Vous êtes sur la page 1sur 6

Why Abstract Mathematics Matters1 It often seems, to those with only a high school (or even basic university)

education in the subject, that mathematics is largely a solved problem. Sure, the thinking goes, there are those arcane and abstruse little corners of irrelevancy with which academics busy themselves, but for the most part we know it all, right? In fact there is a great deal that we simply do not know, and do not understand; and a lot of it is much simpler, and much closer to home than you might think. In many ways our understanding of the subject only really began very recently. Lets take a brief tour of some of what we dont know. Certainly at the cutting face of research mathematics there is much to learn and much that is not understood. It can be argued that these topics are precisely those obscure irrelevancies, mathematical chimera, that were dismissed earlier; but they provide a point of entry and are, perhaps, not as irrelevant as they may seem at rst glance. The diculty of describing these areas is the layered scaolding of denitions and intermediary results that mathematicians have constructed beneath them as they climb towards their goal. Such constructions are often daunting to the outsider, and serve to obfuscate what lies at their heart. This means that the depth of the underlying elegance may not be clear as we climb, but hopefully we will catch a glimpse as we climb down the other side. The particular edice I would like to climb for the purposes of this discussion is that of Group Theory and Representation Theory. Along the way Ill try and point out some of the various directions where our understanding is less than perfect without necessarily getting into the details of those specic elds. Lets begin with the denition of a group. Well sacrace a little rigour and precision for the sake of brevity: A group is a set of objects with a particular algebraic structure. By an algebraic structure we mean a set of operations that take one or more elements of the set as arguments and return a single element of the set. A natural concrete example is the algebraic structure provided to the set of natural numbers by addition and multiplication. Both of these operations take two numbers and return a single number. The more such operations you have the greater the scope for complex interplay between them. For the particular algebraic structure that denes a group we ask for things to be (almost) as simple as possible. We require a single operation that is associative (the order in which we apply the operation doesnt aect the end result), and takes exactly two arguments. We also require the existence of special elements with regard to how they behave under this operation. First, we require an identity element: an element e such that for any other element x we have f (x, e) = f (e, x) = x where f is our associative operation. Second, we require the existence of inverse elements: for each x we require some element x such that f (x, x) = f (x, x) = e. Any set of objects with an algebraic structure satisfying all these constraints is called a group. Adding more operations and
1

This is a public domain article whose author is unknown.

dierent constraints leads to a profusion of other things to study: rings, domains, elds, modules, algebras, etc. Given just this abstract axiomatic denition, just the raw scaolding, it is hard to see the motivation for studying such things - why do we abstract away from the regular numbers in this way? Surprisingly, groups, rings, elds, and other such structures, crop up in all manner of dierent contexts - once you think to look for them you start to nd them everywhere. It turns out that we require more study of such things, not less. For example, in todays fast paced IT world, programming multi-threaded and multi-process applications is considered hard, largely because it is hard to think and reason about many dierent things all happening at same timein parallel. The study of communicating concurrent processes, which is to say multi-threaded and multi-process systems, can be reduced to a simple structure that is almost (but not quite - it lacks a couple of constraints) a ring. The subtle dierence puts it just outside the reach of the vast array of mathematical tools and theorems we have developed. A better understanding of those structures could quickly make reasoning about and programming massively concurrent multithreaded systems exceptionally simple. For now it remains one of those areas about which we know far too little. Let us get back to groups however. For something as surprisingly prevalent as it is, we know a lot less about groups than you might think. For our purposes, we require only some very general properties of the algebraic structure imposed upon the set. The specic properties of the set itself, along with the algebraic structure imposed, can lead to a wide variety of dierent groups each with their own unique character and properties. Perhaps the most immediate division is between groups with only nitely many elements and innite groups. One might expect that when restricted to just groups with nitely many elements it should be relatively simple to work out what all the possibilities are. Classifying nite groups is, however, a remarkably dicult problem and remains to this day unsolved. That is not to say that a lot of work has not been done. So called Abelian groups, where the operation is commutative (that is f (x, y) = f (y, x) for any x, y in the group) are well understood and can essentially be characterised knowing little more than the size of the set on which the algebraic structure is dened. Similarly, after a vast eort spanning decades, there is a classication of all so called nite simple groups. The best existing theorems classifying general nite groups are the Sylow Theorems which decompose groups into what are called p-groups, which are closely related to the prime numbers. The classication of nite p-groups remains an open problem on which much work is currently being done using methods and techniques from seemingly unrelated elds, including theorems about Lie algebras. It appears that the structure of nite groups is a very complex problem that relies on a much deeper understanding of algebraic structures other than just groups. Let us turn our attention then to Representation Theory. For Representation Theory we will require a little more scaolding. We begin with group actions. A group action is a way of thinking of a group (call it G) as acting on some other set (call it A): We can think of each element of G as a function from A to A, and to make sure it all make 2

sense we will require that for any x, y G and a A it follows that x(y(a)) = f (x, y)(a) where f is, as before, the group operation. That sounds rather obscure and technical, but essentially we are just asking that the if we view elements of G as functions, those functions should still respect the algebraic structure of the group. We also require that the identity element of the group should act as an identity function on A. If A has some structure (be it algebraic or otherwise) of its own we can ask the inverse question: what of the set of all (bijective) functions from A to A that preserve the structure? Surprisingly that set of functions turns out to behave precisely as a group with a natural pre-dened action on A. It is in this sense that groups are often referred to as symmetries. At their heart, symmetries are structure preserving functions; wherever you nd symmetries, you nd a group that describes them perfectly. The next ingredient we require for Representation Theory is group homomorphisms. These are, simply put, maps or functions from one group to another. At rst glance this is no dierent than functions on the underlying setswhich most people should be familiar. What makes it a function between groups is the necessity that the function respect the structure of the groups involved. Specically, we require that a function from a group G to a group H obey (f (x, y)) = g((x), (y)) where f and g are the group operations for G and H respectively. This doesnt seem particularly momentous, but a surprising amount of theory ows from this very simple denition. It turns out that we can learn everything there is to know about the internal structure of a group simply by looking at how it relates to other groups through homomorphisms: the external relationships dene the internal structures - this is a surprisingly deep insight and is one of the founding observations for Category Theory, a language and theory that is having revolutionising impacts on both mathematics and computer science. We have another topic to cover, and that is vector spaces. Ill try to leave out details that wont concern us so as to keep things moving along. Probably the easiest way to think of a vector space is as an n-dimensional spacethat is, a space of points where each point requires n pieces of information to uniquely specify it. We then allow scaling of points, multiplying by a scalar which is just an ordinary number. In essence this is simply stretching the point away from (or toward if the scalar is less than 1) the origin. We also need to be able to add points together, which works in a vector style parallelogram addition. Now, to say that each point can be uniquely specied by n pieces of information is to say that there is a set of n points such that each point can be uniquely expressed as a sum of multiples of those n points. We can ask if there is just one such set of n points that will let us do that. The answer is nothere are lots! So what if we start with one set of n base points that work, and have a function that maps them to another set that also works? This is what is called a linear transformation, and as long as we use linear transformations the structure of the vector space (the ability to scale and add points and express any point as a combination of n base points) is always preserved. If the whole concept of preserving structure is sounding familiar, then you can guess where this is headed. 3

Take a vector space and consider the set of all nonsingular (invertible) linear transformations on the vector space. That set of linear transformations, as noted when discussing group actions, forms a group. It is possible then, given some group G, to wonder about all the group homomorphisms from G into the group of linear transformations of some vector space. Each such homomorphism is called a representation of G, and the study of such representations is called Representation Theory. A couple more facts: the degree of a representation is the dimension of the vector space, and a representation is called irreducible if it cannot be broken down and expressed as a sum of simpler representations. Again, this sounds like a mounthfulan unnecessarily abstract and futile exercise, but given that relationships to other groups can tell us a great deal about the potentially complex internal structure of a group, and that linear transformations of vector spaces are easy to deal with (they are just matrices), it promises a way to better understand the internal structure of complicated groups. In practice Representation Theory provides an extremely rich and enlightening view of groups, the details proving to be far more intricate and beautiful (at least in an abstract mathematical sense) than can really be imagined without an in depth study. For example, other new interesting structures called group-rings and D modules arise in a natural way, and both prove to have their own intriguing and poorly understood complications. A new and barely explored expanse of open problems beckons uswe discover just how little we actually know. There is a great deal more to learn. Mathematicians continue to expand our understanding of these subjects, and there is much more that could be said before we truly arrive at the cutting edge of mathematical research. I would like to stop at this point, in part because many readers are undoubtedly already feeling that we went too far into the abstract airy heights of irrelevancy, and also because weve arrived at a convenient point to take a surprising new look at weve got. So far weve been discussing the completely abstract. Weve climbed up a path of increasingly theoretical constructions built one atop the other to reach a point of potentially perilous irrelevance. Now its time to step back into the concrete world. Our universe can be thought of as a space with some kind of structurethe laws of physics. We can ask what functions or transformations of space and time will preserve certain structures (like, say, Maxwells equations). This leads us to a group of transformations of spacetime that preserve the desired properties. As I said, groups, once you know to look for them, crop up everywhere. Now something intriguing happens: the internal structure of a group corresponds in a natural way with our physical conservation laws, such as the law of conservation of energy, or the law of conservation of momentum. Things get really interesting, however, when we realise that, given a group, each irreducible representation of the group very naturally corresponds to a dierent fundamental particle of physics, force-carrying and non-force-carrying particles neatly delineated from one another according to the degree of the irreducible representation! All of a sudden what had seemed like a little axiomatic game that mathematicians were o playing by themselves is having 4

very real impacts on the world in which we live. All of a sudden weve come crashing back down into relevance. Which brings us to the start of the truly hard questions: why does it work?! Mathematics works. That much is obvious to anybody. Mathematics is essentially the language of much of science, and is seemingly unendingly applicable to the real world. The problem is that it is not at all obvious why this should actually be the case. Philosophers have struggled with this question since Pythagoras, all to no avail. It is entirely all too common for mathematicians to embark on a purely theoretical exploration of the truly abstract and abstruse as a mental game, only to have, decades or centuries later, their work prove stunningly applicable to some very real world problem. The mathematics that laid the foundations for General Relativity began with mathematicians wondering, purely hypothetically, what would happen if they ignored Euclids fth postulate; much of the work of G.H. Hardy, who famously claimed no discovery of his would ever make a practical dierence in the world, has found considerable application in cryptography. To explain its eectiveness it seems we need rst to explain what mathematics actually is. If mathematics is simply the abstraction of our intuitions about the physical world, then why is it so universal? Surely each mathematician would then develop a personal mathematics from their own intuitions and claims of mathematical objectivity would become muddied. Equally one can ask why it is that a rejection of intuition on purely logical grounds, such as the rejection of Euclids fth postulate, should lead to mathematics that is surprisingly more applicable rather than less. Can it then be argued that mathematics is simply logic? This would explain, to some degree, its applicability: mathematics would simply be the expression of fundamental eternal universal truths. Unfortunately, despite truly heroic eorts by some of histories greatest mathematicians and logicians (Frege in Foundations of Arithmetic Russell and Whitehead in Principia Mathematica) the reduction of mathematics to pure logic failed. While Russell and Whitehead managed to construct a remarkable edice out of pure logic, introducing along the way many new concepts such as Type theory which is now of great importance in computer science, they fell short of building mathematics as we know it: their system was too weak and required extra axioms (the axiom of innity and the axiom of reducibility) which cannot reasonably be called purely logical principles. Can we not, then, dene our base set of rules and proceed by logic from there? This was the approach to mathematics favoured by Hilbert and his formalist school. While this approach did much to shore up the shaky foundations of pure mathematics before running afoul of Godels Incompleteness Theorems, it simply sidesteps our initial question. Mathematics, in this line of thought, becomes reduced to a game played with symbols on paper and an arbitrary set of rules (we ask only for consistency). Why should such an arbitrary game apply to reality? That is not at all clear. The debate as to what mathematics actually is continues, mostly in philosophic circles. No school of thought has yet given a satisfactory answer. Personally I lean towards a 5

more recent view that mathematics is about structure and can be grounded best in a structure oriented language such as that of Category Theory. However, Category Theory does not yet provide a suciently rigorous foundation to make that claim truly defensible. Only time will tell. Thus we began wondering what there is in mathematics that we dont know, and conclude realising that we dont even know what mathematics is.

Vous aimerez peut-être aussi