Você está na página 1de 7

New Scientist Archive | Selected Article

New Scientist

The missing link...


New Scientist vol 176 issue 2368 - 09 November 2002, page 30

1, 2, 4, 8... What comes next? Nothing. If we're talking numbers, the


obvious next term is 16. But if we're talking a particular kind of
algebra, there is no next term. And it turns out that this is highly
significant. The ultimate number - the humble 8 - lies at the heart of
a mathematical system known as the octonions, and this system
appears to be the key that will allow physicists to fit quantum theory
and gravity together. Strange as it may seem, the number 8 may
provide us with a "theory of everything".

The tale of the octonions begins in the mid-16th century. Until that
time, mathematicians had thought that numbers were God-given, a
done deal. No one could contemplate inventing a new number. But
around 1550 the Italian algebraists Girolamo Cardano and Raphael
Bombelli did just that, by writing down the square root of -1. It took
about 400 years to sort out what the thing meant, but only 300 to
convince mathematicians that it was too useful to be ignored.

By the 1800s, Cardano and Bombelli's concoction had crystallised


into a new kind of number, i, whose square is -1. The square of a
"real number" - the usual kind that we all know - is always positive.
So whatever i may be, it's not a real number, and mathematicians
call it an "imaginary" number to make this clear. A combination of
real and imaginary numbers, like 4 + 5i, is said to be "complex".

We live in a curious Universe in which, as physicist Eugene Wigner


memorably announced, mathematics is "unreasonably effective".
Complex numbers may seem weird, but they turn out to be a
marvellous tool for understanding physics. Problems of heat, light,
sound, vibration, elasticity, gravity, magnetism, electricity and fluid
flow all succumbed to this complex weaponry - but only for physics
in two dimensions.

Our own Universe, however, has three dimensions of space - if not


more. So, since the two-dimensional system of complex numbers was
so effective for two-dimensional physics, might there be an
analogous three-dimensional number system that could be used for
physics in the real world?

The answer is a resounding no. The Irish mathematician William


Rowan Hamilton spent years trying to find a three-dimensional
number system - but with no success. Then, on 16 October 1843 he
had a flash of insight: don't look in three dimensions, look in four.
And it worked. Hamilton named his new numbers "quaternions".

1
New Scientist Archive | Selected Article

Two months later, having heard about quaternions from Hamilton,


John Graves - a British mathematician and an old college friend of
Hamilton's - announced he had found an eight-dimensional number
system. He called it the "octaves". But before Graves could publish,
the British lawyer-mathematican Arthur Cayley made the same
discovery, and published it as an addendum to an otherwise awful
paper on elliptic functions. He called the system "octonions".

The discovery of the octonions was ever after credited to the wrong
person (they are often known as Cayley numbers, even today), but it
didn't really matter because nobody took any notice of them anyway.
The octonions appeared to be nothing more than Victorian
mathematical whimsy.

Graves was not to be put off though, and spent a long while
convinced that his method of going from 4 to 8 could be repeated,
leading to algebras with dimensions of 16, 32, 64 and so on for any
power of 2. He called his 16-dimensional algebra the sedenions, but
he couldn't find a way to make it - or any of the others - work, and
began to doubt whether it could exist.

His doubt was well-founded. We now know that those four algebras,
of dimensions 1, 2, 4 and 8, are the only ones that behave remotely
like ordinary real numbers. The reason is that, with increasing
numbers of dimensions, these systems obey fewer and fewer
algebraic laws - the amount of algebraic structure keeps decreasing.
Put rather too simply, by the time we reach Graves's sedenions,
there's pretty much no algebraic structure left.

Real, complex, quaternion, octonion; 1, 2, 4, and 8 dimensions: even


by mathematical standards this is an odd set of tools. These four
number systems have several features in common, the most striking
being that they are "division algebras". There are many number
systems in which notions of addition, subtraction and multiplication
hold good: when these notions are applied to the integers (... -2, -1,
0, 1, 2, 3, ...), for example, they transform two integers into another
integer. But the same can't be said for division: divide some integers
by others, for example, and the result is often not an integer. But in
these four number systems, you can always divide and yet remain
within the same system.

And that's not the only mathematical operation that sets them apart.
Numbers in these systems are the only ones to have a "norm",
effectively the number's distance from the origin (see Graphic). With
the complex numbers, the norm of x + iy is x 2 + y2. Because of the
existence of a norm, and their divisibility, these number systems are
known as "normed division algebras".

2
New Scientist Archive | Selected Article

This is all very pretty, to mathematicians at least. But surely the


only really important cases are the real and complex numbers. Well,
not quite: the quaternions have shown up in some useful if esoteric
researches - fields such as abstract algebra and topology.

But it's certainly true that the octonions remained in the shadows
for a long time. In 1925 Wigner, working with the mathematician
John von Neumann, tried to make the octonions the basis of
quantum mechanics. But he failed, and the octonions slipped back
into obscurity. Until now, that is.

Rather surprisingly, the octonions have revealed themselves as the


most important system of all. That's because they are crucial to
string theory, the best candidate for a physical theory of everything.
After 150 years, physics is finally telling us the purpose of the
octonions: they are essential to space and time.

String theory is an attempt to marry the large-scale geometry of


Einstein's general relativity to the small-scale uncertainties inherent
in quantum theory. Both these theories are brilliantly successful in
their own domain. But they can't be fitted together: put into the
same framework, they effectively contradict each other. So the
search has been on for a unified theory that modifies them well
enough to fit them together consistently but doesn't destroy their
existing successes.

The current front runner in this search is known as string theory.


Very roughly, the traditional idea that a fundamental particle is a
featureless point is rejected, and particles are modelled instead as
tiny loops of energy - the aforementioned strings. The loops can
vibrate in ways that give them integer quantum numbers, such as
spin, charge and charm.

But all this only works if the loop is a many-dimensional surface that
protrudes beyond the familiar four-dimensional space-time, and one
of the burning questions is just how many dimensions there are. At
the moment, finding the answer seems to depend on finding the
number of dimensions where the theories work most elegantly. And
though physicists have not pinned it down precisely, they have
noticed that something rather pleasing occurs when they work with
3, 4, 6 and 10 dimensions. Interestingly, each of these numbers is 2
greater than that of a normed division algebra: subtract 2 from 3, 4,
6 and 10, and you get 1, 2, 4 and 8. And that's no coincidence:
these algebras are a vital part of the theory.

Consider, for example, the relationship between two mathematical


objects: vectors and spinors. A vector is essentially a way to

3
New Scientist Archive | Selected Article

describe the size and orientation of something. Velocity, for


example, is a vector because it describes a body's speed and the
direction in which it is moving. The spinor is a more esoteric
mathematical gadget invented by Paul Dirac to describe electron
spin. It turns out that the relationship between vectors and spinors
holds precisely (and only) in space-times of 3, 4, 6, and 10
dimensions. This happens because, in 3, 4, 6, and 10-dimensional
string theory, every spinor can be represented using two numbers
in the associated normed division algebra. This doesn't happen for
any other number of dimensions, and it has lots of nice
consequences for physics.

So we have four candidate string theories here: real, complex,


quaternionic, and octonionic. The one that is thought to have the
best chance of corresponding to reality is the 10-dimensional one,
because it neatly avoids a mass of mathematical obstacles while
allowing the physics to work properly. And, in this system, the
relationships between the properties of matter are specified by the
octonions: if this particular theory really does correspond to reality,
then our Universe is built from pairs of octonions.

If 10 dimensions turns out to be not quite enough, however, it seems


that the octonions will still be found to play a vital role in the theory
of everything. The other very fashionable candidate string theory,
"M-theory", involves 11-dimensional space-time. Although that
means the vector-spinor relationship won't hold, something almost
as good does. In M-theory, the extra dimensions don't need to be
curled up tightly, so the restriction to six extra dimensions can be
relaxed to allow a seventh, but again it doesn't work without the
octonions.

In order to reduce the perceptible part of space-time from 11


dimensions to the familiar four (three space and one time), we have
to hide seven of them. We do that by rolling them up so tightly that
they can't be detected. And how do you do that? You make use of the
octonions' symmetry.

The idea of symmetry - a property that allows you to move something


in a certain way and leave it looking the same - has turned out to be
central to physics, especially the quantum world. All our theories of
fundamental particles, and their strange properties such as spin,
charge and charm, which come in whole-number chunks, boil down
to symmetries. And the use of octonionic symmetry in M-theory even
gives a purpose to a mathematical peculiarity, discovered around
the same time as the octonions, whose existence has always
mystified mathematicians (see "The eightfold way"). So the efficacy
of the octonions here is doubly pleasing.

4
New Scientist Archive | Selected Article

While the octonions started out as mathematical curiosities, and


were almost entirely ignored for 150 years, their time has come.
They are no longer quaint Victoriana, but a hefty clue to a possible
theory of everything. Daunting though their mathematics is,
physicists are beginning to take up this new set of tools and work
with it. A paper published this year by John Baez of the University of
Califonia, Riverside, has prompted much Web-based discussion
between string theorists. It all boils down to one extraordinary
realisation: the humble 8 is no longer just a number. It's our key to
the Universe.

The eightfold way

The best candidate we have for a theory of everything is string


theory. This idea suggests that the fundamental particles are loops
of energy that exist in many more dimensions than the four we
experience. Multidimensional loops can in principle take on lots of
shapes. The big task facing physicists is to find the right one.

In string theory, as in old-fashioned quantum theory, a key principle


to pin down correct theories is symmetry. Physicists often settle on a
particular shape as the correct description of something because it
has the right kinds of symmetry: the Universe seems to like
symmetric characteristics. And in string theory, it turns out, the
symmetry of the octonions is crucial.

A symmetry of something is a way to transform it so that when


you've finished it looks the same as it did at the start. If you take a
featureless circular disc, for example, and rotate it through any
angle, it looks just the same. This is an example of a continuous
symmetry. But with a square, only 90° rotations will do that.

And it's not only objects such as circles and squares that can have
symmetry operations applied to them. Hard as it may be to envisage,
algebras have symmetries too. The collection of all symmetries of a
given shape or algebra or whatever, is called its symmetry group. In
the 19th century, the Norwegian mathematician Sophus Lie
captured such symmetries using an algebraic structure that is
known nowadays as a Lie group. An example is the set of rotations of
an object in three-dimensional space. One symmetry in that set
would be a rotation that turns this magazine through 180° until it's
upside down.

Lie groups - the fundamental kinds of symmetry - can be divided into


four main families. For instance, one family consists of the rotation
group in the plane, the rotation group in space, the rotation group
in four-dimensional space, and so on. Each dimension of space

5
New Scientist Archive | Selected Article

corresponds to one member of the family. Another family describes


all the possible ways to distort space of n dimensions while keeping
straight lines straight. Again, there is one such group for each
dimension n. These are "linear mappings" and they do things like (in
the case n = 2) stretch the plane in a north-south direction while
leaving east-west unchanged; or tilt the north-south axis by 45°
while leaving the east-west one unchanged. It's as if someone has
leaned against the vertical axis and pushed it over. A similar thing
happens with larger n.

But there are five curious symmetry groups that don't fit into any
family. The very existence of these "exceptional" Lie groups, which
rejoice in the names G2, F4, E6, E7 and E8, is a puzzle and a pain to
mathematicians, who like everything to fit into some pattern or
other. One exasperated mathematician even declared them a "brutal
act of Providence".

For decades, no one could find any use for the exceptional groups,
or any reason for their existence, and it was tempting simply to
ignore them. However, it has now been realised that all five of them
can be explained in terms of the octonions. In effect, they form a
small family of their own, one with only five members. And it looks as
though the octonions actually hold together what may prove to be
the theory of everything.

For the 11 dimensions of M-theory to be reduced to the four that we


experience, physicists need to carry out some particular
mathematical transformations on space-time. The only way to do that
is with the exceptional Lie group G2. And what else is this heroic
group that rescues a vital theory of physics? It's the symmetry group
of the octonion algebra.

The octonions have eight units: an ordinary number and seven


others called e1, e2 and so on up to e7. The square of any of these is
-1. The multiplication rule for the units is determined by the "Fano
plane", shown on the left. Suppose you want to multiply e3 by e7,
say. Look at the diagram for the corresponding points, find the line
that joins them, and you will see that there is a third point on the
line, e1. Following the arrows, you go from e3 to e7 to e1, so e3 × e7
= e1. If the ordering is the other way round, throw in a minus sign:
e7 × e3 = -e1. What's more, all the lines are considered to loop back
to the start, so e1 × e3 = e7, and e3 × e1 = -e7. Do this for all the
possible pairs of units and you know how to multiply octonions.

Ian Stewart
Ian Stewart is a professor of mathematics based at the University of Warwick

6
New Scientist Archive | Selected Article

© Copyright Reed Business Information Ltd.

Você também pode gostar