Dualizing sup-preserving endomaps of a complete lattice
Abstract
It is argued in [6] that the quantale of sup-preserving endomaps of a complete lattice is a Girard quantale exactly when is completely distributive. We have argued in [17] that this Girard quantale structure arises from the dual quantale of inf-preserving endomaps of 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 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 developed in [17] by investigating its dualizing elements. We argue that if has the structure a Frobenius quantale, that is, if it has a dualizing element, not necessarily a cyclic one, then 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 and automorphisms of . Finally, we also argue that if is finite and is autodual, then is distributive.
1 Lattice structure of the homsets in SLatt
The homset in SLatt.
The category SLatt of complete lattices and sup-preserving functions is a well-known -autonomous category [2, 10, 9, 6]. For complete lattices , we denote by the homset in this category. The two-element Boolean algebra is a dualizing element and is, as a lattice, isomorphic to the dual lattice . More generally, the functor is naturally isomorphic to the functor where, for , is the right adjoint of , noted here by (the left adjoint of an inf-preserving function shall be denoted by ).
Let us describe the internal structure of the homset as a complete lattice. For and , we define the following elements of :
Lemma 1.1.
For each , , and , if and only if . Consequently, for each ,
and the sup-preserving functions of the form generate under arbitrary meets.
The tensor notation arises from the canonical isomorphism of -autonomous categories. That is, is dual to the tensor product and the functions correspond to elementary tensors of . For , , and , let us recall that there exists uniquely determined maps and satisfying
The binary operations and are known under several names: they yield left and right Kan extensions and (often when ) 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) , (ii) , (iii) , (iv) .
Proof.
The relations and are well known, see e.g. [20], and and are immediate consequences of these relations.
Let us focus on and notice that, in order to make sense of these relations, we need to assume , , , and . Observe that if and only if and therefore, in view of Lemma 1.1 and of the definion of , the relation . Next, the condition amounts to , that is, for all , if then . Clearly this condition is equivalent to , and therefore to . Finally, the relation is directly verified. However, let us observe the abuse of notation, since for the maps and to be composable we need to assume either or and . ∎
Inf-preserving functions as tensor product.
Let denote the poset of inf-preserving functions from to , with the pointwise ordering. Observe that, as a set, equals . Yet, as a poset or a lattice, the equality is the correct one. As a matter of fact, we have iff , all , iff , all , iff . Using standard isomorphisms of -autonomous categories, we have
That is, the set of inf-preserving functions from to can be taken as a concrete realization of the tensor product . This should not come as a surprise, since it is well-known that the set of Galois connections from to —that is, pairs of functions such that iff —realizes the tensor product 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 to .
Notice now that
(1) |
from which we derive the following principle:
Fact 1.3.
There is a bijection beweeen sup-preserving functions from to and sup-preserving functions from to .
The map yielding the isomorphism in equation (1) is , the operation of taking the right adjoint. The bijection stated in Fact 1.3 is therefore obtained by precomposing with .
We exploit now the work done for to recap the structure of as a tensor product. Consider the maps
By dualizing Lemma 1.1, we observe that the relation holds, the maps realize the elementary tensors of the (abstract) tensor product , is join-generated by these maps, and every can be canonically written as .
Recall that a bimorphism is a function that is sup-preserving in each variable, separately. This in particular means that infs in are transformed into sups in . The universal property of as a tensor product can be therefore stated as follows:
Fact 1.4.
Given a bimorphism , there exists a unique sup-preserving functions such that . For , is defined by
2 Raney’s transforms
For and , define
It is easily seen that has a right adjoint, so , and that has a left adjoint, so belongs to . We call the operations and 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 if and only if , so is right adjoint to . Raney’s transforms have been the key ingredient allowing us to prove in [18] that is a Girard quantale if 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 sending to and its extension
By evaluating at , we obtain
that is, . Remark now that is the (set-theoretic) transpose of the trimorphism
Consequently, Raney’s transform is the transpose of the map
(2) |
In the category SLatt, is both the unit for the tensor product and its dual . -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 . That Raney’s transforms are maps was recognized in [9] where also the nuclear objects—i.e. those objects whose 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 -autonomous, it is compact closed [11].
A key property of Raney’s transforms, importantly used in [18, 17], is the following. For and , the relations
(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:
Naturality of Raney’s transforms.
Observe now that, for and , we have
implying that the following diagram commutes:
That is, Raney’s transform is natural in both its variables. Let us remark on the way the following:
Proposition 2.1.
There are exactly two natural arrows from to , the trivial one and Raney’s transform.
In order to simplify reading, we use both for a bimorphism and for its extension to the tensor product .
Proof.
If is natural, then
since . Let . If , then is the trivial map. Otherwise, and . Then, observing that and that for , it follows that
∎ |
Remark 2.2.
Similar considerations can be developed if naturality is required in just one variable. For example, if the bimorphism is such that , then for some .
For in the category SLatt, let and . Denote by (resp. ) the set of fixed points of (resp., of ). Then, we have a standard (epi,iso,mono)-factorization
Thus, is mono if and only and is epic if and only if . Notice that is the image of under , while is the image of under . We apply this factorization to Raney’s transforms.
Definition 2.3.
A sup-preserving function is tight if , or, equivalently, if it belongs to the image of via the Raney’s transform . We let be the set of tight functions from to .
By its definition, is the sub-join-semilattice of generated by the . Moreover, it is easily seen that , for each and , and that if and only of or . From these relations, yields a concrete representation of Wille’s tensor product , 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:
-
(i)
is a bi-ideal of .
-
(ii)
For a complete lattice , the transform is surjective if and only if , that is, if .
-
(iii)
For each complete lattice , the pair is a quantale.
Let us recall that Raney’s Theorem [14] characterizes completely distributive lattices as those complete lattices satisfying the identity
This identity is exactly the identity or, as we have seen in Proposition 2.4, the identity holding for each . Since complete distributivity is autodual (at least in a classical context), we derive that Raney’s transform is surjective if and only if it is injective.
We conclude this section with a glance at the quantale of tight maps, where is an arbitrary complete lattice . We pause before for a technical lemma needed end the section and later on as well. Recalling the equations in (3), let us define
and observe the following:
Lemma 2.5.
For each , , and , the following conditions are equivalent: (i) for all , or , (ii) (iii) (iv) (v) (vi) .
Proof.
Proposition 2.6.
Unless is a completely distributive lattice (in which case ), the quantale is not unital.
Proof.
Let be unit for and write . For an arbitrary , evaluate at the identity
and deduce that, for each , . This happens exactly when or , that is, when . Since and are arbitrary, we have, within , . Again, for arbitrary, evaluate at the identity
and deduce that . Considering that , then we have and therefore
Since this holds for any , and since the opposite inclusion always holds, then . By Raney’s Theorem, is a completely distributive lattice. ∎
Recall that a dualizing element in a quantale is an element such that , for each . As consequences of Proposition 2.6, we obtain:
Corollary 2.7.
Unless is a completely distributive lattice,
-
(i)
the quantale has no dualizing element,
-
(ii)
the interior operator obtained by composing the two Raney’s transform is not a conucleus on .
Proof.
∎
3 Dualizing elements of
We investigate in this section dualizing elements of . Proposition 2.6.18 in [6] states that if is dualizing, then is completely distributive. Recall that a cyclic element in a quantale is an element such that , for each . Trivially, the top element of a quantale is cyclic. Our work [17] proves that if has a non-trivial cyclic element, then this element is and, once more, cyclicity of implies that is completely distributive. It was still open the possibility that 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 (with atoms ) and has the following multiplication table:
It is verified that is the only non-cyclic element and that, at the same time, it is the only dualizing element. For the quantale we shall see that existence of a dualizing element again implies complete distributivity of (and therefore existence of a cyclic and dualizing element).
As this might be of more general interest, we are going to investigate how divisions and by an arbitrary act on the and . To this end, we start remarking that Raney’s transforms intervene in the formulas for computing left adjoints of the maps and .
Lemma 3.1.
The functions and have both a left and a right adjoint. Namely, for each , , and , the following relations hold:
Using the relations stated in Lemma 3.1 computing divisions becomes an easy task.
Lemma 3.2.
For each , , and , we have
Proof.
We compute as follows:
where we have used Lemma 2.5. Verification that is similar. The other two identities are verified as follows:
and, dually, | ||||
∎ |
The relations in the following proposition are then easily derived.
Proposition 3.3.
The following relations hold:
From Proposition 3.3, it follows that
If is invertible, then its inverse is also its right adjoint. From the uniqueness of the right adjoint it follows that is inverted by its right adjoint . Thus, is invertible if and only if is invertible, in which case we have and , since both and are bicontinuous. In some important case, the expressions exhibited in Proposition 3.3 simplify:
Corollary 3.4.
If (resp., ), then
These relations hold as soon as either or is invertible.
Example 3.5.
If is a completely distributive lattice, then the relation holds, by Raney’s Theorem [14]. Let , then is invertible, since it is the identity. Necessarily, we also have and therefore:
Theorem 3.6.
If is dualizing, then and are inverse to each other and is completely distributive.
Proof.
If is dualizing, then, for all ,
and since both and are injective, then and . Thus, and are inverse to each other. Remark now that , since . Then , since is an ideal of . Then is completely distributive by Raney’s Theorem. ∎
It is not difficult to give a direct proof of the converse, namely that if is invertible, then is dualizing. We prove this as a general statement about involutive residuated lattices. Notice that the map sending to is definable in the language of involutive residuated lattices, since , where is the canonical cyclic dualizing element of (if 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 , is dualizing if and only if is invertible.
Proof.
If is dualizing, then , for each . Using well known identities of involutive residuated lattices, compute as follows:
Letting , then . We derive similarly, from which it follows that is inverted by .
Conversely, suppose that is invertible, say . It immediately follows that . Again, for each ,
and then, dualizing this relation, we obtain
Since always hold, we have . The identity is derived similarly. ∎
Example 3.8.
If , then is dualizing if and only if it is invertible. Indeed, is dualizing iff is invertible iff is invertible. Now, if is invertible, then it is continuous and ; therefore is invertible. Similarly, if is invertible, then it is continuous and ; therefore is invertible.
Example 3.9.
Consider a poset , the complete lattice of downsets of , and recall that is completely distributive. The quantale is isomorphic to the quantale of weakening relations (profuctors/bimodules) on the poset . These are the relations such that , , and imply (for all ). Thus, weakening relations are downsets of and the bijection between and goes along the lines described in previous sections, since
Explicitly, this bijection, sending to , is such that
(4) |
where, for , . We have seen that dualizing elements of are in bijection with automorphisms of which in turn are in bijection with automorphisms of . Given such an automorphism, we aim at computing the dualizing weakening relation corresponding to this automorphism. To this end, let us recall that, when is completely distributive, is the unique non-trivial cyclic element. Observe that since , is also dualizing and that
For , recall that . We use the relations in (4) to compute the dualizing element of corresponding to an invertible order preserving map , as follows:
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 and . By Fact 1.3, this question amounts to understand whether is autodual.
Let us discuss the case when is a finite lattice. We use for the set of join-irreducible elements of and for the set of meet-irreducible elements of . The reader will have no difficulties convincing himself of the following statement:
Lemma 4.1.
A map is meet-irreducible if and only if it is an elementary tensor of the form with and .
The following statement might instead be less immediate:
Lemma 4.2.
For each and , the map is join-irreducible.
Proof.
Let and , and let us use to denote the unique upper cover of . Suppose that . By evaluating the two sides of this equality at , we obtain and therefore for some . If , then . Suppose now that , so and . Observe also that , since . Then , so . We have argued that , for all , and therefore that . ∎
Theorem 4.3.
If is a finite lattice and is autodual, then is distributive.
Proof.
If is invertible, then restricts to a bijection , so these two sets have same cardinality. For and , the as well as the elementary tensors are pairwise distinct. Therefore, we have
and . That is, the elements are all the join-irreducible elements of and therefore the set generates under joins. It follows that and that 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 that is not completely distributive and such that is autodual. It is clear, however, that in order to construct such a fancy lattice, properties of bimorphisms need to be investigated. What are the properties of a bimorphism forcing to be completely distributive when is surjective? Taking the bimorphism as example, let us abstract part of Raney’s Theorem:
Proposition 4.4.
Let be a bimorphism such that for each , the image of under is a finite chain. If belongs to the image of , then is a completely distributive lattice.
Proof.
Let such that . We aim at showing that , with the index ranging on choice functions ( is a choice function if , for each ). Since , we also have and therefore, in order to achieve our goal, it will be enough to show that for each , if , then . Let be such that , fix , and observe then that
since . Since the set is finite and directed (it is a finite chain), it has a maximum: there exists such that . It follows that
since . By letting vary, we have constructed a choice function such that , and consequently . ∎
Bimorphisms satisfying the conditions of Proposition 4.4 might be easily constructed by taking , (resp., and ), and defining then
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 .
References
- [1]
- [2] Michael Barr (1979): -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.