\,y\overline{\otimes}x\mid f(x)\leq y\,\}=\bigwedge_{x\in X}f(x)\overline{\otimes}x\,,

and the sup-preserving functions of the form y¯xy\overline{\otimes}x generate [X,Y][X,Y]_{\vee} under arbitrary meets.

The tensor notation arises from the canonical isomorphism [X,Y](YopX)op[X,Y]_{\vee}\simeq(Y^{op}\otimes X)^{op} of \ast-autonomous categories. That is, [X,Y][X,Y]_{\vee} is dual to the tensor product YopXY^{op}\otimes X and the functions y¯xy\overline{\otimes}x correspond to elementary tensors of YopXY^{op}\otimes X. For f[X,Y]f\in[X,Y]_{\vee}, g[Y,Z]g\in[Y,Z]_{\vee}, and h[X,Z]h\in[X,Z]_{\vee}, let us recall that there exists uniquely determined maps g\h[X,Y]g\backslash h\in[X,Y]_{\vee} and h/f[Y,Z]h/f\in[Y,Z]_{\vee} satisfying

gf\displaystyle g\circ f hifffg\hiffgh/f.\displaystyle\leq h\;\;\text{iff}\;\;f\leq g\backslash h\;\;\text{iff}\;\;g\leq h/f\,.

The binary operations \\backslash and // are known under several names: they yield left and right Kan extensions and (often when X=Y=ZX=Y=Z) they are named residuals or division operations [7] or right and left implication [15, 6]. With the division operations at hand, let us list some elementary relations between the functions previously defined:

Lemma 1.2.

The following relations hold: (i) y¯x=cyax=cy/cx=ay\axy\overline{\otimes}x=c_{y}\vee a_{x}=c_{y}/c_{x}=a_{y}\backslash a_{x}, (ii) ey,x=cyax=cyaxe_{y,x}=c_{y}\land a_{x}=c_{y}\circ a_{x}, (iii) cy=y¯=cyac_{y}=y\overline{\otimes}\top=c_{y}\circ a_{\bot}, (iv) ax=¯x=caxa_{x}=\bot\overline{\otimes}x=c_{\top}\circ a_{x}.

Proof.

The relations y¯x=cyaxy\overline{\otimes}x=c_{y}\vee a_{x} and ey,x=cyaxe_{y,x}=c_{y}\land a_{x} are well known, see e.g. [20], and (iii)(iii) and (iv)(iv) are immediate consequences of these relations.

Let us focus on y¯x=cy/cx=ay\axy\overline{\otimes}x=c_{y}/c_{x}=a_{y}\backslash a_{x} and notice that, in order to make sense of these relations, we need to assume cx:XXc_{x}:X^{\prime}\xrightarrow{\;\;\;\;\;}X, cy:XYc_{y}:X^{\prime}\xrightarrow{\;\;\;\;\;}Y, ax:XYa_{x}:X\xrightarrow{\;\;\;\;\;}Y^{\prime}, and ay:YYa_{y}:Y\xrightarrow{\;\;\;\;\;}Y^{\prime}. Observe that f(x)yf(x)\leq y if and only if fcxcyf\circ c_{x}\leq c_{y} and therefore, in view of Lemma 1.1 and of the definion of //, the relation y¯x=cy/cxy\overline{\otimes}x=c_{y}/c_{x}. Next, the condition fay\axf\leq a_{y}\backslash a_{x} amounts to ayfaxa_{y}\circ f\leq a_{x}, that is, for all tXt\in X, if txt\leq x then f(t)yf(t)\leq y. Clearly this condition is equivalent to f(x)yf(x)\leq y, and therefore to fy¯xf\leq y\overline{\otimes}x. Finally, the relation ey,x=cyaxe_{y,x}=c_{y}\circ a_{x} is directly verified. However, let us observe the abuse of notation, since for the maps axa_{x} and cyc_{y} to be composable we need to assume either X=YX=Y or ax:X𝟐a_{x}:X\xrightarrow{\;\;\;\;\;}\mathbf{2} and cy:𝟐Yc_{y}:\mathbf{2}\xrightarrow{\;\;\;\;\;}Y. ∎

Inf-preserving functions as tensor product.

Let [X,Y][X,Y]_{\land} denote the poset of inf-preserving functions from XX to YY, with the pointwise ordering. Observe that, as a set, [X,Y][X,Y]_{\land} equals [Xop,Yop][X^{op},Y^{op}]_{\vee}. Yet, as a poset or a lattice, the equality [X,Y]=[Xop,Yop]op[X,Y]_{\land}=[X^{op},Y^{op}]_{\vee}^{op} is the correct one. As a matter of fact, we have f[X,Y]gf\leq_{[X,Y]_{\land}}g iff f(x)Yg(x)f(x)\leq_{Y}g(x), all xXx\in X, iff g(x)Yopf(x)g(x)\leq_{Y^{op}}f(x), all xXx\in X, iff g[Xop,Yop]fg\leq_{[X^{op},Y^{op}]_{\vee}}f. Using standard isomorphisms of \ast-autonomous categories, we have

[X,Y]\displaystyle[X,Y]_{\land} =[Xop,Yop]opYXop.\displaystyle=[X^{op},Y^{op}]_{\vee}^{op}\simeq Y\otimes X^{op}\,.

That is, the set of inf-preserving functions from XX to YY can be taken as a concrete realization of the tensor product YXopY\otimes X^{op}. This should not come as a surprise, since it is well-known that the set of Galois connections from XX to YY—that is, pairs of functions (f:XY,g:YX)(f:X\xrightarrow{\;\;\;\;\;}Y,g:Y\xrightarrow{\;\;\;\;\;}X) such that yf(x)y\leq f(x) iff xg(y)x\leq g(y)—realizes the tensor product YXY\otimes X in SLatt, see e.g. [19, 13] or [6, §2.1.2]. Such a pair of functions is uniquely determined by its first element, which is an inf-preserving functions from XopX^{op} to YY.

Notice now that

[X,Y]op\displaystyle[X,Y]_{\vee}^{op} YopXXYop[Y,X],\displaystyle\simeq Y^{op}\otimes X\simeq X\otimes Y^{op}\simeq[Y,X]_{\land}\,, (1)

from which we derive the following principle:

Fact 1.3.

There is a bijection beweeen sup-preserving functions from [X,Y]op[X,Y]_{\vee}^{op} to [X,Y][X,Y]_{\vee} and sup-preserving functions from [Y,X][Y,X]_{\land} to [X,Y][X,Y]_{\vee}.

The map yielding the isomorphism in equation (1) is ρ\rho, the operation of taking the right adjoint. The bijection stated in Fact 1.3 is therefore obtained by precomposing with ρ\rho.

We exploit now the work done for [X,Y][X,Y]_{\vee} to recap the structure of [X,Y][X,Y]_{\land} as a tensor product. Consider the maps

γy(t)\displaystyle\gamma_{y}(t) :={,t=,y,otherwise,\displaystyle:=\begin{cases}\top\,,&t=\top\,,\\ y\,,&\text{otherwise}\,,\end{cases} αx(t)\displaystyle\alpha_{x}(t) :={,xt,,otherwise,\displaystyle:=\begin{cases}\top\,,&x\leq t,\\ \bot\,,&\text{otherwise}\,,\end{cases}
y¯x(t)\displaystyle y\underline{\otimes}x(t) ={,t=,y,xt,,otherwise.\displaystyle=\begin{cases}\top\,,&t=\top\,,\\ y\,,&x\leq t\,,\\ \bot\,,&\text{otherwise}\,.\end{cases}

By dualizing Lemma 1.1, we observe that the relation y¯x=γyαxy\underline{\otimes}x=\gamma_{y}\land\alpha_{x} holds, the maps y¯xy\underline{\otimes}x realize the elementary tensors of the (abstract) tensor product YXopY\otimes X^{op}, [X,Y][X,Y]_{\land} is join-generated by these maps, and every g[X,Y]g\in[X,Y]_{\land} can be canonically written as g=xXg(x)¯xg=\bigvee_{x\in X}g(x)\underline{\otimes}x.

Recall that a bimorphism ψ:Y×XopZ\psi:Y\times X^{op}\xrightarrow{\;\;\;\;\;}Z is a function that is sup-preserving in each variable, separately. This in particular means that infs in XX are transformed into sups in ZZ. The universal property of [X,Y][X,Y]_{\land} as a tensor product can be therefore stated as follows:

Fact 1.4.

Given a bimorphism ψ:Y×XopZ\psi:Y\times X^{op}\xrightarrow{\;\;\;\;\;}Z, there exists a unique sup-preserving functions ψ~:[X,Y]Z\tilde{\psi}:[X,Y]_{\land}\xrightarrow{\;\;\;\;\;}Z such that ψ~(y¯x)=ψ(y,x)\tilde{\psi}(y\underline{\otimes}x)=\psi(y,x). For g[X,Y]g\in[X,Y]_{\land}, ψ~(g)\tilde{\psi}(g) is defined by

ψ~(g)\displaystyle\tilde{\psi}(g) :=xLψ(g(x),x).\displaystyle:=\bigvee_{x\in L}\psi(g(x),x)\,.

2 Raney’s transforms

For g[X,Y]g\in[X,Y]_{\land} and f[X,Y]f\in[X,Y]_{\vee}, define

g(x)\displaystyle{g}{}^{\vee}(x) :=xtg(t),\displaystyle:=\bigvee_{x\not\leq t}g(t)\,, f(x)\displaystyle{f}{}^{\land}(x) :=txf(t).\displaystyle:=\bigwedge_{t\not\leq x}f(t)\,.

It is easily seen that g{g}{}^{\vee} has a right adjoint, so g[X,Y]{g}{}^{\vee}\in[X,Y]_{\vee}, and that f{f}{}^{\land} has a left adjoint, so f{f}{}^{\land} belongs to [X,Y][X,Y]_{\land}. We call the operations (){(\,\cdot\,)}{}^{\vee} and (){(\,\cdot\,)}{}^{\land} the Raney’s transforms, even if Raney defined these transforms on Galois connections. (In [17] we explicitly related these maps to Raney’s original way of defining them). Notice that gf{g}{}^{\vee}\leq f if and only if gfg\leq{f}{}^{\land}, so (){(\,\cdot\,)}{}^{\land} is right adjoint to (){(\,\cdot\,)}{}^{\vee}. Raney’s transforms have been the key ingredient allowing us to prove in [18] that [C,C][C,C]_{\vee} is a Girard quantale if CC is a complete chain and, lately in [17], that the full-subcategory of SLatt whose objects are the completely distributive lattices is a Girard quantaloid.

Consider the bimorphism e:Y×Xop[X,Y]e:Y\times X^{op}\xrightarrow{\;\;\;\;\;}[X,Y]_{\vee} sending y,xy,x to ey,x=cyax[X,Y]e_{y,x}=c_{y}\circ a_{x}\in[X,Y]_{\vee} and its extension

e~(f)\displaystyle\tilde{e}(f) ={cf(t)attX}.\displaystyle=\bigvee\{\,c_{f(t)}\circ a_{t}\mid t\in X\,\}\,.

By evaluating e~(f)\tilde{e}(f) at xLx\in L, we obtain

e~(f)(x)\displaystyle\tilde{e}(f)(x) ={(cf(t)at)(x)tL}=xtf(t)=f(x),\displaystyle=\bigvee\{\,(c_{f(t)}\circ a_{t})(x)\mid t\in L\,\}=\bigvee_{x\not\leq t}f(t)={f}{}^{\vee}(x)\,,

that is, e~(f)=f\tilde{e}(f)={f}{}^{\vee}. Remark now that e:Y×Xop[X,Y]e:Y\times X^{op}\xrightarrow{\;\;\;\;\;}[X,Y]_{\vee} is the (set-theoretic) transpose of the trimorphism

y,x,t\displaystyle\langle y,x,t\rangle ={,txy,otherwise.\displaystyle=\begin{cases}\bot\,,&t\leq x\\ y\,,&\text{otherwise}\,.\end{cases}

Consequently, Raney’s transform (){(\,\cdot\,)}{}^{\vee} is the transpose of the map

YXopXY[X,𝟐]XYevalY2Y.\displaystyle Y\otimes X^{op}\otimes X\xrightarrow{\;\;\simeq\;\;}Y\otimes[X,\mathbf{2}]_{\vee}\otimes X\xrightarrow{\;\;Y\otimes\,eval\,\;\;}Y\otimes 2\xrightarrow{\;\;\simeq\;\;}Y\,. (2)

In the category SLatt, 𝟐\mathbf{2} is both the unit for the tensor product YXY\otimes X and its dual (YopXop)op[X,Yop](Y^{op}\otimes X^{op})^{op}\simeq[X,Y^{op}]_{\vee}. \ast-autonomous categories with this property are examples of isomix categories in sense of [4, 5, 3] where the transpose of the map in (2) is named 𝗆𝗂𝗑\mathsf{mix}. That Raney’s transforms are 𝗆𝗂𝗑\mathsf{mix} maps was recognized in [9] where also the nuclear objects—i.e. those objects whose 𝗆𝗂𝗑\mathsf{mix} maps are invertible—in the category SLatt were characterized (using Raney’s Theorem) as the completely distributive lattices. The importance of this characterization stems from the fact that the nucleus of a symmetric monoidal closed category—that is, the full subcategory of nuclear objects—yields a right adjoint to the forgetful functor from the category of compact closed categories to that of symmetric monoidal closed ones, as mentioned in [16]. In particular, the full subcategory of SLatt whose objects are the completely distributive lattices is more than symmetric monoidal closed or \ast-autonomous, it is compact closed [11].

A key property of Raney’s transforms, importantly used in [18, 17], is the following. For g[X,Y]g\in[X,Y]_{\land} and f[X,Y]f\in[X,Y]_{\vee}, the relations

ρ(g)\displaystyle\rho({g}{}^{\vee}) =(g),\displaystyle={\ell(g)}{}^{\land}\,, (f)\displaystyle\ell({f}{}^{\land}) =ρ(f),\displaystyle={\rho(f)}{}^{\vee}\,, (3)

hold. This property might be directly verified, as we did in [18, 17]. It might also be inferred from the commutativity of each square in the diagram below:

[X,Y]{[X,Y]_{\land}}[X,Y]{[X,Y]_{\vee}}YXop{Y\otimes X^{op}}XopY{X^{op}\otimes Y}[Yop,Xop]{[Y^{op},X^{op}]_{\land}}[Yop,Xop]{[Y^{op},X^{op}]_{\vee}}[Y,X]op{[Y,X]_{\vee}^{op}}[Y,X]op{[Y,X]_{\land}^{op}}()X,Y\scriptstyle{{(\,\cdot\,)}{}^{\vee}_{X,Y}}\scriptstyle{\ell}ρ\scriptstyle{\rho}σ\scriptstyle{\sigma}𝗆𝗂𝗑Y,X\scriptstyle{\mathsf{mix}_{Y,X}}𝗆𝗂𝗑Xop,Yop\scriptstyle{\mathsf{mix}_{X^{op},Y^{op}}}()Yop,Xop\scriptstyle{{(\,\cdot\,)}{}^{\vee}_{Y^{op},X^{op}}}()Y,X\scriptstyle{{(\,\cdot\,)}{}^{\land}_{Y,X}}

Naturality of Raney’s transforms.

Observe now that, for g:XXg:X^{\prime}\xrightarrow{\;\;\;\;\;}X and f:YYf:Y\xrightarrow{\;\;\;\;\;}Y^{\prime}, we have

fcy\displaystyle f\circ c_{y} =cf(y),\displaystyle=c_{f(y)}\,, axg\displaystyle a_{x}\circ g =aρ(g)(x),andfey,xg=ef(y),ρ(g)(x),\displaystyle=a_{\rho(g)(x)}\,,\quad\,\text{and}\,\quad f\circ e_{y,x}\circ g=e_{f(y),\rho(g)(x)}\,,

implying that the following diagram commutes:

YXop{Y\otimes X^{op}}[X,Y]{[X,Y]_{\vee}}YXop{Y^{\prime}\otimes X^{\prime}{}^{op}}[X,Y]{[X^{\prime},Y^{\prime}]_{\vee}}fgop\scriptstyle{f\otimes g^{op}}()Y,X\scriptstyle{{(\,\cdot\,)}{}^{\vee}_{Y,X}}[g,f]\scriptstyle{[g,f]_{\vee}}()Y,X\scriptstyle{{(\,\cdot\,)}{}^{\vee}_{Y^{\prime},X^{\prime}}}

That is, Raney’s transform (){(\,\cdot\,)}{}^{\vee} is natural in both its variables. Let us remark on the way the following:

Proposition 2.1.

There are exactly two natural arrows from YXopY\otimes X^{op} to [X,Y][X,Y]_{\vee}, the trivial one and Raney’s transform.

In order to simplify reading, we use ψ\psi both for a bimorphism ψ:Y×XopZ\psi:Y\times X^{op}\xrightarrow{\;\;\;\;\;}Z and for its extension to the tensor product ψ~:YXopZ\tilde{\psi}:Y\otimes X^{op}\xrightarrow{\;\;\;\;\;}Z.

Proof.

If ψ\psi is natural, then

ψ(y,x)\displaystyle\psi(y,x) =ψ(cy(),γx())=cyψ(,)ax,\displaystyle=\psi(c_{y}(\top),\gamma_{x}(\bot))=c_{y}\circ\psi(\top,\bot)\circ a_{x}\,,

since ρ(ax)=γx\rho(a_{x})=\gamma_{x}. Let f=ψ(,)f=\psi(\top,\bot). If f=f=\bot, then ψ\psi is the trivial map. Otherwise, fcf\neq c_{\bot} and f()f(\top)\neq\bot. Then, observing that fax=cf()axf\circ a_{x}=c_{f(\top)}\circ a_{x} and that cycz=cyc_{y}\circ c_{z}=c_{y} for zz\neq\bot, it follows that

ψ(y,x)\displaystyle\psi(y,x) =cyfax=cycf()ax=cyax.\displaystyle=c_{y}\circ f\circ a_{x}=c_{y}\circ c_{f(\top)}\circ a_{x}=c_{y}\circ a_{x}\,.
Remark 2.2.

Similar considerations can be developed if naturality is required in just one variable. For example, if the bimorphism ψ:Y×Xop[X,Y]\psi:Y\times X^{op}\xrightarrow{\;\;\;\;\;}[X,Y]_{\vee} is such that ψ(y,x)g=ψ(y,ρ(g)(x))\psi(y,x)\circ g=\psi(y,\rho(g)(x)), then ψ(y,x)=χ(y)ax\psi(y,x)=\chi(y)\circ a_{x} for some χ:Y[X,Y]\chi:Y\xrightarrow{\;\;\;\;\;}[X,Y]_{\vee}. \Diamond

For f:XYf:X\xrightarrow{\;\;\;\;\;}Y in the category SLatt, let j=ρ(f)fj=\rho(f)\circ f and o=fρ(f)o=f\circ\rho(f). Denote by XjX_{j} (resp. YoY_{o}) the set of fixed points of jj (resp., of oo). Then, we have a standard (epi,iso,mono)-factorization

X{X}Y{Y}Xj{X_{j}}Yo{Y_{o}}f\scriptstyle{f}j\scriptstyle{j}\scriptstyle{\simeq}

Thus, ff is mono if and only j=idXj=id_{X} and ff is epic if and only if o=idYo=id_{Y}. Notice that YoY_{o} is the image of XX under ff, while XjX_{j} is the image of YY under ρ(f)\rho(f). We apply this factorization to Raney’s transforms.

Definition 2.3.

A sup-preserving function f:XYf:X\xrightarrow{\;\;\;\;\;}Y is tight if f=f{f}{}^{\land}{}{}^{\vee}=f, or, equivalently, if it belongs to the image of [X,Y][X,Y]_{\land} via the Raney’s transform (){(\,\cdot\,)}{}^{\vee}. We let [X,Y]𝚝[X,Y]_{\vee}^{\mathtt{t}} be the set of tight functions from XX to YY.

By its definition, [X,Y]𝚝[X,Y]_{\vee}^{\mathtt{t}} is the sub-join-semilattice of [X,Y][X,Y]_{\vee} generated by the cyaxc_{y}\circ a_{x}. Moreover, it is easily seen that w¯z[X,Y]𝚝w\overline{\otimes}z\in[X,Y]_{\vee}^{\mathtt{t}}, for each wYw\in Y and zXz\in X, and that cyaxw¯zc_{y}\circ a_{x}\leq w\overline{\otimes}z if and only of ywy\leq w or zxz\leq x. From these relations, [X,Y]𝚝[X,Y]_{\vee}^{\mathtt{t}} yields a concrete representation of Wille’s tensor product Y^XopY\widehat{\otimes}X^{op}, see [20], which, for finite lattices, coincides with the Box tensor product of [8].

Next, we list some immediate consequences of naturality of Raney’s transforms:

Proposition 2.4.

The following statements hold:

  1. (i)

    [X,Y]𝚝[X,Y]_{\vee}^{\mathtt{t}} is a bi-ideal of [X,Y][X,Y]_{\vee}.

  2. (ii)

    For a complete lattice LL, the transform ():[L,L][L,L](\,\cdot\,)^{\vee}:[L,L]_{\land}\xrightarrow{\;\;\;\;\;}[L,L]_{\vee} is surjective if and only if idL[L,L]𝚝id_{L}\in[L,L]_{\vee}^{\mathtt{t}}, that is, if id=idid={id}{}^{\land}{}{}^{\vee}.

  3. (iii)

    For each complete lattice LL, the pair ([L,L]𝚝,)([L,L]_{\vee}^{\mathtt{t}},\circ) is a quantale.

Let us recall that Raney’s Theorem [14] characterizes completely distributive lattices as those complete lattices satisfying the identity

z\displaystyle z =zxyxy.\displaystyle=\bigvee_{z\not\leq x}\bigwedge_{y\not\leq x}y\,.

This identity is exactly the identity id=idid={id}{}^{\land}{}{}^{\vee} or, as we have seen in Proposition 2.4, the identity f=ff={f}{}^{\land}{}{}^{\vee} holding for each f[L,L]f\in[L,L]_{\vee}. Since complete distributivity is autodual (at least in a classical context), we derive that Raney’s transform ():[L,L][L,L](\,\cdot\,)^{\vee}:[L,L]_{\land}\xrightarrow{\;\;\;\;\;}[L,L]_{\vee} is surjective if and only if it is injective.

We conclude this section with a glance at the quantale ([L,L]𝚝,)([L,L]_{\vee}^{\mathtt{t}},\circ) of tight maps, where LL is an arbitrary complete lattice LL. We pause before for a technical lemma needed end the section and later on as well. Recalling the equations in (3), let us define

f\displaystyle f{}^{\ast} :=(f)(=ρ(f)),\displaystyle:=\ell({f}{}^{\land})\quad(\,={\rho(f)}{}^{\vee}\,)\,,

and observe the following:

Lemma 2.5.

For each xXx\in X, yYy\in Y, and f[X,Y]f\in[X,Y]_{\vee}, the following conditions are equivalent: (i) for all tXt\in X, xtx\leq t or yf(t)y\leq f(t), (ii) cyaxfc_{y}\circ a_{x}\leq f (iii) y¯xfy\underline{\otimes}x\leq{f}{}^{\land} (iv) yf(x)y\leq{f}{}^{\land}(x) (v) f(y)xf{}^{\ast}(y)\leq x (vi) fx¯yf{}^{\ast}\leq x\overline{\otimes}y.

Proof.
(i)(ii):(i)\Leftrightarrow(ii):  direct verification. (ii)(iii):(ii)\Leftrightarrow(iii):  since cyax=(y¯x)c_{y}\circ a_{x}={(y\underline{\otimes}x)}{}^{\vee}, cyaxfc_{y}\circ a_{x}\leq{f}{}^{\land} and by the adjunction ()(){(\,\cdot\,)}{}^{\vee}\dashv{(\,\cdot\,)}{}^{\land}. (iii)(iv):(iii)\Leftrightarrow(iv):  by the dual of Lemma 1.1. (iv)(v):(iv)\Leftrightarrow(v):  since fff{}^{\ast}\dashv{f}{}^{\land}. (iv)(v):(iv)\Leftrightarrow(v):  by Lemma 1.1.

Proposition 2.6.

Unless LL is a completely distributive lattice (in which case [L,L]𝚝=[L,L][L,L]_{\vee}^{\mathtt{t}}=[L,L]), the quantale ([L,L]𝚝,)([L,L]_{\vee}^{\mathtt{t}},\circ) is not unital.

Proof.

Let uu be unit for ([L,L]𝚝,)([L,L]_{\vee}^{\mathtt{t}},\circ) and write u=iIcyiaxiu=\bigvee_{i\in I}c_{y_{i}}\circ a_{x_{i}}. For an arbitrary xLx\in L, evaluate at xx the identity

ax\displaystyle a_{x} =axu=axcyiaxi\displaystyle=a_{x}\circ u=\bigvee a_{x}\circ c_{y_{i}}\circ a_{x_{i}}

and deduce that, for each iIi\in I, =axcyiaxi(x)\bot=a_{x}\circ c_{y_{i}}\circ a_{x_{i}}(x). This happens exactly when xxix\leq x_{i} or yixy_{i}\leq x, that is, when cyiaxix¯xc_{y_{i}}\circ a_{x_{i}}\leq x\overline{\otimes}x. Since xLx\in L and iIi\in I are arbitrary, we have, within [L,L][L,L], u=iIcyiaxixLx¯x=idLu=\bigvee_{i\in I}c_{y_{i}}\circ a_{x_{i}}\leq\bigwedge_{x\in L}x\overline{\otimes}x=id_{L}. Again, for yLy\in L arbitrary, evaluate at \top the identity

cy\displaystyle c_{y} =ucy=cyiaxicy.\displaystyle=u\circ c_{y}=\bigvee c_{y_{i}}\circ a_{x_{i}}\circ c_{y}\,.

and deduce that y=yxiyiy=\bigvee_{y\not\leq x_{i}}y_{i}. Considering that cyiaxiidLc_{y_{i}}\circ a_{x_{i}}\leq id_{L}, then we have yiid(xi)y_{i}\leq{id}{}^{\land}(x_{i}) and therefore

y\displaystyle y =yxiyiyxiid(xi)ytid(t)=id(y).\displaystyle=\bigvee_{y\not\leq x_{i}}y_{i}\leq\bigvee_{y\not\leq x_{i}}{id}{}^{\land}(x_{i})\leq\bigvee_{y\not\leq t}{id}{}^{\land}(t)={id}{}^{\land}{}{}^{\vee}(y)\,.

Since this holds for any yLy\in L, ididid\leq{id}{}^{\land}{}{}^{\vee} and since the opposite inclusion always holds, then id=idid={id}{}^{\land}{}{}^{\vee}. By Raney’s Theorem, LL is a completely distributive lattice. ∎

Recall that a dualizing element in a quantale (Q,)(Q,\circ) is an element 0Q0\in Q such that 0/(x\0)=(0/x)\0=x0/(x\backslash 0)=(0/x)\backslash 0=x, for each xQx\in Q. As consequences of Proposition 2.6, we obtain:

Corollary 2.7.

Unless LL is a completely distributive lattice,

  1. (i)

    the quantale ([L,L]𝚝,)([L,L]_{\vee}^{\mathtt{t}},\circ) has no dualizing element,

  2. (ii)

    the interior operator (){{(\,\cdot\,)}{}^{\land}}{}^{\vee} obtained by composing the two Raney’s transform is not a conucleus on [L,L][L,L]_{\vee}.

Proof.
(i) If 0 is dualizing, then 0\00\backslash 0 is a unit of the quantale. (ii) We argue that the inclusion (gf)gf{{(g\circ f)}{}^{\land}}{}^{\vee}\leq{{g}{}^{\land}}{}^{\vee}\circ{{f}{}^{\land}}{}^{\vee}, required for (){{(\,\cdot\,)}{}^{\land}}{}^{\vee} to be a conucleus on [L,L][L,L]_{\vee}, does not hold, unless LL is completely distributive. The opposite inclusion gf(gf){{g}{}^{\land}}{}^{\vee}\circ{{f}{}^{\land}}{}^{\vee}\leq{{(g\circ f)}{}^{\land}}{}^{\vee} holds since gf{{g}{}^{\land}}{}^{\vee}\circ{{f}{}^{\land}}{}^{\vee} belongs to [L,L]𝚝[L,L]_{\vee}^{\mathtt{t}}, gfgf{{g}{}^{\land}}{}^{\vee}\circ{{f}{}^{\land}}{}^{\vee}\leq g\circ f, and (gf){{(g\circ f)}{}^{\land}}{}^{\vee} is the greatest element of [L,L]𝚝[L,L]_{\vee}^{\mathtt{t}} below gfg\circ f. If (gf)=gf{{(g\circ f)}{}^{\land}}{}^{\vee}={{g}{}^{\land}}{}^{\vee}\circ{{f}{}^{\land}}{}^{\vee} for each f,g[L,L]f,g\in[L,L]_{\vee}, then 1{{1}{}^{\land}}{}^{\vee} is a unit for [L,L]𝚝[L,L]_{\vee}^{\mathtt{t}} and LL is completely distributive.

3 Dualizing elements of [L,L][L,L]_{\vee}

We investigate in this section dualizing elements of [L,L][L,L]_{\vee}. Proposition 2.6.18 in [6] states that if idL{id_{L}}{}^{\vee} is dualizing, then LL is completely distributive. Recall that a cyclic element in a quantale (Q,)(Q,\circ) is an element 0Q0\in Q such that 0/x=x\00/x=x\backslash 0, for each xQx\in Q. Trivially, the top element of a quantale is cyclic. Our work [17] proves that if [L,L][L,L]_{\vee} has a non-trivial cyclic element, then this element is idL{id_{L}}{}^{\vee} and, once more, cyclicity of idL{id_{L}}{}^{\vee} implies that LL is completely distributive. It was still open the possibility that [L,L][L,L]_{\vee} might have dualizing elements and no non-trivial cyclic elements. This is possible in principle, since the tool Mace4 [12] provided us with an example of a quantale where the unique dualizing element is not cyclic. The quantale is built on the modular lattice M5M_{5} (with atoms u,d,a,b,cu,d,a,b,c) and has the following multiplication table:

udabcuudabcddaadbbdccd\begin{array}[]{r|ccccccc}&\bot&u&d&a&b&c&\top\\ \hline\cr\bot&\bot&\bot&\bot&\bot&\bot&\bot&\bot\\ u&\bot&u&d&a&b&c&\top\\ d&\bot&d&\top&\top&\top&\top&\top\\ a&\bot&a&\top&\top&\top&d&\top\\ b&\bot&b&\top&d&\top&\top&\top\\ c&\bot&c&\top&\top&d&\top&\top\\ \top&\bot&\top&\top&\top&\top&\top&\top\end{array}

It is verified that dd is the only non-cyclic element and that, at the same time, it is the only dualizing element. For the quantale [L,L][L,L]_{\vee} we shall see that existence of a dualizing element again implies complete distributivity of LL (and therefore existence of a cyclic and dualizing element).

As this might be of more general interest, we are going to investigate how divisions \f\,\cdot\,\backslash f and f/f/\,\cdot\, by an arbitrary f[L,L]f\in[L,L]_{\vee} act on the ey,xe_{y,x} and y¯xy\overline{\otimes}x. To this end, we start remarking that Raney’s transforms intervene in the formulas for computing left adjoints of the maps cc and aa.

Lemma 3.1.

The functions c:Y[X,Y]c:Y\xrightarrow{\;\;\;\;\;}[X,Y]_{\vee} and a:Xop[X,Y]a:X^{op}\xrightarrow{\;\;\;\;\;}[X,Y]_{\vee} have both a left and a right adjoint. Namely, for each xXx\in X, yYy\in Y, and f[X,Y]f\in[X,Y]_{\vee}, the following relations hold:

cyf\displaystyle c_{y}\leq f iffyf(),\displaystyle\;\;\text{iff}\;\;y\leq{f}{}^{\land}(\bot)\,, fcy\displaystyle f\leq c_{y} ifff()y,\displaystyle\;\;\text{iff}\;\;f(\top)\leq y\,,
axf\displaystyle a_{x}\leq f ifff()x,\displaystyle\;\;\text{iff}\;\;f{}^{\ast}(\top)\leq x\,, fax\displaystyle f\leq a_{x} iffxρ(f)().\displaystyle\;\;\text{iff}\;\;x\leq\rho(f)(\bot)\,.

Using the relations stated in Lemma 3.1 computing divisions becomes an easy task.

Lemma 3.2.

For each xXx\in X, yYy\in Y, and f,g[X,Y]f,g\in[X,Y]_{\vee}, we have

f/ax\displaystyle f/a_{x}\ =f(x)¯(=cf(x)),\displaystyle={f}{}^{\land}(x)\overline{\otimes}\top\quad(\,=c_{{f}{}^{\land}(x)}\,)\,, cy\f\displaystyle c_{y}\backslash f =¯f(y)(=af(y)),\displaystyle=\bot\overline{\otimes}f{}^{\ast}(y)\quad(\,=a_{f{}^{\ast}(y)}\,)\,,
f/cy\displaystyle f/c_{y} =f()¯y,\displaystyle={f}{}^{\land}(\bot)\overline{\otimes}y\,, ax\f\displaystyle a_{x}\backslash f =x¯f().\displaystyle=x\overline{\otimes}f{}^{\ast}(\top)\,.
Proof.

We compute as follows:

gcy\f\displaystyle g\leq c_{y}\backslash f iffcygfiffcyagfiffcyaρ(g)()fifff(y)ρ(g)()iffgaf(y).\displaystyle\;\;\text{iff}\;\;c_{y}\circ g\leq f\;\;\text{iff}\;\;c_{y}\circ a_{\bot}\circ g\leq f\;\;\text{iff}\;\;c_{y}\circ a_{\rho(g)(\bot)}\leq f\;\;\text{iff}\;\;f{}^{\ast}(y)\leq\rho(g)(\bot)\;\;\text{iff}\;\;g\leq a_{f{}^{\ast}(y)}\,.

where we have used Lemma 2.5. Verification that f/ax=cf(x)f/a_{x}=c_{{f}{}^{\land}(x)} is similar. The other two identities are verified as follows:

hf/cy\displaystyle h\leq f/c_{y} iffhcyfiffch(y)fiffh(y)f()iffhf()¯y,\displaystyle\;\;\text{iff}\;\;h\circ c_{y}\leq f\;\;\text{iff}\;\;c_{h(y)}\leq f\;\;\text{iff}\;\;h(y)\leq{f}{}^{\land}(\bot)\;\;\text{iff}\;\;h\leq{f}{}^{\land}(\bot)\overline{\otimes}y\,,
and, dually,
hax\f\displaystyle h\leq a_{x}\backslash f iffaxhfiffaρ(h)(x)fifff()ρ(h)(x)iffh(f())xiffhx¯f().\displaystyle\;\;\text{iff}\;\;a_{x}\circ h\leq f\;\;\text{iff}\;\;a_{\rho(h)(x)}\leq f\;\;\text{iff}\;\;f{}^{\ast}(\top)\leq\rho(h)(x)\;\;\text{iff}\;\;h(f{}^{\ast}(\top))\leq x\;\;\text{iff}\;\;h\leq x\overline{\otimes}f{}^{\ast}(\top)\,.

The relations in the following proposition are then easily derived.

Proposition 3.3.

The following relations hold:

f/(cyax)\displaystyle f/(c_{y}\circ a_{x}) =f(x)¯y,\displaystyle={f}{}^{\land}(x)\overline{\otimes}y\,, (cyax)\f\displaystyle(c_{y}\circ a_{x})\backslash f =x¯f(y),\displaystyle=x\overline{\otimes}f{}^{\ast}(y)\,,
f/(y¯x)\displaystyle f/(y\overline{\otimes}x) =f(x)¯f()¯y,\displaystyle={f}{}^{\land}(x)\overline{\otimes}\top\land{f}{}^{\land}(\bot)\overline{\otimes}y\,, (y¯x)\f\displaystyle(y\overline{\otimes}x)\backslash f =¯f(y)x¯f().\displaystyle=\bot\overline{\otimes}f{}^{\ast}(y)\land x\overline{\otimes}f{}^{\ast}(\top)\,.

From Proposition 3.3, it follows that

cf(x)ay\displaystyle c_{{f}{}^{\land}(x)}\circ a_{y} f/(y¯x),\displaystyle\leq f/(y\overline{\otimes}x)\,, cxaf(y)\displaystyle c_{x}\circ a_{f{}^{\ast}(y)} (y¯x)\f.\displaystyle\leq(y\overline{\otimes}x)\backslash f\,.

If ff{}^{\ast} is invertible, then its inverse is also its right adjoint. From the uniqueness of the right adjoint it follows that ff{}^{\ast} is inverted by its right adjoint f{f}{}^{\land}. Thus, ff{}^{\ast} is invertible if and only if f{f}{}^{\land} is invertible, in which case we have f()=f{}^{\ast}(\top)=\top and f()={f}{}^{\land}(\bot)=\bot, since both ff{}^{\ast} and f{f}{}^{\land} are bicontinuous. In some important case, the expressions exhibited in Proposition 3.3 simplify:

Corollary 3.4.

If f()={f}{}^{\land}(\bot)=\bot (resp., f()=f{}^{\ast}(\top)=\top), then

f/(y¯x)\displaystyle f/(y\overline{\otimes}x) =cf(x)ay(resp.,(y¯x)\f=cxaf(y)).\displaystyle=c_{{f}{}^{\land}(x)}\circ a_{y}\quad(\text{resp.,}\;\;(y\overline{\otimes}x)\backslash f=c_{x}\circ a_{f{}^{\ast}(y)})\,.

These relations hold as soon as either f{f}{}^{\land} or ff{}^{\ast} is invertible.

Example 3.5.

If LL is a completely distributive lattice, then the relation idL=idL{id_{L}}{}^{\land}{}{}^{\vee}=id_{L} holds, by Raney’s Theorem [14]. Let o=idLo={id_{L}}{}^{\vee}, then o{o}{}^{\land} is invertible, since it is the identity. Necessarily, we also have o=idLo{}^{\ast}=id_{L} and therefore:

o/(cyax)\displaystyle o/(c_{y}\circ a_{x}) =(cyax)\o=x¯y,\displaystyle=(c_{y}\circ a_{x})\backslash o=x\overline{\otimes}y\,, o/(y¯x)\displaystyle o/(y\overline{\otimes}x) =(y¯x)\o=cxay.\displaystyle=(y\overline{\otimes}x)\backslash o=c_{x}\circ a_{y}\,. \Diamond
Theorem 3.6.

If f[L,L]f\in[L,L]_{\vee} is dualizing, then f{f}{}^{\land} and ff{}^{\ast} are inverse to each other and LL is completely distributive.

Proof.

If ff is dualizing, then, for all x,yLx,y\in L,

cy\displaystyle c_{y} =f/(cy\f)=f/af(y)=cf(f(y)),ax=(f/ax)\f=cf(x)\f=af(f(x)),\displaystyle=f/(c_{y}\backslash f)=f/a_{f{}^{\ast}(y)}=c_{{f}{}^{\land}(f{}^{\ast}(y))}\,,\;\quad a_{x}=(f/a_{x})\backslash f=c_{{f}{}^{\land}(x)}\backslash f=a_{f{}^{\ast}({f}{}^{\land}(x))}\,,

and since both cc and aa are injective, then y=f(f(y))y={f}{}^{\land}(f{}^{\ast}(y)) and x=f(f(x))x=f{}^{\ast}({f}{}^{\land}(x)). Thus, f{f}{}^{\land} and ff{}^{\ast} are inverse to each other. Remark now that f[L,L]𝚝f{}^{\ast}\in[L,L]_{\vee}^{\mathtt{t}}, since f=(f)=ρ(f)f{}^{\ast}=\ell({f}{}^{\land})={\rho(f)}{}^{\vee}. Then idL=ff[L,L]𝚝id_{L}={f}{}^{\land}\circ f{}^{\ast}\in[L,L]_{\vee}^{\mathtt{t}}, since [L,L]𝚝[L,L]_{\vee}^{\mathtt{t}} is an ideal of [L,L][L,L]_{\vee}. Then LL is completely distributive by Raney’s Theorem. ∎

It is not difficult to give a direct proof of the converse, namely that if ff{}^{\ast} is invertible, then ff is dualizing. We prove this as a general statement about involutive residuated lattices. Notice that the map sending ff to ff{}^{\ast} is definable in the language of involutive residuated lattices, since f=f\of{}^{\ast}=f\backslash o, where o=idLo={id_{L}}{}^{\vee} is the canonical cyclic dualizing element of [L,L][L,L]_{\vee} (if LL is completely distributive). The statement in the following Proposition 3.7 is implicit in the definition of a (symmetric) compact closed category in [11] (see [21, §5] for the non symmetric version of this notion).

Proposition 3.7.

In every involutive residuated lattice QQ, ff is dualizing if and only if ff{}^{\ast} is invertible.

Proof.

If fQf\in Q is dualizing, then f/(x\f)=(f/x)\f=xf/(x{}^{\ast}\backslash f)=(f/x{}^{\ast})\backslash f=x{}^{\ast}, for each xQx\in Q. Using well known identities of involutive residuated lattices, compute as follows:

x\displaystyle x =x=(f/(x\f))=(x\f)f=(x/f)f.\displaystyle=x{}^{\ast}{}^{\ast}=(f/(x{}^{\ast}\backslash f)){}^{\ast}=(x{}^{\ast}\backslash f)\circ f{}^{\ast}=(x/f{}^{\ast})\circ f{}^{\ast}\,.

Letting x=1x=1, then 1=(1/f)f1=(1/f{}^{\ast})\circ f{}^{\ast}. We derive 1=f(f\1)1=f{}^{\ast}\circ(f{}^{\ast}\backslash 1) similarly, from which it follows that ff{}^{\ast} is inverted by 1/f=f\11/f{}^{\ast}=f{}^{\ast}\backslash 1.

Conversely, suppose that ff{}^{\ast} is invertible, say fg=gf=1f{}^{\ast}\circ g=g\circ f{}^{\ast}=1. It immediately follows that g=1/f=f\1g=1/f{}^{\ast}=f{}^{\ast}\backslash 1. Again, for each xQx\in Q,

x\displaystyle x{}^{\ast} =x(1/f)f(x/f)f,\displaystyle=x{}^{\ast}\circ(1/f{}^{\ast})\circ f{}^{\ast}\leq(x{}^{\ast}/f{}^{\ast})\circ f{}^{\ast}\,,

and then, dualizing this relation, we obtain

x=x\displaystyle x=x{}^{\ast}{}^{\ast} ((x/f)f)=f/(x\f).\displaystyle\geq((x{}^{\ast}/f{}^{\ast})\circ f{}^{\ast}){}^{\ast}=f/(x\backslash f)\,.

Since xf/(x\f)x\leq f/(x\backslash f) always hold, we have x=f/(x\f)x=f/(x\backslash f). The identity x=(f/x)\fx=(f/x)\backslash f is derived similarly. ∎

Example 3.8.

If L=[0,1]L=[0,1], then ff is dualizing if and only if it is invertible. Indeed, ff is dualizing iff ff{}^{\ast} is invertible iff f{f}{}^{\land} is invertible. Now, if f{f}{}^{\land} is invertible, then it is continuous and f=f{f}{}^{\land}=f; therefore ff is invertible. Similarly, if ff is invertible, then it is continuous and f=f{f}{}^{\land}=f; therefore f{f}{}^{\land} is invertible. \Diamond

Example 3.9.

Consider a poset PP, the complete lattice 𝒟(P){\mathcal{D}}(P) of downsets of PP, and recall that 𝒟(P){\mathcal{D}}(P) is completely distributive. The quantale [𝒟(P),𝒟(P)][{\mathcal{D}}(P),{\mathcal{D}}(P)]_{\vee} is isomorphic to the quantale of weakening relations (profuctors/bimodules) on the poset PP. These are the relations RP×PR\subseteq P\times P such that yRxyRx, yyy^{\prime}\leq y, and xxx\leq x^{\prime} imply xRyx^{\prime}Ry^{\prime} (for all x,x,y,yPx,x^{\prime},y,y^{\prime}\in P). Thus, weakening relations are downsets of P×PopP\times P^{op} and the bijection between [𝒟(P),𝒟(P)][{\mathcal{D}}(P),{\mathcal{D}}(P)]_{\vee} and 𝒟(P×Pop){\mathcal{D}}(P\times P^{op}) goes along the lines described in previous sections, since

[𝒟(P),𝒟(P)]\displaystyle[{\mathcal{D}}(P),{\mathcal{D}}(P)]_{\vee} 𝒟(P)𝒟(Pop)𝒟(P×Pop).\displaystyle\simeq{\mathcal{D}}(P)\otimes{\mathcal{D}}(P^{op})\simeq{\mathcal{D}}(P\times P^{op})\,.

Explicitly, this bijection, sending ff to RfR_{f}, is such that

(y,x)Rf\displaystyle(y,x)\in R_{f} iffyf(x)iffcyaxf,\displaystyle\;\;\text{iff}\;\;y\in f(\downarrow x)\;\;\text{iff}\;\;c_{\downarrow y}\circ a_{\downarrow x}\leq f\,, (4)

where, for xPx\in P, x:={yPyx}\downarrow x:=\{\,y\in P\mid y\leq x\,\}. We have seen that dualizing elements of [𝒟(P),𝒟(P)][{\mathcal{D}}(P),{\mathcal{D}}(P)]_{\vee} are in bijection with automorphisms of 𝒟(P){\mathcal{D}}(P) which in turn are in bijection with automorphisms of PP. Given such an automorphism, we aim at computing the dualizing weakening relation corresponding to this automorphism. To this end, let us recall that, when LL is completely distributive, o=id=(id)o={id}{}^{\vee}=\ell({id}{}^{\land}) is the unique non-trivial cyclic element. Observe that since o=ido=id{}^{\ast}, oo is also dualizing and that

(y,x)Ro\displaystyle(y,x)\in R_{o} iffxy.\displaystyle\;\;\text{iff}\;\;x\not\leq y\,.

For f[𝒟(P),𝒟(P)]f\in[{\mathcal{D}}(P),{\mathcal{D}}(P)]_{\vee}, recall that f=f\of{}^{\ast}=f\backslash o. We use the relations in (4) to compute the dualizing element of 𝒟(P×Pop){\mathcal{D}}(P\times P^{op}) corresponding to an invertible order preserving map g:PPg:P\xrightarrow{\;\;\;\;\;}P, as follows:

(y,x)R𝒟(g)\displaystyle(y,x)\in R_{{\mathcal{D}}(g){}^{\ast}} iffcyax𝒟(g)=𝒟(g)\o\displaystyle\;\;\text{iff}\;\;c_{\downarrow y}\circ a_{\downarrow x}\leq{\mathcal{D}}(g){}^{\ast}={\mathcal{D}}(g)\backslash o
iff𝒟(g)cyax=c𝒟(g)(y)ax=cg(y)ax0\displaystyle\;\;\text{iff}\;\;{\mathcal{D}}(g)\circ c_{\downarrow y}\circ a_{\downarrow x}=c_{{\mathcal{D}}(g)(\downarrow y)}\circ a_{\downarrow x}=c_{\downarrow g(y)}\circ a_{\downarrow x}\leq 0
iffxg(y).\displaystyle\;\;\text{iff}\;\;x\not\leq g(y)\,. \Diamond

4 Further bimorphisms, bijections, directions

Even when Raney’s transforms are not inverse to each other, it might still be asked whether there are other isomorphisms between [L,L][L,L]_{\land} and [L,L][L,L]_{\vee}. By Fact 1.3, this question amounts to understand whether [L,L][L,L]_{\vee} is autodual.

Let us discuss the case when LL is a finite lattice. We use 𝖩(L)\mathsf{J}(L) for the set of join-irreducible elements of LL and 𝖬(L)\mathsf{M}(L) for the set of meet-irreducible elements of LL. The reader will have no difficulties convincing himself of the following statement:

Lemma 4.1.

A map f[L,L]f\in[L,L]_{\vee} is meet-irreducible if and only if it is an elementary tensor of the form m¯jm\overline{\otimes}j with m𝖬(L)m\in\mathsf{M}(L) and j𝖩(L)j\in\mathsf{J}(L).

The following statement might instead be less immediate:

Lemma 4.2.

For each j𝖩(L)j\in\mathsf{J}(L) and m𝖬(L)m\in\mathsf{M}(L), the map ej,me_{j,m} is join-irreducible.

Proof.

Let m𝖬(L)m\in\mathsf{M}(L) and j𝖩(L)j\in\mathsf{J}(L), and let us use mm^{\ast} to denote the unique upper cover of mm. Suppose that ej,m=iIfie_{j,m}=\bigvee_{i\in I}f_{i}. By evaluating the two sides of this equality at mm^{\ast}, we obtain j=iIfi(m)j=\bigvee_{i\in I}f_{i}(m^{\ast}) and therefore j=fi(m)j=f_{i}(m^{\ast}) for some iIi\in I. If tmt\leq m, then fi(t)em,j(t)=f_{i}(t)\leq e_{m,j}(t)=\bot. Suppose now that tmt\not\leq m, so m<tmm<t\vee m and mtmm^{\ast}\leq t\vee m. Observe also that fi(t)ej,m(t)=jf_{i}(t)\leq e_{j,m}(t)=j, since ej,m=iIfie_{j,m}=\bigvee_{i\in I}f_{i}. Then j=fi(m)fi(mt)=fi(m)fi(t)=fi(t)=fi(t)j=f_{i}(m^{\ast})\leq f_{i}(m\vee t)=f_{i}(m)\vee f_{i}(t)=\bot\vee f_{i}(t)=f_{i}(t), so j=fi(t)j=f_{i}(t). We have argued that fi(t)=ej,mf_{i}(t)=e_{j,m}, for all tLt\in L, and therefore that fi=ej,mf_{i}=e_{j,m}. ∎

Theorem 4.3.

If LL is a finite lattice and [L,L][L,L]_{\vee} is autodual, then LL is distributive.

Proof.

If ψ:[L,L]op[L,L]\psi:[L,L]_{\vee}^{op}\xrightarrow{\;\;\;\;\;}[L,L]_{\vee} is invertible, then ψ\psi restricts to a bijection 𝖬([L,L])𝖩([L,L])\mathsf{M}([L,L]_{\vee})\xrightarrow{\;\;\;\;\;}\mathsf{J}([L,L]_{\vee}), so these two sets have same cardinality. For m𝖬(L)m\in\mathsf{M}(L) and j𝖩(L)j\in\mathsf{J}(L), the ej,me_{j,m} as well as the elementary tensors m¯jm\overline{\otimes}j are pairwise distinct. Therefore, we have

|𝖬(L)|×|𝖩(L)||𝖩([L,L])|=|𝖬([L,L])|=|𝖬(L)|×|𝖩(L)||\mathsf{M}(L)|\times|\mathsf{J}(L)|\leq|\mathsf{J}([L,L]_{\vee})|=|\mathsf{M}([L,L]_{\vee})|=|\mathsf{M}(L)|\times|\mathsf{J}(L)|

and |𝖬(L)|×|𝖩(L)|=|𝖩([L,L])||\mathsf{M}(L)|\times|\mathsf{J}(L)|=|\mathsf{J}([L,L]_{\vee})|. That is, the elements ej,me_{j,m} are all the join-irreducible elements of [L,L][L,L]_{\vee} and therefore the set {ej,mj𝖩(L),m𝖬(L)}\{\,e_{j,m}\mid j\in\mathsf{J}(L),m\in\mathsf{M}(L)\,\} generates [L,L][L,L]_{\vee} under joins. It follows that [L,L]=[L,L]𝚝[L,L]_{\vee}=[L,L]_{\vee}^{\mathtt{t}} and that LL is distributive. ∎

We do not know yet if the theorem above can be generalized to infinite complete lattices or whether there is some fancy infinite complete lattice LL that is not completely distributive and such that [L,L][L,L]_{\vee} is autodual. It is clear, however, that in order to construct such a fancy lattice, properties of bimorphisms ψ:L×Lop[L,L]\psi:L\times L^{op}\xrightarrow{\;\;\;\;\;}[L,L]_{\vee} need to be investigated. What are the properties of a bimorphism ψ\psi forcing LL to be completely distributive when ψ~\tilde{\psi} is surjective? Taking the bimorphism ee as example, let us abstract part of Raney’s Theorem:

Proposition 4.4.

Let ψ:L×Lop[L,L]\psi:L\times L^{op}\xrightarrow{\;\;\;\;\;}[L,L]_{\vee} be a bimorphism such that for each x,yLx,y\in L, the image of LL under ψ(y,x)\psi(y,x) is a finite chain. If idLid_{L} belongs to the image of ψ~:[L,L][L,L]\tilde{\psi}:[L,L]_{\land}\xrightarrow{\;\;\;\;\;}[L,L]_{\vee}, then LL is a completely distributive lattice.

Proof.

Let zZz\in Z such that z=iIjJizjz=\bigwedge_{i\in I}\bigvee_{j\in J_{i}}z_{j}. We aim at showing that zsiIzs(i)z\leq\bigvee_{s}\bigwedge_{i\in I}z_{s(i)}, with the index ss ranging on choice functions s:IiIJis:I\xrightarrow{\;\;\;\;\;}\bigcup_{i\in I}J_{i} (ss is a choice function if s(i)Jis(i)\in J_{i}, for each iIi\in I). Since idL={ψ(y,x)ψ(y,x)idL}id_{L}=\bigvee\{\,\psi(y,x)\mid\psi(y,x)\leq id_{L}\,\}, we also have z={ψ(y,x)(z)ψ(y,x)idL}z=\bigvee\{\,\psi(y,x)(z)\mid\psi(y,x)\leq id_{L}\,\} and therefore, in order to achieve our goal, it will be enough to show that for each y,xLy,x\in L, if ψ(y,x)idL\psi(y,x)\leq id_{L}, then ψ(y,x)(z)siIzs(i)\psi(y,x)(z)\leq\bigvee_{s}\bigwedge_{i\in I}z_{s(i)}. Let y,xy,x be such that ψ(y,x)idL\psi(y,x)\leq id_{L}, fix iIi\in I, and observe then that

ψ(y,x)(z)\displaystyle\psi(y,x)(z) ψ(y,x)(jJizj)=jJiψ(y,x)(zj),\displaystyle\leq\psi(y,x)(\bigvee_{j\in J_{i}}z_{j})=\bigvee_{j\in J_{i}}\psi(y,x)(z_{j})\,,

since zjJizjz\leq\bigvee_{j\in J_{i}}z_{j}. Since the set {ψ(y,x)(zj)jJi}\{\,\psi(y,x)(z_{j})\mid j\in J_{i}\,\} is finite and directed (it is a finite chain), it has a maximum: there exists j(i)Jjj(i)\in J_{j} such that jJiψ(y,x)(zj)=ψ(y,x)(zj(i))\bigvee_{j\in J_{i}}\psi(y,x)(z_{j})=\psi(y,x)(z_{j(i)}). It follows that

ψ(y,x)(z)\displaystyle\psi(y,x)(z) ψ(y,x)(zj(i))zj(i),\displaystyle\leq\psi(y,x)(z_{j(i)})\leq z_{j(i)}\,,

since ψ(y,x)idL\psi(y,x)\leq id_{L}. By letting ii vary, we have constructed a choice function j:IiIJij:I\xrightarrow{\;\;\;\;\;}\bigcup_{i\in I}J_{i} such that ψ(y,x)(z)iIzj(i)\psi(y,x)(z)\leq\bigwedge_{i\in I}z_{j(i)}, and consequently ψ(y,x)(z)siIzs(i)\psi(y,x)(z)\leq\bigvee_{s}\bigwedge_{i\in I}z_{s(i)}. ∎

Bimorphisms satisfying the conditions of Proposition 4.4 might be easily constructed by taking f[L,[L,L]]f\in[L,[L,L]_{\vee}]_{\vee}, g[Lop,Lop]g\in[L^{op},L^{op}]_{\vee} (resp., f[L,L]f\in[L,L]_{\vee} and g[Lop,[L,L]]g\in[L^{op},[L,L]_{\vee}]_{\vee}), and defining then

ψ(y,x)\displaystyle\psi(y,x) :=f(y)ag(x)(resp., ψ(y,x):=cf(y)g(x)).\displaystyle:=f(y)\circ a_{g(x)}\quad(\,\text{resp., }\psi(y,x):=c_{f(y)}\circ g(x)\,)\,.

These bimorphisms satisfy the conditions of Proposition 4.4, since they only take two values. As a consequence of the proposition, they cannot be used to construct a fancy dual isomorphism of [L,L][L,L]_{\vee}.

References

  • [1]
  • [2] Michael Barr (1979): \ast-autonomous categories. Lecture Notes in Mathematics 752, Springer, Berlin, 10.1007/BFb0064582.
  • [3] R. F. Blute, J. R. B. Cockett & R. A. G. Seely (2000): Feedback for linearly distributive categories: traces and fixpoints. J. Pure Appl. Algebra 154(1-3), pp. 27–69, 10.1016/S0022-4049(99)00180-2.
  • [4] R. F. Blute, J. R. B. Cockett, R. A. G. Seely & T. H. Trimble (1996): Natural deduction and coherence for weakly distributive categories. J. Pure Appl. Algebra 113(3), pp. 229–296, 10.1016/0022-4049(95)00159-X.
  • [5] J. R. B. Cockett & R. A. G. Seely (1997): Proof theory for full intuitionistic linear logic, bilinear logic, and MIX categories. Theory Appl. Categ. 3, pp. No. 5, 85–131.
  • [6] Patrik Eklund, Javier Gutiérrez García, Ulrich Höhle & Jari Kortelainen (2018): Semigroups in complete lattices. Developments in Mathematics 54, Springer, Cham, 10.1007/978-3-319-78948-4.
  • [7] Nikolaos Galatos, Peter Jipsen, Tomasz Kowalski & Hiroakira Ono (2007): Residuated Lattices: An Algebraic Glimpse at Substructural Logics. Studies in Logic and the Foundations of Mathematics 151, Elsevier, 10.1016/S0049-237X(07)80005-X.
  • [8] G. Grätzer & F. Wehrung (1999): A new lattice construction: the box product. J. Algebra 221(1), pp. 315–344, 10.1006/jabr.1999.7975.
  • [9] D. A. Higgs & K. A. Rowe (1989): Nuclearity in the category of complete semilattices. J. Pure Appl. Algebra 57(1), pp. 67–78, 10.1016/0022-4049(89)90028-5.
  • [10] André Joyal & Myles Tierney (1984): An extension of the Galois theory of Grothendieck. Mem. Amer. Math. Soc. 51(309), pp. vii+71, 10.1090/memo/0309.
  • [11] G. M. Kelly & M. L. Laplaza (1980): Coherence for compact closed categories. J. Pure Appl. Algebra 19, pp. 193–213, 10.1016/0022-4049(80)90101-2.
  • [12] W. McCune (2005–2010): Prover9 and Mace4. http://www.cs.unm.edu/~mccune/prover9/.
  • [13] Evelyn Nelson (1976): Galois connections as left adjoint maps. Comment. Math. Univ. Carolinae 17(3), pp. 523–541.
  • [14] George N. Raney (1960): Tight Galois connections and complete distributivity. Trans. Amer. Math. Soc. 97, pp. 418–426, 10.2307/1993380.
  • [15] Kimmo I. Rosenthal (1990): Quantales and their applications. Pitman Research Notes in Mathematics Series 234, Longman Scientific & Technical, Harlow.
  • [16] K. A. Rowe (1988): Nuclearity. Canad. Math. Bull. 31(2), pp. 227–235, 10.4153/CMB-1988-035-5.
  • [17] Luigi Santocanale (2020): The Involutive Quantaloid of Completely Distributive Lattices. In Uli Fahrenberg, Peter Jipsen & Michael Winter, editors: RAMiCS 2020, Palaiseau, France, April 8-11, 2020, Proceedings [postponed], Lecture Notes in Computer Science 12062, Springer, pp. 286–301, 10.1007/978-3-030-43520-2_18.
  • [18] Luigi Santocanale & Maria João Gouveia (2020): The continuous weak order. Journal of Pure and Applied Algebra, 10.1016/j.jpaa.2020.106472. In press.
  • [19] Zahava Shmuely (1974): The structure of Galois connections. Pacific J. Math. 54(2), pp. 209–225, 10.2140/pjm.1974.54.209.
  • [20] Rudolf Wille (1985): Tensorial decomposition of concept lattices. Order 2(1), pp. 81–95, 10.1007/BF00337926.
  • [21] David N. Yetter (2001): Functorial knot theory. Series on Knots and Everything 26, World Scientific Publishing Co., Inc., River Edge, NJ, 10.1142/9789812810465.