Let me state first off that I am a little uneasy posting to sci.physics.research since I understand so little of physics. But a recent thread turned to Chern classes, which is not only mathematics but a part of mathematics I understand moderately well, so after a couple of email messages I have been persuaded to write in with some explanatory material. The question I can answer is, "what are Chern classes?" (Sorry, I can't really help with the question, "why are they important to physics?") The short answer is: they are cohomology classes associated to vector bundles over a topological space M. I will try to outline a few of their properties, which will be a little tricky since I am not going to assume much familiarity with the terms "cohomology classes" and "vector bundles", either. Let me start by asking for a model of two topological spaces E and B and a two-to-one map f: E-> B between them. Here are some examples: * E is a set with two points, B is a set with one points. No choice for f ! * E and B are both the unit circle in the plane, and f(z) = z^2 * E and B are both cylinders, and f can be described as "rolling up the cylinder E twice as tightly to get the cylinder B". I intend the last example to sound like cheating after the second example. If you recognize a cylinder as being a product S^1 x R of a circle and a line, you can describe the map I intended as f(z,t) = (z^2, t). In other words, we have a diagram of this shape: E = circle | | squaring map \|/ v projection B' = cylinder ------> B = circle and we take advantage of it to get a two-to-one map of a cylinder E' onto B' sort of by using the "squaring" map on the right and combining it with the projection map on the bottom somehow. There is a trick to doing this, which you don't really have to understand, but it works: from the diagram above we construct another space E' and a mapping E' --> B' which, exactly as with the mapping on the right, is two-to-one. This mapping E' --> B' is known as the "pull-back" of the squaring map E --> B (via the projection map on the bottom). So we have this machine for making examples of two-to-one mappings: start with one such -- say the squaring map I've been using -- and then get oodles more by taking arbitrary maps into B and constructing the pull-backs (whatever they are). (OK, I can't resist giving the construction, because it's so easy. Given any map f : E --> B between two spaces, and any other map g : B' --> B into the "base" space B, we "pull back f via g" as follows: let E' be the set of pairs (b', e) in the Cartesian product B' x E for which g(b') = f(e) (both sides are members of B). Let f' : E' --> B' be the projection onto the first component. That's all there is to it.) Now, there are plenty more examples of two-to-one maps. You may be familiar with the covering map S^2 --> RP^2 which amounts to gluing each point of the Southern hemisphere to its antipode in the Northern hemisphere. This map isn't so easily related to the squaring map I already mentioned. Well, actually, it is, but this new example is more complicated -- you can derive the squaring-map example from this new one, but not the other way around. What I mean is the observation that there is a subspace B of RP^2 which is homeomorphic to the circle, and which is covered by the equator in S^2, also a circle. In other words, the squaring map is really just the pull-back, via the inclusion B --> RP^2, of the covering map S^2 --> RP^2. And by the way, you can construct a pull-back of a pull-back; it's another pull-back (exercise: what does this mean?). Therefore, all the examples we could concoct starting with the squaring map may now be created from the covering map S^2 --> RP^2, too. To recap: we had one example of a two-to-one map, we observed we could get many more from this, but there were other examples which didn't arise this way (trust me on this: the covering by S^2 does not arise as a pullback of the squaring map). So we added another starting example, then found that it and its pullbacks accounted for all the examples known so far. Are we done? Well, no: there's another example of a two-to-one map, namely the covering S^3 --> RP^3. It's not a pullback of the previous example, although the previous one IS a pullback of this one! So this one accounts for all the examples listed ... until we think about the covering S^4 --> RP^4, when the cycle repeats. OK, this may seem like overkill, but it turns out you can lump together all of these examples into one: there is a space called S^\infty (the "infinite dimensional sphere"), which includes all the spheres S^1, S^2, ... (each is the equator of the next), and which covers a sphace known as RP^\infty in a two-to-one way. AND, at last, this example is "universal": one can prove that ALL two-to-one maps (in a certain category of topological spaces which is plenty general enough for physicists :-) ) may be obtained as pullbacks of this one covering S^\infty --> RP^\infty, using all possible maps of spaces into RP^\infty. This space we have now constructed, RP^\infty, is then known as the "classifying space for Z/2Z bundles"; the notation B(Z/2Z) is sometimes used ("B" as in "base space", see above). (With some categorical nonsense, it follows that once there is such an object, it's essentially unique. So I can use the word "the" with confidence...) Now, you can sort of argue that RP^\infty is "made to order" for two-to-one maps. I will press that idea in my next construction. First, let me observe that RP^n can be thought of as the set of all lines (through the origin) in (n+1)-dimensional space. After all, every such line meets the sphere exactly twice at antipodal points, so once you glue those pairs together, you have one point of RP^n for every line, and vice versa. This is sort of where all the 2's come from: the group of isometries of the real line is Z/2Z. Or you can collect together all the lines, in the manner of the pullbacks described earlier: consider the set E of all the pairs (r, v) where r is an element of RP^\infinity (i.e. a line) and v is any vector on that line. All the points in E with a given first coordinate r form a line, and indeed if you take a fairly small patch U of RP^\infty, all the pairs (r,v) in E whose first coordinate lies in the patch U will collectively form a subspace of E which is homeomorphic to the product U x R; we say the pair of space E --> RP^\infty is a "line bundle" for this reason. All right, now let's play this game with complex geometry. There's a space called CP^n which consists of all the complex lines (one-dimensional subspaces) in C^(n+1). Each such line meets the unit sphere in this space (which would be S^(2n+1), if you're keeping score); indeed, there's a circle's worth of points in each such linear subspace at a distance of 1 from the origin. So you can keep track of these 1-dimensional subspaces by taking S^(2n+1) and sewing together these various circles together, collapsing each to a point. If you're like me you have some trouble visualizing this, but you needn't bother: all you have to know is that there is a map S^(2n+1) --> CP^n with the property that it's a "circle-to-one" map (the inverse image of each point in CP^n is a circle in S^(2n+1) ) and that the points of CP^n correspond to complex lines. (Indeed, you can go whole hog and think of this map as simply being a restriction of the map C^(n+1) - {0} --> CP^n which simply assigns a nonzero vector in C^(n+1) to the complex line which contains it. In the language of equivalence relations, this is the "natural map" v |--> [v], but no one ever seems to think of the natural map as being natural the first time they see it.) The constructions for the various n's hang together nicely, and with suitable topological niceties, one obtains a pair of space S^\infty --> CP^\infty which turn out to be "universal" for "circle-to-one" maps; you'll see CP^\infty referred to as B(S^1) or B(U(1)) for this reason (U(n) is the unitary group. Oh yes, there are some n's coming...) As in the real case, we may also form the total space E consisting of all pairs (r, v) where r is an element of CP^\infty, that is, a complex line, and v is a vector in this line. Then E --> CP^\infty is a complex line bundle: not only is the inverse image of each point a copy of the complex line, but locally the inverse images hang together (i.e., the inverse image of sufficiently small neighborhoods U will be homeomorphic to products U x C). This space is universal for complex line bundles: in a suitable category of spaces, every complex line bundle may be constructed as a pullback of the bundle map E --> CP^\infty via some map into the base space. You can, of course, play all these games with higher-dimensional subspaces of real or complex vector spaces than these simple one-dimensional subspaces we've covered. Instead of projective spaces you get Grassmannian manifolds, and instead of Z/2Z or S^1 you get orthogonal groups or unitary groups. Instead of line bundles you get vector bundles. (You can even do it all over the quaternions if you're careful about non-commutativity.) The point is that all the bundles people worry about can be assumed to be created once and for all in a universal way: there exists a universal (real/complex) n-bundle E --> B for each n, and all the other bundles people concoct can be described simply as pullbacks of this one universal one, pulled back along various maps into the base space B. It's sort of nice to be able to separate all the bundle stuff into one giant concrete example, together with more or less arbitrary maps into this scheme. If you need to stop for a cup of coffee, now would be a good time. Ok, welcome back. Now I'd like to talk about cohomology classes. Cohomology is a very general term for a lot of very useful tools in mathematics. The big plus is that they take some complicated (e.g. geometrical) information and distill it into something algebraic. One difficulty with the usual presentations, however, is that once someone has computed a cohomology group, or a cohomology class, they don't have a way of describing just what it is they have found. Here is a way around it. From general categorical nonsense, one can show that, whatever the homology group H^n( X, G ) means for a topological space, it must be true that (with n and G fixed) there is a space K for which H^n(X, G) is naturally isomorphic to the set of (homotopy classes of) maps from X into K. Here G is assumed to be an abelian group, and then an abelian group structure can be put on this set of maps, and the isomorphism is an isomorphism of groups. To keep the various K's straight, we might write this as K(G,n) rather than K. It's called an Eilenberg-Maclane space. If you grant that it exists, then some basic information about spheres forces this space to have the property that all its homotopy groups are zero except the n-th, and that one is isomorphic to G. Almost none of these K's is a space you'd recognize unless you've been keeping really disreputable company. So that's what a "cohomology class" is (may be considered to be): it's a (homotopy equivalence class of) map from the given space X into a particular and peculiar space K(G,n). This can be a useful perspective for proving a few common facts about cohomology. For example, suppose we have a map f : X --> Y between two spaces. Given a cohomology class u in H^n(Y,G), we think of u as a map u : Y --> K(G,n); this of course leads to a composite map (u o f) : X --> K(G,n), which is then an element of H^n(X,G). Typically we write f^*(u) instead of u o f but in any event it's clear that f has created for us a function f^* : H^n(Y,G) --> H^n(X,G). It's not so clear but is true that f^* is actually a homomorphism of groups, meaning that H^n( ---, G) is a functor from the category of spaces to the category of groups. (You may notice the change in direction of the morphisms; to a mathematician this means it's a contravariant functor, although I have been told that physicists use the words covariant and contravariant differently so I'll just not call it anything.) A little more yet is true (and not so accessible from this point of view): if G happens to be (the underlying additive group of) a ring, then the collection of groups H^*(X,G) also forms a ring, and f^* is a homomorphism of rings. Since the most natural choices for G (namely Z, the integers, and R, the reals) are also rings, this fact is quite handy. For example, one can give quite succint results this way: Theorem: H^*( CP^\infty, Z ) = Z[x] with dim(x) = 2 This says that the cohomology ring is the ring of polynomials in one variable. Then H^n is the subspace of homogeneous polynomials of dimension n, meaning the set of constant multiples of x^(n/2) (and if n is odd, H^n contains only 0 ). Additively, this is just the group of integers, so we have simply asserted that H^n is {0} if n is odd, and Z if n is even. The real case is a little more peculiar: Theorem: H^*(RP^\infty, Z) = Z[x]/(2x) with dim(x) = 1 (In other words, H^n is Z/2Z for all n except n=0; H^0 is Z.) Recall that CP^\infty = B(U(1)) was to be just the first in a series of spaces. I'll reveal the Secret for the next one and let you guess the pattern for the others. (Your guess will be right :-) ) Theorem: H^*( B(U(2)), Z ) = Z[ c_1, c_2 ] with dim(c_n) = 2n. So now we are able to reveal the Magic Decoder Ring for Chern Classes. Suppose E --> B is a complex line bundle. By universality, there is a map f: B --> CP^\infty such that E is simply the pullback of the universal line bundle S^\infty --> CP^\infty. By the first calculation (Theorem) above, H^*(CP^\infty, Z) = Z[c_1] where dim(c_1) = 2. By functoriality there is a ring homomorphism f^* : Z[c_1] --> H^*(B,Z). By simple ring theory, this homomorphism is completely determined by the image f^*(c_1). By quite a few homotopy-theory details I have glossed over :-) all of this depends only on the line bundle E-->B with which we began. Therefore we are justified in choosing a notation which involves only this initial data. We write c_1(E-->B) (or c_1(E) or something) for this cohomology class f^*(c_1) and call it the first Chern class of the line bundle. If we had begun with a vector bundle with higher dimension, we would need to invoke a different classifying space, whose cohomology would be a polynomial ring in several variables c_i; we write c_i(E) for f^*(c_i), which is an element of H^(2i)(B,Z). If we had begun with a real vector bundle, we would be computing elements w_i(E) = f^*(w_i) where w_i is an element of H^i, and is of additive order at most 2 (unless i=0); these are the Stiefel-Whitney classes. I want to stress that many of the facts one might wish to prove about Chern classes is now rather easy, or at least can be pushed away from bundels altogether and reduced to certain calculations about the classifying spaces. (Product formulas, action of Steenrod algebras, etc.) Also I can't resist drawing the parallels: n-plane bundles over B are determined by maps into B(U(n)) elements of H^n(B,Z) are just maps into K(Z,n) This is particularly cool in the case n=1 because B(U(1)) is a K(Z,2) ! (Also in the real case B(O(1)) = K(Z/2Z, 1).) Thus we find that line bundles are in one-to-one correspondence with elements of H^2(B,Z), and in fact it's the (first) Chern class which sets up this correspondence.