Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Topology and Geometry for Physicists
Topology and Geometry for Physicists
Topology and Geometry for Physicists
Ebook505 pages7 hours

Topology and Geometry for Physicists

Rating: 3.5 out of 5 stars

3.5/5

()

Read preview

About this ebook

Differential geometry and topology are essential tools for many theoretical physicists, particularly in the study of condensed matter physics, gravity, and particle physics. Written by physicists for physics students, this text introduces geometrical and topological methods in theoretical physics and applied mathematics. It assumes no detailed background in topology or geometry, and it emphasizes physical motivations, enabling students to apply the techniques to their physics formulas and research.
"Thoroughly recommended" by The Physics Bulletin, this volume's physics applications range from condensed matter physics and statistical mechanics to elementary particle theory. Its main mathematical topics include differential forms, homotopy, homology, cohomology, fiber bundles, connection and covariant derivatives, and Morse theory.
LanguageEnglish
Release dateAug 16, 2013
ISBN9780486318363
Topology and Geometry for Physicists

Read more from Charles Nash

Related to Topology and Geometry for Physicists

Titles in the series (100)

View More

Related ebooks

Mathematics For You

View More

Related articles

Reviews for Topology and Geometry for Physicists

Rating: 3.5 out of 5 stars
3.5/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Topology and Geometry for Physicists - Charles Nash

    CHAPTER 1

    Basic Notions of Topology and the Value of Topological Reasoning

    1.1 INTRODUCTION

    Topology can be thought of as a kind of generalization of Euclidean geometry, and also as a natural framework for the study of continuity. Euclidean geometry is generalized by regarding triangles, circles, and squares as being the same basic object. Continuity enters because in saying this one has in mind a continuous deformation of a triangle into a square or a circle, or indeed any arbitrary shape. A disc with a hole in the centre is topologically different from a circle or a square because one cannot create or destroy holes by continuous deformations. Thus using topological methods one does not expect to be able to identify a geometrical figure as being a triangle or a square. However, one does expect to be able to detect the presence of gross features such as holes or the fact that the figure is made up of two disjoint pieces etc. This leads to the important point that topology produces theorems that are usually qualitative in nature—they may assert, for example, the existence or non-existence of an object. They will not in general, provide the means for its construction.

    Let us begin by looking at some examples where topology plays a role.

    Example 1. Cauchy’s residue theorem

    Consider the contour integral for a meromorphic function f(z) along the path Γ1 which starts at a and finishes at b (c.f. Fig. 1.1). Let us write

    Now deform the path Γ1 continuously into the path Γ2 shown in Fig. 1.1. Provided we cross no poles of f(z) in deforming Γ1 into Γ2, then Cauchy’s theorem for meromorphic functions allows us straightaway to deduce that

    Figure 1.1

    This is just the statement that

    where C is the closed contour made up by joining Γ1 to Γ2 and reversing the arrow on Γ1 so as to give an anticlockwise direction to the contour C. The intuitive content of this result is that to integrate f(z) from a to b in the complex plane is independent of the path joining a to b (under the conditions stated). Even if we relax these conditions and allow that the deformation of Γ1 into Γ² may entail the crossing of some poles of f(z), then we still have complete knowledge of the relationship of the two integrals. It is simply

    where the sum is over the residues, if any, of the poles inside C. This simple example uncovers some topological properties underlying the familiar Cauchy theorem.

    Example 2. The fundamental theorem of algebra

    In this example we shall use the Cauchy theorem to give a proof of the fundamental theorem of algebra. First of all a simple consequence of Cauchy’s theorem is that for a meromorphic function f(z) we have

    where n0 and np are the number of zeroes and the number of poles respectively of f(z) lying inside C Let f(z) be the polynomial Q(z) of degree q where

    Then since Q(z) has no poles we have

    It is clear that n0 is a continuous function of q of the q + 1 coefficients a0, . . ., aq. Let us select the coefficients a0, . . ., aq − 1 and write

    However, n0 also only takes integer values. Now since it is impossible to continuously jump from one integer value to another it follows that n0 must be invariant under continuous change of the ai’s. This state of affairs permits the following argument: for large |z|, say |z| > R, |Q(z)|, grows as |aq||z|q which is a large number. Then if C is a circular contour of radius R′ > R, C will contain all the zeroes of Q(z). Next we continuously deform ai, by choosing

    Evidently,

    thus when ε = 0 the integral (1.7) becomes trivial to evaluate and we obtain the desired result

    (We cannot, of course, require aq to tend to zero since this would change the degree of Q(z)).

    Example 3. Fixed points and their applications

    Let f be a function from a set X into itself. The question, does f possess any fixed points?, is one of great interest, and is, in general, tackled using topology. For example if we refer to example 2 above, and choose for f the function

    Then the existence of fixed points, f(z) = z, of f amounts to the fundamental theorem of algebra.

    Let us have a look at the proof of fixed point theorems for functions of 1 and 2 real variables. For functions of 1 variable from an interval (a, b) into itself the proof is simply the assertion that the graph of f(x) must cross the line y = x somewhere, c.f. Fig. 1.2.

    Figure 1.2

    For functions of 2 variables let us choose f to be a function from the disc B² into itself. B² is the disc defined by x² + y² a². To prove that f(x) has a fixed point we start by considering the effect of f on points on the boundary of B². Let x be on the boundary of B² and let it be mapped to f(x), then construct the vector

    shown in Fig. 1.3.

    Figure 1.3

    We can of course use the Definition (1.12) for all points x in B². This gives us the vector field υ(x) on B². Returning to the point x shown in Fig. 1.3, we rotate the point x through 2 π about the centre of B². Notice that in doing so υ(x) also rotates through 2 π once. We then describe this situation by saying that υ(x) has index 1. Clearly the index of a vector field, when rotated through 2 π must be an integer. Thus if we continuously deform the boundary of B² into the dotted curve shown in Fig. 1.4, then the index of υ(x) for this path must still be one—the index is a topological invariant.

    We can now prove our result by reductio ad absurdum. First of all note that if υ(x) has a zero then, since the zero vector has an undefined direction, the index definition breaks down and cannot be defined in the way we gave. Now consider the point p inside B² and assume that u(p) ≠ 0 and let

    Then consider a small circle c of radius ε < L about p c.f. Fig. 1.5.

    Figure 1.4

    By continuity the image of C under f is the dotted circle C′ about f(p) of radius |f′(p)|. We can choose ε small enough so that |ε + f′(P)ε)| < L so that C and C′ do not intersect. Now consider the index of υ(x) computed by going round C once. The vector field is graphed in Fig. 1.6 purely assuming ε ~ 0 and continuity of f.

    Clearly, since υ(x) points from left to right everywhere the index cannot be 1. Thus this assumption of constant unit index must be rejected. This means that υ(x) must have a zero somewhere i.e. that f(x) has a fixed point. This may be extended to the case of the solid sphere or ball Bn, n = 2, 3, . . . etc. This is the celebrated Brouwer fixed point theorem.

    The utility of fixed point theorems is enormous. For example if A is a differential operator, say A = ∆ the Laplacian then the existence of solutions to

    is equivalent to the existence of fixed points of the operator A + I Further important examples are provided by systems of linear differential equations

    Figure 1.5

    Figure 1.6

    or

    Zeroes of the vector field v, commonly called singularities of v, have a special significance. If t is the time, and (1.15) is a set of equations for a dynamical problem, then the singularities of v are the equilibrium configurations of the dynamical system. In general, if we have a function f(x1, . . ., xn) then the singularities of the vector field v where

    are of considerable interest.

    The singularities of v are just the extrema or critical points of f. Their study, using topological methods, provides the foundation for Morse theory, a topic which we shall come to in a later chapter.

    1.2 BASIC TOPOLOGICAL NOTATIONS

    Having taken a topological view of some mathematical situations, it is now necessary to equip ourselves with some mathematical techniques to progress further. These techniques will be supplied to the reader in the chapters that follow. Some of their names, and perhaps only their names, are already known to the reader, and exposure to them now should breed familiarity for what follows. The names are, homotopy, homology, cohomology and fibre bundles to mention the main ones. The mathematical setting, for the development of these techniques, requires some preliminary definition and illustrative examples. This is where we now begin.

    Topological space

    The mathematical structure within which we work is called a topological space. There is no reason to restrict ourselves unduly, so we expect this to be a fairly general definition.

    Let X be any set and Y = {Xα} denote a collection, finite or infinite of subsets of X. Then X and Y form a topological space provided the Xα and Y satisfy:

      i. ∈ Y, X Y

     ii. Any finite or infinite subcollection {Zα} of the Xα has the property that ∪Zα Y

    iii. Any finite subcollection {Zα1, . . . ., Zαn} of the Xα has the property that ∩Zαi Y    (1.17)

    The set X is then called a topological space and the Xα are called open sets. The choice of Y satisfying (1.17) is said to give a topology to X.

    Examples

    (a)   If X is a set and Y is the collection of all subsets of X, i.e. the set 2X, then (1.17) is clearly satisfied. This is called the discrete topology of X.

    (b)   If X is any set, then the simple choice satisfies (1.17). This is called the indiscrete or trivial topology.

    (c)   Let X = R, the real numbers, and let Xα be those subsets of R which satisfy the following condition. For each x Xα there is an open interval (a, b) containing x, i.e. a < x < b, and also (a, b) is contained in Xα, i.e. (a, b) ∈ Xα. This gives a topology to R called the usual topology.

    Having given the Definition (1.17) and the examples (a), (b) and (c), we now need to discuss them. Let us begin with the definition of topological space, what is its underlying purpose or motivation? For example the topologically ignorant inquiring mind asks: why do open sets remain open under infinite union but only under finite intersection? Also, what does this abstract Definition (1.17) have to do with the intuitive discussion of continuous deformation which preceded it? In answer to all this it should be gently emphasised that the, admittedly abstract, Definition (1.17) is the end product of many mathematical experiments with various definitions and concepts. These experiments finally gave forth the present definition of a topological space. The purpose of this work was to build a mathematical structure within which continuity, realised by continuous functions, would naturally reside. In other words, the definition of topological space was to be used to give a definition (as general as could be managed, as is always the case in mathematics, especially in this century), of continuous functions. We need therefore, if we are to understand what is going on, to look at this definition.

    Definition

    A function f mapping from the topological space X to the topological space Y is continuous if the inverse image of an open set in Y is an open set in X.    (1.18)

    Let us examine this definition using example (c) above of the usual topology on the real numbers R. We take X and Y both to be R and so we wish to study real valued functions f(x) of a real variable x. Now note that according to example (c) all open intervals (a, b) are automatically open sets. Providing ourselves with a graph of f(x) we can proceed to apply Definition (1.18).

    Figure 1.7

    Notice that we have given the function f(x) what would normally be called a discontinuity at x = 0. To test our Definition (1.18) we take the open set (5,17) and compute its inverse image f−1{(5, 17)}. It is immediate to see that

    But (−16, −4) is an open set thus we satisfy (1.18). It is now clear that if we replace (5,17) by the interval (a, b) and if 1 < a < b, or, a < b < then f−1{(a, b)} will be an open interval which is an open set. This is sufficient to establish the continuity of f(x) according to the usual ε, δ method for values of x ≠ 0. Clearly at x = 0 something different must happen. To see that this is indeed the case we compute the inverse of the open set (1 − ε, 1 + ε), ε > 0. Evidently f − 1{(1 − ε, 1 + ε)} is the set of X such that − ε < x 1 or

    But (− ε, 1 ] is not an open set. This is because the point 1 cannot be contained in an interval (a, b) which is itself contained in (− ε, 1 ]. At least for this example then, continuity defined by (1.18) coincides with our usual notion of continuity.

    Another point worth examining is the reason why the Definition (1.18) demands that the inverse of an open set remains open under f, rather than the reverse. That is to say would it not be simpler and more natural to define f to be continuous if f maps open sets in X into open sets in Y? We can dispose of this point by a counter example. We display a function f which is clearly discontinuous by reference to our usual intuition and also by reference to Definition (1.18). The function f will, however, have the property that it maps open sets into open sets.

    Figure 1.8

    We display the actual function f(x) in Fig. 1.8, and we continue to use the usual topology for the real numbers. As can be seen from Fig. 1.8, the topological space X is R, and the union of two disjoint pieces give Y. To show that f is discontinuous is an easy exercise. Nevertheless it is easy to convince oneself by taking examples that f always maps open sets into open sets. This explains why the definition (1.18) needs to be stated in terms of inverses of open sets.

    Having attempted to show how the abstract looking objects called open sets can be used to define continuity, we go on to answer another question that we raised earlier. It is the question: why not allow open sets to be open under infinite intersection? The answer is that to allow this in the Definition (1.17) of a topological space would render most topological spaces useless. For example, take the usual topology for R obtained as above, now supplement it by the requirement that open sets remain open under infinite intersection. The result of this is that any real number X is now also an open set, and so, because of the union axiom, is any subset of R. This is because if x R and we consider the open sets Xα given by

    then the intersection of all the xα is just the point x, i.e.

    So the usual topology would, under this stringent requirement concerning intersection, be reduced to the discrete topology on R. This would be of no use for studying continuity. To see this note that under the discrete topology since every subset of R is open, all maps from R to R are continuous. Clearly we do not want a definition of continuity as broad as this. Now we can begin to see why we only require open sets to remain open under finite intersections. We hope that this examination of some of the motivation behind the Definition (1.17,18) has helped the reader.

    Returning now to examples (a) and (b) above consider the discrete and the indiscrete or trivial topology for X. It is clear that the discrete topology is the largest topology that we can give to X, largest that is in terms of number of open sets. Also the trivial topology is the smallest topology that we can give to X. These two extremes motivate the idea of comparing two topologies on X; this can be done if the open sets defining one of the topologies on X are completely contained in the open sets defining the other topology on X. More precisely let T1 = {Xα} and be the open sets defining two topologies T1 and T2 on X. Then if T1 ⊃ T2 the topology T1 is said to be larger than the topology T2, or, alternatively, T2 is said to be smaller than T1. In other words, an open set for T2 is also an open set for T1 but not the otherway round. Another terminology used describes the above situation by saying that T2 is coarser than T1, or equivalently, T1 is finer than T2. The origin of this terminology is that we can think of the open sets of T1 and T2 as being like a kind of grid on X. This being so, the more open sets we provide, the finer the grid on X. There are more open sets in T1 than T2 hence T1 is finer than T2.

    The situation T1 ⊃ T2 is also referred to by saying T1 is stronger than T2 or T2 is weaker than T1. Thus we have stronger, larger and finer and their opposites: weaker, smaller and coarser. Finally T1 and T2 may be such that neither T1 ⊃ T2 nor T2 ⊃ T1, and then T1 and T2 are not comparable.

    There now follows a series of short sections where we introduce, and illustrate by examples, the definitions of some standard and regularly occurring properties of topological spaces.

    Neighbourhoods and closed sets

    Neighbourhoods

    Given a topology T on X, then N is a neighbourhood of a point x X if N is a subset of X and N contains some open set Xα to which x belongs. Notice that we do not require N itself to be an open set. However, all open sets Xα which contain x are neighbourhoods of x since they are contained in themselves. Thus neighbourhoods are a little bit more general than open sets.

    Examples

    (i) R: the real line. Take the usual topology on R then the interval [−1, 15] is a neighbourhood of the point 0. However, [−1, 15] is not an open set because the points −1 and 15 cannot be contained in an open interval (a, b) which also lies within [−1, 15]. Clearly [—1, 15] is a neighbourhood of the points 1, 2, 5, 14 etc. In fact [−1, 15] is a neighbourhood of all x that belong to the open interval (−1, 15).

    (ii) R²: two dimensional Euclidean space. The usual topology for R² is constructed as follows: consider all rectangles in R² of the form (a, b) × (c, d) where a, b, c and d are all rationals. Then the open sets of the usual topology are given by all such rectangles, and all possible unions of such rectangles. Select now a point (x1, x2) in R², say (x1, x2) = (1, 2). Then the three sets: (1, 3) × (−l, 4), [0, 3) × [−l, 4) and the disc (x − 1)² + (y − 2)² ε², ε > 0 are all neighbourhoods of the point (1, 2). The first is an open neighbourhood while the latter two are not.

    Closed sets

    Let T be a topology on X, then any subset U of X is closed if the complement of U in X (which we write as X U), is an open set. Since the complement of the complement of U is U itslef then we could also infer from this that a set is open when its complement is closed. Notice also that this definition applied to the sets X and Σ shows that they are both open and closed regardless of the details of the topology T.

    Figure 1.9

    Example 1. The interval [a, b]. Take the pair of open intervals (−∞, a) and (b, ∞), their union, (−∞, a) ∪ (b. ∞) i, is open under the usual topology on R. The complement of (−∞, a) ∪ (b. ∞) is the closed interval [a, b], Thus we see that all closed intervals are closed.

    Example 2. The plane R² with the usual topology has an obvious closed set: any rectangle of the form [a, b] × [c, d]. In fact by intersecting enough rectangles so as to form polygons with n vertices, and then taking the limit n → ∞ to form a circle x² + y² = a², we can see that the disc B² defined by x² + y² a² is also closed.

    Closure of a set

    Another closely related notion to that of a closed set is that of the closure of a set. Consider a set U, in general there will be many closed sets which contain U. Denote the family of closed sets with this property by {Fα}. Then the intersection of all the Fα written ∩Fα, is called the closure of U and is denoted by . In intuitive terms, the intersection of all the Fα is the smallest closed set which contains U.

    Example. For the real line R we can see immediately that if U = (a, b), then = [a, b]. Notice that the operation of taking the closure of U rendered the open set (a, b) closed. This will always be so. Further, there is nothing to be gained from taking the closure twice, or . If then a set is already closed than it is its own closure, the converse is also true. All these statements became very reasonable and intuitively clear if one works through some simple examples in Rn such as the ones we have used above.

    Boundary and interior

    The interior of a set U is the union of all open subsets Oα of U, and is written U⁰; evidently U⁰ = ∪α Oα. In intuitive terms U⁰, the interior of U, is the largest open subset of U.

    Example 1. Take U to be the closed set B² in R². Then U⁰ is the open set x² + y² < a². On the other hand if we take U to be the open set x² + y² < a², then U⁰ = U. In fact as should be expected by now, U⁰ = U if and only if U is open.

    The boundary of a set U which we shall write b(U) is the complement of the interior of u in the closure of U: b(U) = − U⁰.

    Example 2. If we take the real line R, and U is the set [a, b) then we see that U⁰ = (a, b) and = [a, b] so that

    Note that the sets (a, b), [a, b], [a, b) and (a, b] all have the same boundary namely the pair of points {a, b}. One can also see from this sort of example that open sets are always disjoint from their boundaries, and that closed sets always contain their boundaries:

    Example 3. The boundary of B² in the usual topology on R² is the circle x² + y² = a².

    Example 4. With the usual topology on R² and the choice: U is the set all points with rational coordinates (p/q, p′/q′), we find that b(U) is the whole set R². This latter example is an instance of an interesting situation. Rather than the boundary b(U) being disjoint from U, the boundary b(U) contains U. It is also worth noticing that the interior of U is the empty set ∅. This is because none of the subsets of U are open. If we refer now to the definition of b(U)

    Then we have

    But since

    b(U) = R²

    then

    In other words the closure of U is also the whole of R². When this occurs the set U is said to be dense in R²: more formally, a set U is dense in X if the closure or U is X.

    Example 5. Take the usual topology on R, then it is clear from the above that the set of all rationals Q is dense in R. So also though, is the set of all irrationals.

    Compactness

    Compactness is a very important notion in topology. As with the definition of a topological space its definition needs to be illuminated by some additional discussion. We begin with the definition and follow with the discussion.

    A preliminary definition that we need is that of a cover of a set U. Given a family of sets {Fα} = F say, then F is a cover of U if ∪Fα contains U. If all the Fα happen to be open sets then the cover is called an open cover. Now consider the set U and all its possible open coverings. The set U is compact if for every open covering {Fα} with ∪Fα ⊃ U there always exists a finite subcovering {F1, . . ., Fn} of U with F¹ ∪ F2 . . . ∪ Fn ⊃ U.

    A

    Enjoying the preview?
    Page 1 of 1