ANSWERS TO SOME QUESTIONS ASKED BY STUDENTS IN MATH 245, taught in Fall 2017 from my notes "An Invitation to General Algebra and Universal Constructions", /~gbergman/245. These are my responses to questions submitted by students in the course, the answers to which I did not work into my lecture, but which I thought might be of interest to more students than the ones who asked them. ---------------------------------------------------------------------- You ask whether the two downward arrows in the diagram in Definition 3.1.3 represent the same map h, or whether there is a distinction between the map regarded as a set map and as a group homomorphism. They do indeed represent the same map h. Since a group homomorphism is a set-map that satisfies appropriate conditions with respect to the group operations, the same map can be both a set-map and a homomorphism. Its occurrence in the triangle is based on the fact that it is a set map, and so can be composed with the horizontal arrow of that triangle, and the result equated with the diagonal arrow, giving us a triangle of set maps. Its appearance on the right, on the other hand, is based on its being a group homomorphism. It is unique for having the combination of the set-theoretic and group-theoretic properties shown by these two appearances. ---------------------------------------------------------------------- You ask about the remark after the last diagram in section 3.1, that since any two free groups on X are "essentially" the same, one often speaks of "the" free group on X. Whether it is reasonable to use "the" in such a situation depends on what one is focusing on. If one is interested in the group-theoretic properties of the group, these are the same for any two isomorphic groups, so one thinks of all free groups on X as being "the same for the purposes of discussion", and it is natural to use "the". On the other hand, if one is interested in questions such as whether the set X is actually contained in F, or mapped into it by a map u which is not an inclusion map, then this will differ for different realizations of the free group structure, and one would call each of them "a free group on X". Regardless of one's point of view, it is not quite precise to say "the" free group, since the different objects with that universal property are not literally the same. But it is OK to speak a little imprecisely as long as the speaker and hearer understand what is actually meant. ---------------------------------------------------------------------- Concerning the proof of Proposition 4.6.5, you say > I don't quite see what the purpose of constructing this set A is ... I hope that what I said in class answered your question. Though it is easy to see that every element of the universal group generated by images of G and H can be written in the form (4.6.4), and to define operations on such expressions, it is not easy to show that these operations satisfy the group axioms; equivalently, that the group identities, together with the relations satisfies by our generators, don't somehow imply equality among two different expressions (4.6.4). But by finding a group of permutations of a set on which all those relations are satisfied, and in which different expressions (4.6.4) yield different permutations, we get this conclusion, and hence have the desired description of the "coproduct" of the groups G and H. ---------------------------------------------------------------------- You ask whether Lemma 4.17.3 can be proved without the Hausdorffness condition on K. No. For example, let K be any set, given with the topology under which the closed subsets are K itself and its finite subsets. Then any infinite subset of K is dense. Hence if we let X be any infinite set, we can map it into topological spaces K constructed as above so as to have dense image, though there is no upper bound on the cardinalities of such spaces K. ---------------------------------------------------------------------- In connection with the paragraph before Definition 5.1.4, which notes that partially ordered sets are not algebras in the sense used in this class, you ask "Could we not get around this issue by concentrating on relations instead of operations?" Yes; but then objects such as groups, rings, etc. would have to be described as having certain relations R subject to the condition that for every x and y there exists a unique z such that (x,y,z)\in R; and results on free groups, rings, etc., groups, rings, etc. presented by generators and relations, and so on, would all have to be formulated in terms of relations with this particular property. So to create a comfortable context for studying these things, I have chosen to work with algebras, defined by sets with operations, and to consider more general relational structures to be nearby relatives which we visit when we need them. ---------------------------------------------------------------------- In connection with the suggestion preceding Exercise 5.6:2, that if you haven't seen proofs by Zorn's Lemma before you might look at such proofs in standard graduate algebra texts, or ask your instructor for some elementary examples, you ask for such examples. Most basic graduate-level algebra texts have many such proofs, but it takes some work to find where they are. One text where one can find these by an online search is Hungerford's "Algebra". If you go to https://books.google.com/books?id=e-YlBQAAQBAJ and type "zorn" in the box saying "Search inside", you will get 23 results, and turning to the corresponding pages in a physical copy of the text, you can find the proofs in question. However, I suspect you can do *most* of Exercises 5.6:2-5.6:14 without looking at additional examples. It is unfortunate that I began that string of exercises with 5.6:2, which, in its present form, does not make it at all obvious how Zorn's Lemma would be used. So look at the next few, and see whether you have better luck. ---------------------------------------------------------------------- You asked about my use of the phrase "(generally infinite)" in the fourth-from-last paragraph of chapter 6. The reason I put that phrase in was, of course, that we usually see the conjunction symbol used to connect two propositions, or, if several conjunction symbols appear, finitely many; and I wanted to make clear that the finiteness restriction that is automatic in such cases was not being assumed there. I could equally well have expressed this by writing "(possibly infinite)". In choosing to say "generally", I was implicitly assuming that in mathematics, infinite structures (such as the ring of integers, or the real numbers) are more often of interest, and finite structures a more special case. But for mathematicians who specialize in the study of finite objects (e.g., finite groups, finite lattices, etc.), the reverse is true. So there was no absolute justification for my choice of word. ---------------------------------------------------------------------- You ask about the relation between closed subsets, i.e., sets fixed under "**", where "*" is the operator in a Galois connection, and topological closure, such as comes up in Galois theory of infinite field extensions. Well, closure operations occur throughout mathematics, as the examples given in section 6.3 show. It happens that in one area, topology, one deals with a closure operation that is simply called "closure". This is perhaps what led you to look at the word "closed" in that way. I guess when a closure operator is not finitary, i.e., when the closure of the union of a chain of subsets can contain more elements than the union of the closures of the sets in the chain, then the way these new elements arise is often given by a topological closure. So when this happens, as in the field extension case you mentioned, topology will be involved. I don't know whether every closure operator on a set can be decomposed somehow into a finitary closure operator and a topology ... . But anyway, as Exercise 6.5:3 shows, every closure operator on a set can be looked at as arising from some Galois connection, but as Exercise 6.3:17(i) shows, those closure operators that come from topologies alone are highly restricted. ---------------------------------------------------------------------- You ask, concerning the first sentence of Example 7.5.9, where I refer to "objects of this sort", whether this means sets with operations of a given signature; and also whether there are any restrictions on the language in which the propositions comprising T are expressed. My intent was to be very general. Rereading that Example, I think that in the first line, after "Let S be a set of mathematical objects", I ought to add, "of a given sort (e.g., of groups, of positive integers, of topological spaces, ...)". Then, hopefully, the sense of "of this sort" later in the sentence will be clear. The propositions can likewise be in any language -- all that is needed to get a Galois connection is that for each object s\in S and each proposition t\in T, it makes sense to say whether s satisfies t. The point of this Example is show that the basic ideas of studying sets of mathematical objects determined by propositions that they satisfy, and sets of propositions determined by the mathematical objects that satisfy them, can be looked at as an example of the concept of Galois connection. Of course, when one wants to study such a situation further, one will generally want restrictions of one sort or another. For instance, the paragraph following Exercise 6.3:8 discusses certain properties that follow if the language in question includes the operators "or" and "and" (interpreted in the standard way), and T is closed under those operators. The material of Chapter 9 concerns the situation you asked about, where S consists of all sets with operations of a given signature. There, T consists of all identities in those operations (but does not contain expressions formed with "or". Whether we allow "and" (written "\wedge") doesn't really matter, because, e.g., if t_1, t_2, t_3 are identities, then the set of algebras satisfying all members of, say \{t_1\wedge t_2, t_3\} is also the set satisfying all members of \{t_1, t_2, t_3\}.) ---------------------------------------------------------------------- You ask about the meaning of "isomorphic" in exercise 7.2:1. An isomorphism i: C --> D of categories, like an isomorphism between other mathematical objects, means a way of mapping the elements comprising C bijectively to the elements comprising D so that the structure is exactly preserved. In detail, this means a bijection i_{Ob} of object-sets, and for each pair of objects X, Y \in Ob(C), a bijection i_{X,Y}: C(X,Y) --> D(i_{Ob}(X), i_{Ob}(Y)), which respects composition (and identity morphisms, though that follows from the other conditions). ---------------------------------------------------------------------- In connection with Definition 7.4.4, you asked whether the concept of an object such as a group G being U-small referred to the whole structure, e.g., the group (|G|, \mu_G, \iota_G, e_G), being a small set (a member of U), or just the set |G|. Those two conditions are equivalent, since once |G| is U-small, the map \mu_G, as a map |G| x |G| --> |G|, and hence a subset of (|G| x |G|) x |G|, will also be U-small, and likewise for \iota_G and e_G; hence the 4-tuple (|G|, \mu_G, \iota_G, e_G) will be U-small. Anyway, the intended meaning (in case there are cases where the equivalence is evident) is that whole structure (|G|, \mu_G, \iota_G, e_G) is a U-small set. ---------------------------------------------------------------------- > ... you mention that in categories there is no real way to > distinguish isomorphic objects. Is there any danger, then, of > always looking at the skeleton category, which identifies > isomorphic objects? ... From an abstract point of view, no. But in specific cases, it could well be confusing. For instance, suppose we are looking at the category Set, and are considering the ordinals as objects of that category. Then the countably infinite ordinals (of which we know there are uncountably many) all become isomorphic. Suppose we go to a skeleton category of Set in which the only representative of that isomorphism class is \omega. We could still think about the chain of inclusions of countable ordinals by taking a retraction of Set to a skeleton, and looking at the images of the inclusion maps \alpha --> \beta among the countably infinite ordinals as certain endomorphisms of the object \omega to which they have all been retracted; and we could, for instance, describe the first uncountable ordinal as the direct limit of that chain of endomorphisms of \omega. But it would be much easier to think of the category Set in which the countable ordinals are separate objects. ---------------------------------------------------------------------- You ask about (co)products that are preserved under all functors (possibly assuming that the codomain category has coproducts). Well, this will certainly be true of the coproduct of any 1-object family! My guess is that it will not be true in any other cases; but I don't know. ---------------------------------------------------------------------- > As is mentioned in Section 8.1, all universal objects are initial > objects in some category. But to show this, don't we need some > abstract way of constructing a category in which a universal object > is initial? The examples given in 8.1 seem rather ad hoc, as they > depend on the specific universal object being considered. Well, we don't yet have a general definition of "universal object"! In Chapter 4 we displayed a large bunch of interesting constructions which seemed to have a feature of "universality" in common; and in section 7.8 we found that some families of these could be gathered under common category-theoretic descriptions. In the first half of Chapter 8 we will look at these more systematically. From the point of view of section 8.1, "universal object" might be defined as "structure corresponding to an initial object in some category"; in later sections, we look at specific sorts of properties that can be described as in this way, but that have more natural descriptions in terms of other categories, and how to pass between the two sorts of description. ---------------------------------------------------------------------- Regarding Definition 8.3.9 you asked, > If we have three functors F,G,H, where G is a right adjoint of F > and a left adjoint of H, is there anything interesting we can say > about F and H? For instance, in the example you give with products, > coproducts, and the diagonal functor, the product and coproduct > functors are essentially dual to each other. That the product and coproduct are "essentially dual" just means that they are obtained by dual constructions -- left adjoint and right adjoint -- from the diagonal functor, which has no left/right asymmetry. It refers to the way we look at them, not to properties that they will have in a particular category. But the question of whether, when a functor has a left and a right adjoint, those two adjoints have properties relating them, is interesting. I don't know the answer, but I would guess that some properties of G, concerning what distinctions it preserves and what distinctions it loses, would force some dual properties on F and H, which they would share. But I don't know anything concrete about this. Alexey Zoubov has pointed out to me a result along these lines: F is full and faithful if and only if H is full and faithful. This is Lemma 1.3 in "Exponentiable morphisms, partial products and pullback complements" by Roy Dyckhoff and Walter Tholen, Journal of Pure and Applied Algebra, 49(1987) 103-116, https://doi.org/10.1016/0022-4049(87)90124-1 ---------------------------------------------------------------------- I meant to work the answer to your question into my lecture, but didn't get to it. You asked whether there is a notion of convergence associated with the direct limit. I think that the main connection is conceptual -- that if X is the direct limit of a family of objects X_i, then the successive X_i are "more and more like X". E.g., in the case of a direct limit of groups, they in general have more and more of the elements that will show up in X, and these satisfy more and more of the relations that are satisfied there. But it may be possible to turn this into topological convergence. If we define a language in which there are symbols for all the elements of X, and for the kinds of relations that these can satisfy, then we might define a topology on any set of objects of the indicated sorts, some elements of which are labeled with some of the element-symbols, taking for a subbasis of open sets the sets of objects characterized by having (or not having) an element symbolized by each symbol, and satisfying (or not satisfying) a given relation on such elements; and then I think the direct limit X would be the topological limit of the system of "points" X_i. (In an abstract category, we might do something similar using the existence of maps from various objects in the category satisfying various composition relations; but I haven't thought this through.) ---------------------------------------------------------------------- You ask about the second sentence after Theorem 8.8.9, saying that if we assume that all functors from D to C have colimits, then the isomorphism between the left and right expressions in (8.8.10) becomes a case of Theorem 8.8.3. I see that that does need a bit of explaining! The first assertion of Theorem 8.8.3 says that if F: C --> D is a left adjoint functor, and S: E --> C has a colimit, then lim_E FS = F(lim_E S). (I will use "lim" in this e-mail for the colimit symbol, "lim" with an arrow "-->" under it.) In the application we want to make, the roles of C and D in that theorem are taken by C^D and C respectively, the role of F by lim_D, and the role of S by B(-,-), regarded as a functor E --> C^D. ("E" is the one symbol that translates to itself!) Then the formula lim_E FS = F(lim_E S) takes the form lim_E lim_D B(-,-) = lim_D lim_E B(-,-). (When I have time, after the end of the semester, I should add a clarification of this to the online revised version of the text.) ---------------------------------------------------------------------- > In proposition 8.9.3, it says that "In Set, direct limits commute > with finite limits." This commutation should be understood as 'up to > isomorphism', correct? Well, limits and colimits are only defined up to isomorphism -- they are objects with universal properties, and if one object has the universal property, then anything isomorphic to it does. It is true that if we have specific constructions of the limit and colimit in one order and in the other, then the "commutativity" statement means that the two resulting objects are isomorphic (by an isomorphism that respects the appropriate structure). See Definition 8.8.1, last sentence and then first two sentences. ---------------------------------------------------------------------- Pointing to the contrast between Exercises 8.10:6 and 8.10:7, you ask about general results on universal constructions in classes of algebras with large sets of operations. I don't know whether such results have been looked at. In General Algebra, it is most natural to consider algebras with small sets of operations. Complete lattices are most naturally thought of as having meets and joins of arbitrary subsets; but operations on subsets don't fit the techniques of General Algebra. The best way one can accommodate them is to treat them as "large" families of operations on tuples of elements. As such, they are more like the kinds of structures studied in General Algebra than sets with operations on subsets would be, but they still push the envelope. I think "random" sorts of algebras with large classes of operations will "almost never" have free objects; it is only certain sorts that arise in special ways -- such as complete semilattices -- that have these. ---------------------------------------------------------------------- You asked about the fact, mentioned following Exercise 9.6:10, that the concept of "heap", developed in that exercise, had being rediscovered many times, and what could have led to that repeated rediscovery. I think that two contrasting aspects of the concept of heap are relevant: That it is fairly natural and moderately useful, but that it does not invite study for its own sake, since the isomorphism classes of heaps (other than the empty heap) are completely determined by the isomorphism classes of the corresponding groups. Naturalness and usefulness leads people to discover the concept, but since the theory completely reduces to that of groups, not many works get written about them, and no one ends up specializing in them; so few people hear about the concept, and if they need them, they often end up reinventing them. When I rediscovered them I called them "isogroups"; see p.60 of https://link.springer.com/article/10.1007%2FBF02188011 , line after display (5). After that appeared, someone told me that they had already been defined. The same situation -- being natural and useful, but having a theory that essentially reduces to other theories -- is also true of preorders. I don't know whether they too have been rediscovered several times. If not, perhaps the situations in which they come up are sufficiently widespread that once a name was assigned to the concept, enough people heard of it to prevent the "rediscovery" syndrome. ---------------------------------------------------------------------- You write > In the middle of section 9.7, it is said there is a connection > between Lie algebras and Lie groups. Does this connection allow us > to identify an algebra with a (possibly) group or vice-versa? As the sentence preceding Exercise 9.7:6 suggests, it is a many-one relationship: The many different Lie groups that look alike in a neighborhood of the identity have the same Lie algebra. For each finite-dimensional Lie algebra L over the real numbers, there is a unique *simply connected* Lie group G having L as its Lie algebra. Other Lie groups having L as their Lie algebra can be obtained by dividing G by a discrete subgroup, and/or throwing in more connected components. For a trivial case: if L is the 1-dimensional Lie algebra, which by (9.7.3) has zero bracket operation, then G is the real line R, while other Lie groups with the same Lie algebra include the circle group R/Z, and various groups with R or R/Z as the connected component of the identity element. ---------------------------------------------------------------------- You ask why the assumption that C is a variety of algebras is needed for condition (iii) of Definition 10.3.5 to be equivalent to the other two. If C is an arbitrary category with appropriate direct products as in the first sentence of the definition, then "elements of A" and "relations" satisfied by such elements aren't meaningful. One might, of course, ask whether some weaker assumptions can be made which would make those concepts meaningful. One can certainly do that. If the conditions are too weak, the concepts might be meaningful but the the equivalence could fail. For instance, if C is the category of all finite groups, one can still speak of elements and relations, but a given system of elements and relations might not determine a finite group so such X and Y might not determine a representing object. (E.g., one element and no relations give in the variety of all groups the infinite cyclic group, but don't determine any object of the category of finite groups.) There are, however, conditions weaker than that C be a variety of algebras, but strong enough to make the equivalence hold. But to develop these would require that we introduce further concepts, and for simplicity, I have restricted the main focus of the text to varieties of algebras. ---------------------------------------------------------------------- You ask about "a nice description" of the left adjoints of the representable functors Monoid --> Monoid (section 10.6). Well, they all have the nice formal description as functors associating to each monoid M a monoid gotten by attaching together a bunch of copies of the representing monoid R, indexed by the elements of M, with relations determined by the relations of M and the comultiplication of R. But what the result looks like for particular comonoids R can be complicated. For the simplest interesting case: The left adjoint of the functor associating to a monoid its group of invertible elements is the functor associating to a monoid M the result of adjoining an inverse to every element. The next-simplest interesting case is the left adjoint of the functor associating to M the monoid of pairs (a,b) in M with ab=e. This adjoint adjoins to M, for every a\in M, an element a' such that aa'=1, in such a way that the map a |-> a' reverses the order of multiplication. Note that if in M one has ab=e, then in the new monoid, b will have both the left inverse a and the right inverse b', so it becomes invertible, so its left inverse a also becomes invertible. So the elements that were 1-sided invertible M all become invertible in the new monoid; hence the submonoid generated by those previously-1-sided-invertible elements is embedded in a group. Elements that were not previously 1-sided invertible need not become invertible, though they do become 1-sided invertible. I haven't studied those functors systematically ... . ---------------------------------------------------------------------- You ask about the meaning of "subobject" in the paragraph following display (10.7.1). Good point. I guess I was implicitly relying on the fact that in most real-world mathematical contexts, an idempotent endomorphism of an object is a retraction to a subobject. This isn't a formal statement true or even meaningful in an arbitrary category; so that comment should be considered a heuristic observation, suggesting what we should look for. It is valid with object taken to mean "category", so it leads us to the right conclusion in this case. ---------------------------------------------------------------------- > Are there any examples of functors T: C^\op \to C such that T and > TT have left and right adjoints but TTT does not? I don't know. One conceivable way to approach the question would be to have C be a direct product of, say, 4 categories C_0 x C_1 x C_2 x C_3, and have T built out of functors C_i --> C_{i+1} (i=0,1,2), such that some sort of bad behavior is reached only when we carry C_0 into C_3. I don't have anything detailed in mind, but you might be able to go somewhere with the idea. ----------------------------------------------------------------------