Homomorphisms as structure-preserving maps

In summary, a homomorphism between groups is a function that "preserves" the product, a homomorphism between fields is a function that "preserves" both addition and multiplication, and homomorphisms are "structure-preserving" if they are defined in terms of a structure and also "preserve" what it means for the structure to be "preserved".
  • #1
Fredrik
Staff Emeritus
Science Advisor
Gold Member
10,877
422
Homomorphisms as "structure-preserving" maps

A function f between groups is said to be a homomorphism if it "preserves" the product in the sense that f(xy)=f(x)f(y). A function f between fields is said to be a homomorphism if it "preserves" both addition and multiplication in the sense that f(xy)=f(x)f(y), f(x+y)=f(x)+f(y). Homomorphisms are often described as "structure-preserving maps". We should be able to define homomorphisms as structure-preserving maps, if we define "structure" first, and also what it means for the structure to be "preserved". I suggest the following definition:

We define a structure as a 4-tuple (S,R,O,C), where S is a set, R is a set of relations on S, O is a set of operations on S, and C is a subset of S. The members of C are called constants. If (S,R,O,C) and (S',R',O',C') are structures, a map f:S→S' is said to be structure preserving if the following holds for all x,y,x1,...,xn in S:

a) If r is a relation in R such that (x,y) is in r, then there's a r' in R' such that (f(x),f(y)) is in r'.
b) For each n=1,2,..., if m is an operation of arity n in O, then f(m(x1,...,xn))=m(f(x1),...,f(n))
c) If x is in C, then f(x) is in C'.

(A relation of arity n is a subset of Sn. An operation of arity n is a function from Sn into S).

Unless I missed something essential (and I might have), this should take care of the definition of "homomorphism" for groups, rings, fields, topological spaces and a lot more, but not for metric spaces, vector spaces, manifolds and a lot more. A metric space involves a function that isn't an operation. A vector space also involves a function (scalar multiplication) that isn't an operation, and it also involves a previously defined structure (a field). And manifolds...uh...I don't even want to think about it.

What I'd like to know if there is a more general definition that covers these other things as well. I know the term homomorphism is used for vector spaces, but I don't think I've heard it used for metric spaces. I've heard people say that diffeomorphisms are isomorphisms between manifolds, but perhaps they just meant isomorphisms as they are defined in category theory (where they don't have to be structure preserving).

By the way, I have studied some mathematical logic since the other discussion, so I'm now familiar with how structures/algebras are defined in model theory/universal algebra. I understand the reason why texts on those subjects talk about relation symbols instead of relations and so on, and I feel that we don't need to do that here, as long as we're just looking for a single definition of homomorphism.
 
Last edited:
Physics news on Phys.org
  • #2


You should look into category theory
 
  • #3


I'm familiar with the basic definitions of category theory, so I know that it doesn't answer my question. It can't tell us which functions are the homomorphisms of a given class of structures (e.g. groups). Instead, the choice of what functions to call homomorphims (or just morphisms) is part of what defines a category. For example, category theory allows us to define a category that consists of the class of vector spaces and all functions between vector spaces that satisfy f(x+y)=f(x)+f(y) for all x,y. This defines a category of vector spaces, just not the standard one, and its isomorphisms are group isomorphisms, not what we'd normally consider vector space isomorphisms (i.e. linear bijections).
 
  • #4


If you have any prior notion of "object with structure" and "structure-preserving map", you can collect all instances into a category. Then your prior notion of "object with structure" is now the same as the new notion of "object in that category", and similarly for maps.

You can accelerate thing by skipping the first step of trying to come up with a prior notion of structure. :wink:


The members of C are called constants.
A constant is a nullary operation -- i.e. an operation of arity 0. You don't need to treat them separately.


b) For each n=1,2,..., if m is an operation of arity n in O, then f(m(x1,...,xn))=m(f(x1),...,f(n))
m is not an element of O'; it is not an operation on S'.


c) If x is in C, then f(x) is in C'.
Merely mapping constants to constants is inadequate. e.g. in a variety of universal algebras with two constant symbols 0 and 1, the function associated to a morphism must actually satisfy f(0)=0 and f(1)=1.



Fields: Their characterization is essentially non-algebraic. I suppose you luck out because you don't care about characterizing them, and the notions of ring homomorphism between fields and field homomorphism between fields are equivalent.

Topological spaces: You have types of stuff: points and open sets. Morphisms are awkward because the induced functions work in both directions: points are mapped forwards, but open sets are mapped backwards. You could switch to the notion of Locale so that you have only one type of stuff, but the function associated to a morphism still works backwards.

Homotopy types: The homotopy category is not concrete -- roughly speaking its objects cannot be represented as "sets with structure", no matter clever a notion of structure you might come up with.


A diffeomorphism is indeed an isomorphism in the category of smooth manifolds. (Or, more verbosely, the category of smooth manifolds and smooth maps) The reason we have many words like "diffeomorphism", "isometry", "homeomorphism" and so forth is that we are frequently interested in several categories at once. e.g. when studying Riemann manifolds, we might be interested both in smooth metric-preserving maps as well as plain ordinary smooth maps, and from time to time even merely twice-differentiable maps!
 
  • #5


Hurkyl said:
A constant is a nullary operation -- i.e. an operation of arity 0. You don't need to treat them separately.
I know. Also, every operation of arity n is a relation of arity n+1, so we don't even have to treat operations separately. I chose to treat them separately to make the definition look more like the definitions in the mathematical logic books I'm reading and the definition in this book on universal algebra that I downloaded. (Except that they don't include any relations).

Hurkyl said:
m is not an element of O'; it is not an operation on S'.

Merely mapping constants to constants is inadequate. e.g. in a variety of universal algebras with two constant symbols 0 and 1, the function associated to a morphism must actually satisfy f(0)=0 and f(1)=1.
D'oh. That first one was a major brain fart by me, and the second was was pretty silly too. I'm glad you caught them, because now I finally see why these books talk about relation symbols instead of relations and so on. It's not just because it has to be that way when we're dealing with first-order formal languages. It's also needed for the definition of homomorphisms that I'm trying to make. New attempt:

I define a structure to be a triple (S,R,I) where S is a set, R is a set of relation symbols such that the arity of the symbol is explicitly included in the notation (e.g. r(n) is a relation symbol with arity n), and I is a function that assigns a relation on S of the appropriate arity to each relation symbol.

Suppose that (S,R,I) and (S',R,I') are two structures with the same relation symbols. A function f:S→S' is said to be a homomorphism if

[tex](f(x_1),...,f(x_n))\in I'r^{(n)} \Leftrightarrow (x_1,...,x_n)\in Ir^{(n)}[/tex]

for each relation symbol r(n) of arity n and for each n=1,2,... (I think this would be a bit clearer if we treat operations and constants separately).

Hurkyl said:
Fields: Their characterization is essentially non-algebraic. I suppose you luck out because you don't care about characterizing them, and the notions of ring homomorphism between fields and field homomorphism between fields are equivalent.
That's another good catch. You're quite useful. :smile: Now I'm going to have to think about whether my (new) definition allows e.g. a non-abelian division ring to be isomorpic to a field (i.e. abelian division ring).


Hurkyl said:
Topological spaces: You have types of stuff: points and open sets. Morphisms are awkward because the induced functions work in both directions: points are mapped forwards, but open sets are mapped backwards. You could switch to the notion of Locale so that you have only one type of stuff, but the function associated to a morphism still works backwards.
I was thinking that S is the set and that each subset of S is a unary relation on S. You're right about the homomorphisms. My improved definition doesn't work. This makes me wonder why homeomorphic topological spaces are considered equivalent, instead of topological spaces that would be isomorphic according to my definition.

I guess what we really want is a notion of equivalence that ensures that every theorem that we derive using the properties of a specific "structure" (a concept that I still haven't defined satisfactorily) will be guaranteed to hold for any structure that's "equivalent" to the one we used.

If "structure" doesn't apply to all these things that are defined by "a set and then some", like groups, fields, topological spaces, Hilbert spaces and manifolds, then I wonder, isn't there some other term that can be used? Aren't they all special cases of a single more general concept?
 
Last edited:
  • #6


It's OK for other people to contribute here too. :smile: (Not that I don't appreciate Hurkyl's contributions. I always do). Let's focus on one specific detail to make things easier. Why is the category of topological spaces that has continuous maps as morphisms more interesting than the category of topological spaces that has open maps as morphisms? (f:X→Y is said to be open if f(E) is open for every open E).

I've been thinking that we want the morphisms to have the property that the corresponding isomorphisms tell us when two structures are equivalent in the sense that any theorem that holds for a specific structure also holds for all isomorphic structures. Does that hold for topological spaces when the isomorphisms are continuous bijections but not when they're bijective open maps?
 
  • #7


Fredrik said:
Does that hold for topological spaces when the isomorphisms are continuous bijections
I assume you meant "homeomorphism" and not "continuous bijection"?


Incidentally, all isomorphisms in the category of topological spaces and continuous maps are also open.
Additionally, all isomorphisms in the category of topological spaces and open maps are also continuous.
 
  • #8


Hurkyl said:
I assume you meant "homeomorphism" and not "continuous bijection"?
Yes. I assumed those two were the same, but I guess I should have thought it through.

Hurkyl said:
Incidentally, all isomorphisms in the category of topological spaces and continuous maps are also open.
Additionally, all isomorphisms in the category of topological spaces and open maps are also continuous.
So...the isomorphisms of the first category are the same as the isomorphisms of the second category. I guess that solves that problem.

I'll try to prove the things you wrote here as an exercise. Thank you again for your comments.

Edit: I'm too lazy to try hard to find an example of a continuous bijection with an inverse that isn't continuous, but unless I made a mistake, I have verified the other details.
 
Last edited:
  • #9


I think the only way we can have a useful general theory of 'structure preserving maps' is if that theory talks only about properties that all structure preserving maps possess, and this is the case with category theory as far as I can tell.

To me, this seems to imply that this general theory cannot define a structure preserving map as a specific type of map between sets, but only characterize them by their structure-independent features (and therefore the theorems of category theory apply to all 'structures'). For example, we require that composition of structure preserving maps φ : A → B and χ : B → C must be a structure preserving map ψ : A → C and that there exists an 'identity structure preserving map' and so on.

The reason this is powerful is kinda the same as the reason the notion of 'topological space' is powerful. It isolates and characterizes the idea of 'continuity' in a terminology which ensures that no irrelevant features of specific examples enter the description. To study continuous functions on Euclidean space, we don't need much of the information contained in the euclidean metric. So the old definition of continuity (For every epsilon, there exits a delta such that...) is 'redundant'. The modern definition of continuity (If Δ is open, so is φ-1[Δ]) uses a terminology which does not refer to the metric at all, and applies to any topological space.
 
Last edited:
  • #10


dx said:
I think the only way we can have a useful general theory of 'structure preserving maps' is if that theory talks only about properties that all structure preserving maps possess, and this is the case with category theory as far as I can tell.
But category theory doesn't have a concept of "structure preserving" as far as I know. For example, we can define a category of groups with morphisms that don't "preserve" group multiplication.

dx said:
To me, this seems to imply that this general theory cannot define a structure preserving map as a specific type of map between sets, but only characterize them by their structure-independent features (and therefore the theorems of category theory apply to all 'structures'). For example, we require that composition of structure preserving maps φ : A → B and χ : B → C must be a structure preserving map ψ : A → C and that there exists an 'identity structure preserving map' and so on.
I would like these things to follow logically from a general definition of "structure preserving". And I think my second attempt in this thread works (although it would have been clearer if I had mentioned operation symbols and constant symbols explicitly). It seems to work at least for groups, rings (and therefore also fields), modules (and therefore also vector spaces), and topological spaces. I wonder if we can also make it work for metric spaces, manifolds, fiber bundles... I have to get some sleep, but I'll think about that tomorrow.
 
  • #11


Fredrik said:
Now I'm going to have to think about whether my (new) definition allows e.g. a non-abelian division ring to be isomorpic to a field (i.e. abelian division ring).
It doesn't. So that's one less thing to worry about.
 
  • #12


Fredrik said:
But category theory doesn't have a concept of "structure preserving" as far as I know. For example, we can define a category of groups with morphisms that don't "preserve" group multiplication.

In category theory, the 'structure' of the objects in a category is completely encoded in the morphisms and their compositions. So you can't say that the morphisms 'don't preserve the structure', since they define what the 'structure' is in that particular category. For example, consider a topological group. If this is viewed as an object in the category of topological spaces, there is no need for the morphisms to preserve the group structure.

I would like these things to follow logically from a general definition of "structure preserving"

Consider the definition of a topoligical space. The axioms that characterize open sets do not follow from a general definition of 'open set'. This reminds me of something that Niels Bohr said: 'A mutually exclusive relationship always exists between the practical use of any word and attempts at its strict definition'.
 
  • #13


Fredrik said:
But category theory doesn't have a concept of "structure preserving" as far as I know. For example, we can define a category of groups with morphisms that don't "preserve" group multiplication.
Then the structure that is being preserved by the morphisms isn't the group structure. :wink:


One nice concrete1 example of a category is the category of natural numbers and matrices. The objects of this category are natural numbers, and they are only really there to serve as a reminder of when we're allowed to multiply two matrices.

But if you really want to, you can think of the arrows as structure-preserving maps. What structure? The one implicitly defined by matrix arithmetic, of course! And we can use matrices to describe "elements" and even talk about subobjects (via the rowspace of a matrix).

Of course, this category is a concrete category; it's equivalent to the category of finite-dimensional vector spaces and linear transformations. The forgetful functor sends a natural number n to the set Rn, and it sends matrices to the appropriate set functions.


1: Meant in the natural language sense.
 
  • #14


dx said:
In category theory, the 'structure' of the objects in a category is completely encoded in the morphisms and their compositions. So you can't say that the morphisms 'don't preserve the structure', since they define what the 'structure' is in that particular category. For example, consider a topological group. If this is viewed as an object in the category of topological spaces, there is no need for the morphisms to preserve the group structure.
Hurkyl said:
Then the structure that is being preserved by the morphisms isn't the group structure. :wink:
OK, that makes sense. But it makes me wonder if this is a useful point of view. The reason why the explicitly defined "structure-preserving maps" are significant is that they give us a concept of "isomorphism" that ensures that theorems derived using the properties of a specific structure (e.g. SO(3)) holds for all structures that are isomorphic to it (e.g. SU(2)/Z2). That can be proved as a metatheorem. Is there a corresponding metatheorem about the isomorphisms we end up with when we define the structure implicitly by a choice of arrows, and what exactly does it say?
dx said:
Consider the definition of a topoligical space. The axioms that characterize open sets do not follow from a general definition of 'open set'.
I don't really understand this comment. I haven't suggested that they should. What I have suggested is that the appropriate concept of "isomorphism" should come from taking the morphisms to be (explicitly) structure-preserving, i.e. in this case they should be open maps, not continuous maps. However, it also turns out that both choices give us the same isomorphisms.
 
Last edited:
  • #15


Fredrik said:
I'm too lazy to try hard to find an example of a continuous bijection with an inverse that isn't continuous, but unless I made a mistake, I have verified the other details.
The standard example is
[tex][0,2\pi )\to S^1[/tex]
[tex]t\mapsto e^{it}.[/tex]
 
  • #16


Fredrik said:
I don't really understand this comment. I haven't suggested that they should. What I have suggested is that the appropriate concept of "isomorphism" should come from taking the morphisms to be (explicitly) structure-preserving, i.e. in this case they should be open maps, not continuous maps. However, it also turns out that both choices give us the same isomorphisms.

A topological space is defined as a set X together with a collection of subsets of X called its 'topology', such that

i.) If A and B are in the topology, then so is their intersection.

ii.) If Aλ (λ in Λ) are in the topology, so is their union.

iii.) X and {} are in the topology.

These properties of open sets are not derived from a definition of 'open set', but merely postulated. We have an intuitive idea of what an open set is: "a set which contains all points sufficiently close to to each of its points", and such intuitve ideas are always at the back of ones mind when the language of topological spaces is being used. Otherwise it would just be a formal play with symbols.

Similarly, we have an intuitive idea of what a 'structure preserving map' is, and from that idea we abstract the essential features so that they can be discussed independently of specific examples. An isomoprhism is a morphism A → B such that there exists a morphism B → A which composes with the original map to give the identity (both ways). This is a 'categorical definition' of an isomorphism, and does not require any concrete definition of what a 'morphism' is. It is taken to be a primitive notion, and enters only as an element of the Hom(_,_) sets.

Such categorical definitions are useful because they apply to all categories. For example, if you are able to find a categorical definition of the 'direct product' or 'direct sum' of two objects, that definition can be applied to any category, and therefore the language of category theory helps us find appropriate definitions. If you want to consider objects of the form of triples (S, R, I), then category theory will help you find a natural definition of a 'product' of such objects, etc.
 
Last edited:
  • #17


Fredrik said:
Is there a corresponding metatheorem about the isomorphisms we end up with when we define the structure implicitly by a choice of arrows, and what exactly does it say?
In the end, I bet it simply boils down to two things (which I'm not sure are actually different):
  • Theorem: If F is a functor from C to D, and f is an isomorphism in C, then F(f) is an isomorphism in D
  • For any diagram drawn in a category, we can create an 'isomorphic' diagram by replacing its objects with isomorphic ones (choosing a specific isomorphism), and composing the arrows with the isomorphisms and their inverses as appropriate.
(A functor is a "structure-preserving map" between categories, if you will. It only preserves structure up to natural isomorphism, though)

Example 1: "Cartesian product" is a functor from Set x Set to Set. It immediately follows that if X and X' are bijective sets, and Y and Y' are bijective sets, then X x Y and X' x Y' are bijective sets.

Example 2:: The "singular homology" construction defines functors on the category Top of topological spaces and continuous maps. It immediately follows that homeomorphic topological spaces have isomorphic singular homology groups.


Following up on that latter example: there is a relation on continuous maps called "homotopy". This relation is an equivalence relation. It is even a "congruence" on Top: f~g implies fh~gh and kf~kg. It turns out that that homotopic maps turn into equal maps after taking singular homology. There are a variety of other purposes for which the difference between homotopy and equality is insignificant.

Therefore, it is desirable to construct a new category hTop -- the homotopy category -- by modding out by this equivalence relation. Its objects are again topological spaces, but the arrows are now equivalence classes of continuous maps. Isomorphims in this category are homotopy equivalences. An isomorphism class of objects is a homotopy type.

So to reorganize the above data, homology is a property of homotopy types; the structure of "topological space" contains a lot of irrelevant information. Homology defines a functor on hTop, and the various other purposes I mentioned can be carried out in hTop rather than in Top.

So, in these cases, the topic of interest really is this category hTop (or other related constructions). Whatever 'structure' a homotopy type consists of, that's the structure we want to study.

The problem... hTop is not concrete. This is the example I gave earlier of a category that cannot be viewed as "sets with structure". There is a theorem that any functor U:hTop->Set must not be faithful. That means there are homotopy types X and Y, and a pair of distinct arrows f,g from X to Y with the property that U(f) = U(g).
 
  • #18


I should also add you can build a formal language out of diagrams rather than logical connectives as is usually done. I imagine that in such a language, preservation of logic by isomorphism is self-evident.

I didn't pay too much attention to it, but the one construction I saw was used to prove a cute little theorem: a property (described in this language) of categories is preserved and reflected by equivalence of categories if and only if the property can be described on a blackboard. (i.e. which amounts to never asserting as a conclusion that two objects must be equal. Isomorphic is still okay)

("preserved" means that if X has the property, then so does F(X). "reflected" means that if F(X) has the property, then so does X)
 
Last edited:
  • #19


I've attached a diagram that expresses the proposition "A pair of objects has a product".

A model in a category C of this diagram is a functor from the 2-point category into C (i.e. a pair of objects) with the following property:

There exists a functor from the category drawn in the second diagram that is compatable with the first one (i.e. every pair of arrows X <-- Z --> Y such that X and Y were the objects selected by the first functor) with the property that:

For every functor from the third diagram compatible with the second functor:

There exists a unique functor from the fourth diagram compatible with the third. (the triangles commute)



In this language, it's immediate that the property is invariant under isomorphisms; if I replace the two objects with isomorphic ones (and choose a specific isomorphisms), I can produce compatible natural isomorphisms on all of the relevant functors by swapping those two objects for their isomorphic copy, and composing the arrows appropriately.

It's also immediately evident that if XxY is a product of X and Y, then XxY is also a product of X' and Y'. (If X' and Y' are isomorphic to X and Y)
 

Attachments

  • foo.jpg
    foo.jpg
    5.4 KB · Views: 562
  • #20


Hurkyl said:
  • Theorem: If F is a functor from C to D, and f is an isomorphism in C, then F(f) is an isomorphism in D
I understood this part at least, but I had difficulties after that. I'll try again tomorrow.
 
  • #21


Oh, I've worked out something I wanted to mention earlier.


There is a particular kind of structure that can be defined for any small category: that of a category action, which is analogous to a group acting on a set.

A right action of the category C on a set X consists of:
  • An function "domain" that maps each element of X to an object of C.
  • For each arrow f : A --> B of C, there is a partial operation defined on the
    elements with domain B. (turning them into elements with domain A)
  • All of the axioms needed to play nicely with composition within C

Any object A of C can be used to create a set with a right action of C as follows:
  • The elements of this set are all arrows with codomain A.
  • The "domain" function is the domain operation on C
  • The action of an arrow is given by composition in C
(drawing pictures helps, I think)

Any arrow A --> B of C yields a "structure-preserving" homomorphism of right C-sets, turning the C-set defined by A into the C-set defined by B. (This homomorphism is again just composition)

Using this transformation, we can now view every object of the category as a set with structure. and the arrows as structure-preserving maps.


Frequently, we can get away by not taking all arrows with a given codomain, but instead those whose domain lies in some smaller set. (This trick allows us to do all of the above with many large categories)

(It's always fair to use a smaller set, but if it's too small, this representation doesn't work well; it can "forget" the difference between distinct objects and maps)

------------------------------------------------------------------------------

Interesting example: Consider again the category Mat (not a standard name) of natural numbers and matrices. In this example, it is good enough to use 1 as the domain of all of our elements.

Given any natural number n, I create a right Mat-set, which I will suggestively name Rn. It's elements are all arrows whose domain is 1 and whose codomain is n -- in other words, its elements are precisely the set nx1 column vectors.

Since 1 is the only object I'm using for domains, the only operations are those coming from functions 1 --> 1, which are 1x1 matrices; i.e. scalars. The action of a scalar? Right multiply the nx1 column vector by the 1x1 matrix.

Mat is a category with extra structure -- in this case, an addition operation between parallel arrows that distributes over composition. So this addition operation applies to our elements of Rn as well.

After putting together all of the category action axioms and additive category axioms, you'll find that this abstract construction has reproduced the explicit "vector space over R" structure on all of the standard finite-dimensional vector spaces Rn that we're familiar with.


It turns out we've actually reproduced the entire notion of "vector space over R" if we proceed further, but I didn't want to delve into too many details at once.
 
  • #22


Ooh, let's try this recipe on another category.

Let Man be the category of manifolds.
Let Euc be the full subcategory of Euclidean spaces. (its objects are the Rn, its arrows are all continuous maps between them)

Euc generates Man, so we can use it for our elements. Explicitly, a right Euc-action is a set X whose elements I will call "shapes", together with
  • A function X --> N. I will call this "dimension"
  • If f:Rm->Rn is continuous, and x is a shape of dimension n, then xf is a shape of dimension m.
  • The product is "associative": x(fg) = (xf)g, when defined.

A homomorphism of right Euc-sets is a function T with the property that
  • If x is a shape of dimension n, then so is T(x).
  • T(xf) = T(x)f whenever xf is defined

For any manifold M, there is a right Euc-set M whose shapes of dimension n are the continuous functions Rn --> M.


Unfortunately, I'm not sure if "homomorphism" and "continuous" coincide for manifolds. There might be too many homomorphisms, but then again manifolds are nice, and topology can surprise me.

However, I am pretty sure "homomorphism" and "continuous" coincide when we consider just the Euclidean spaces. But I don't know a direct proof -- I am invoking abstract nonsense to make that conclusion. (specifically, the Yoneda lemma)

If they don't coincide, there are notions of topology for categories (e.g. Grothendieck topology), and I am nearly certain an appropriate notion of "continuous right Euc-set" would ensure that homomorphism and continuous coincide.
 
Last edited:
  • #23


I haven't abandoned the discussion. I just feel that I need to understand categories a bit better before I continue, so I started reading a book about it.

By the way, a functor is supposed to be a pair of functions that take objects to objects and arrows to arrows, but the class of objects can be a proper class. So what do we mean by "function" here?
 
Last edited:
  • #24


Fredrik said:
By the way, a functor is supposed to be a pair of functions that take objects to objects and arrows to arrows, but the class of objects can be a proper class. So what do we mean by "function" here?

I'm not sure what the distinction between set and class is with regard to functions. Do we need a different notion of 'function' for classes? To me the phrase "F associates an object F(A) to the object A" seems unambiguous.
 
Last edited:
  • #25


dx said:
I'm not sure what the distinction between set and class is with regard to functions. Do we need a different notion of 'function' for classes? To me the phrase "F associates an object F(A) to the object A" seems unambiguous.
That's not a definition. It's just a comment that tells you how to think about a concept that's left undefined.

The definition goes like this: Suppose X and Y are sets and that [itex]G\subset X\times Y[/itex]. The triple [itex]f=(X,Y,G)[/itex] is said to be a function from X into Y if

(i) for each [itex]x\in X[/itex], there's a [itex]y\in Y[/itex] such that [itex](x,y)\in G[/itex].
(ii) [tex]\Big((x,y)\in G[/tex] and [tex](x,y')\in G\Big)\Rightarrow y=y'[/tex]

Alternatively, you can say that a subset [itex]G\subset X\times Y[/itex] is a function from X into Y if...the same two conditions are satisfied. This defines a function to be what someone who prefers the first definition would call the graph of the function.

I've been reading in Goldblatt's book today. I think he said that we can leave functions undefined instead of set membership. I understand how that makes sense if we want to use categories as the foundation of mathematics instead of set theory, but what if we just want to use category theory to organize and generalize mathematical concepts?
 
Last edited:
  • #26


The notion for function is the same for classes and sets; it's just that the graph of a function whose domain is a proper class is itself a proper class.

You use functions and relations between proper classes all the time -- e.g. in ZFC, [itex]\in[/itex] is a binary relation on Set, and { } is a function from Set to Set (it's the function that sends the set S to the set {S}).

Note that to even reason about proper classes, you have already implicitly chosen a formalism that permits us to talk about such things.
 
  • #27


OK, I've been reading some category theory and mathematical logic, and I still don't see an answer to what I'm mostly concerned about. I thought that there would be a definition of "structure" and "structure-preserving" that covers all of the "sets with something defined on them" that we work with in mathematics.

When you study the definitions of "group", "ring" etc. you can't help noticing how similar the definitions are. It seems obvious that they're special cases of something more general. This something is called an "algebraic structure" or an "algebra". But it doesn't seem to be possible to extend this to include all of the "sets with something else defined on them" in mathematics. (I don't see how to do it for metric spaces. I can do it for modules, if I treat the scalar multiplication function as a unary operation for each scalar, but I don't know if "structures" with uncountably many operation symbols make sense).

I understand that "category" is another generalization of these concepts. While "structure" generalizes "group", "category" generalizes...uh not sure what I should call them...let's go with "L-structures and L-homomorphisms". (For example "groups and group homomorphisms". "L" is the set of symbols that distinguish one first-order language from another. I've seen it called a "language", a "lexicon" or a "signature").

It bugs me that "category" doesn't generalize "homomorphism" in a way that explains why the definitions of the standard choice of arrows in so many categories look so similar. For example, I was able to guess the appropriate way to define the arrows in the category of fiber bundles, and it certainly wasn't the definition of "category" that helped me do that. It was my intuition about what "structure-preserving" means. I just feel that there must be a way to turn that intuition into a definition, and it surprises me a lot that there doesn't seem to be a textbook answer to this.
 
  • #28


Fredrik said:
It bugs me that "category" doesn't generalize "homomorphism" in a way that explains why the definitions of the standard choice of arrows in so many categories look so similar. For example, I was able to guess the appropriate way to define the arrows in the category of fiber bundles, and it certainly wasn't the definition of "category" that helped me do that. It was my intuition about what "structure-preserving" means. I just feel that there must be a way to turn that intuition into a definition, and it surprises me a lot that there doesn't seem to be a textbook answer to this.
I don't think this can be done. A lot of time there are several possible choices of arrows between objects. For example, while SET - the category whose objects are sets, and whose arrows are functions - is the standard category with sets as obhects, we can just as well consider the category whose objects are sets and whose arrows are injective functions. Or Banach Spaces with bounded linear maps, or Banach Spaces with linear contractions (Lipschitz continuous with constat K<=1). Or...
Categoy Theory won't tell you what the "homomorphisms" (arrows) should be, you have to choose them yourself. After all, in any category, the arrows are the most important, not the objects.
 
  • #29


Landau said:
I don't think this can be done. A lot of time there are several possible choices of arrows between objects. For example, while SET - the category whose objects are sets, and whose arrows are functions - is the standard category with sets as obhects, we can just as well consider the category whose objects are sets and whose arrows are injective functions.
That's why I'd like to generalize the "structure" concept instead of talking about "categories". Functions don't have to be injective (or have any special properties) to preserve the "structure" of a set, because there are no operations to preserve. Group homomorphisms on the other hand have to satisfy f(gh)=f(g)f(h), because that's precisely what's required to preserve the "multiplication" operation.

Landau said:
Categoy Theory won't tell you what the "homomorphisms" (arrows) should be...
I know, but model theory does. At least when the class of objects is the the class of structures for a given first-order language. The only problem is that the definition of "structure" isn't general enough to include all those things that I think of as "structures", lacking a better term (metric spaces, manifolds, fiber bundles,...).
 
  • #30


I was just arguing that Category Theory won't do what you want (i.e. I was agreeing with your last post).
 
  • #31


It's interesting you choose fiber bundles as an example. If you generalize to all bundles, the construction is purely formal -- it's called the "slice category" and makes sense for any category whatsoever.

Of course, I believe the slice category was originally inspired by the construction of bundles. But for a kind of "good" category, slicing turns out to be a special case of a particularly natural sort of construction.
 
  • #32


Do "slice categories" go by some other name? I bought Goldblatt's "Topoi" book, but I can't find it in the index.
 
  • #33


How about "comma category"?
 
  • #34


Pages 34-36:

If X is a category, the corresponding "arrow category", X has the arrows of X as objects. The arrows of X are pairs of arrows of X. For example, if f:A→B and g:C→D are arrows of X and therefore objects of X, an arrow from f to g is a pair (h:A→C, k:B→D) of arrows in X, such that [itex]k\circ f=g\circ h[/itex].

A comma category is pretty much the same thing, except that all the arrows have the same domain, or the same codomain. If X is a category, the objects of the comma category X↓A are the arrows of X that have codomain A, and an arrow in X↓A from f:B→A to g:C→A is an arrow h:B→C in X such that [itex]g\circ h=f[/itex].

The objects of the comma category X↑A are the arrows of X that have domain A, and an arrow in X↑A from f:A→B to g:A→C is an arrow h:B→C in X such that [itex]h\circ f=g[/itex].
 
Last edited:
  • #35


The downarrow version is the slice category. (I've heard the uparrow one called the coslice category)

If T is the category of spaces, then T/X is precisely the category of bundles over X. The fiber bundles are a full subcategory therein.


I know, but model theory does.
Not really -- it has the same problem. Model theory tells you what the homomorphism are if and only if the particular structure you are interested in is "models of a particular theory" -- and even then you have to know which theory.

An interesting example are these two alledged theories of monoids:
Theory 1: A monoid consists of a set X with a binary operator * such that
  • a*(b*c) = (a*b)*c
  • There exists an element e such that e*a = a*e = a
Theory 2: A monoid consists of a set X with a binary operator * and constant e such that
  • a*(b*c) = (a*b)*c
  • e*a = a*e = a
Both theories define the class of monoids. However, they define different classes of homomorphisms. As a particular example, consider the multiplicative monoid (Z,*) of integers. The function f(x)=0 is a homomorphism of models of the first theory, but not of the second theory.

A similar example is notorious in ring theory -- if a ring has a multiplicative unit, does a homomorphism have to map it to a multiplicative unit? Differing conventions give different answers.



It's common to define structures to be models of a particular choice of first-order theory. If you do so, then obviously model theory tells you what the homomorphisms are.

Of course, the same is true of category theory. If you define structures as "objects in a category", then obviously category theory tells you what the homomorphism are. :wink:




Incidentally, I'm reminded of the situation of "modules over a graded ring". I don't know the history well so I might be telling it wrong, but as I understood it, at one time there were several possibilities for what precisely that should mean. It wasn't until the theory of Abelian categories was applied to module theory that things became obvious.
 

Similar threads

  • Linear and Abstract Algebra
Replies
1
Views
1K
  • Linear and Abstract Algebra
Replies
8
Views
1K
  • Linear and Abstract Algebra
Replies
22
Views
4K
Replies
4
Views
1K
  • Linear and Abstract Algebra
Replies
1
Views
2K
  • Differential Geometry
Replies
20
Views
2K
  • Linear and Abstract Algebra
Replies
4
Views
3K
  • Linear and Abstract Algebra
Replies
1
Views
912
  • Linear and Abstract Algebra
Replies
2
Views
4K
  • Linear and Abstract Algebra
Replies
6
Views
3K
Back
Top