SlideShare a Scribd company logo
1 of 216
Download to read offline
Preprint typeset in JHEP style - HYPER VERSION January 2009
String Theory
University of Cambridge Part III Mathematical Tripos
Dr David Tong
Department of Applied Mathematics and Theoretical Physics,
Centre for Mathematical Sciences,
Wilberforce Road,
Cambridge, CB3 OWA, UK
http://www.damtp.cam.ac.uk/user/tong/string.html
d.tong@damtp.cam.ac.uk
– 1 –
Recommended Books and Resources
• J. Polchinski, String Theory
This two volume work is the standard introduction to the subject. Our lectures will
more or less follow the path laid down in volume one covering the bosonic string. The
book contains explanations and descriptions of many details that have been deliberately
(and, I suspect, at times inadvertently) swept under a very large rug in these lectures.
Volume two covers the superstring.
• M. Green, J. Schwarz and E. Witten, Superstring Theory
Another two volume set. It is now over 20 years old and takes a slightly old-fashioned
route through the subject, with no explicit mention of conformal field theory. How-
ever, it does contain much good material and the explanations are uniformly excellent.
Volume one is most relevant for these lectures.
• B. Zwiebach, A First Course in String Theory
This book grew out of a course given to undergraduates who had no previous exposure
to general relativity or quantum field theory. It has wonderful pedagogical discussions
of the basics of lightcone quantization. More surprisingly, it also has some very clear
descriptions of several advanced topics, even though it misses out all the bits in between.
• P. Di Francesco, P. Mathieu and D. S´en´echal, Conformal Field Theory
This big yellow book is affectionately known as the yellow pages. It’s a great way
to learn conformal field theory. At first glance, it comes across as slightly daunting
because it’s big. (And yellow). But you soon realise that it’s big because it starts at
the beginning and provides detailed explanations at every step. The material necessary
for this course can be found in chapters 5 and 6.
Further References: “String Theory and M-Theory” by Becker, Becker and Schwarz
and “String Theory in a Nutshell” (it’s a big nutshell) by Kiritsis both deal with the
bosonic string fairly quickly, but include more advanced topics that may be of interest.
The book “D-Branes” by Johnson has lively and clear discussions about the many joys
of D-branes. Links to several excellent online resources, including video lectures by
Shiraz Minwalla, are listed on the course webpage.
Contents
0. Introduction 1
0.1 Quantum Gravity 3
1. The Relativistic String 9
1.1 The Relativistic Point Particle 9
1.1.1 Quantization 11
1.1.2 Ein Einbein 13
1.2 The Nambu-Goto Action 14
1.2.1 Symmetries of the Nambu-Goto Action 17
1.2.2 Equations of Motion 18
1.3 The Polyakov Action 18
1.3.1 Symmetries of the Polyakov Action 20
1.3.2 Fixing a Gauge 22
1.4 Mode Expansions 25
1.4.1 The Constraints Revisited 26
2. The Quantum String 28
2.1 A Lightning Look at Covariant Quantization 28
2.1.1 Ghosts 30
2.1.2 Constraints 30
2.2 Lightcone Quantization 32
2.2.1 Lightcone Gauge 33
2.2.2 Quantization 36
2.3 The String Spectrum 40
2.3.1 The Tachyon 40
2.3.2 The First Excited States 41
2.3.3 Higher Excited States 45
2.4 Lorentz Invariance Revisited 46
2.5 A Nod to the Superstring 48
3. Open Strings and D-Branes 50
3.1 Quantization 53
3.1.1 The Ground State 54
3.1.2 First Excited States: A World of Light 55
– 1 –
3.1.3 Higher Excited States and Regge Trajectories 56
3.1.4 Another Nod to the Superstring 56
3.2 Brane Dynamics: The Dirac Action 57
3.3 Multiple Branes: A World of Glue 59
4. Introducing Conformal Field Theory 61
4.0.1 Euclidean Space 62
4.0.2 The Holomorphy of Conformal Transformations 63
4.1 Classical Aspects 63
4.1.1 The Stress-Energy Tensor 64
4.1.2 Noether Currents 66
4.1.3 An Example: The Free Scalar Field 67
4.2 Quantum Aspects 68
4.2.1 Operator Product Expansion 68
4.2.2 Ward Identities 70
4.2.3 Primary Operators 73
4.3 An Example: The Free Scalar Field 77
4.3.1 The Propagator 77
4.3.2 An Aside: No Goldstone Bosons in Two Dimensions 79
4.3.3 The Stress-Energy Tensor and Primary Operators 80
4.4 The Central Charge 82
4.4.1 c is for Casimir 84
4.4.2 The Weyl Anomaly 86
4.4.3 c is for Cardy 89
4.4.4 c has a Theorem 91
4.5 The Virasoro Algebra 92
4.5.1 Radial Quantization 92
4.5.2 The Virasoro Algebra 94
4.5.3 Representations of the Virasoro Algebra 96
4.5.4 Consequences of Unitarity 98
4.6 The State-Operator Map 99
4.6.1 Some Simple Consequences 101
4.6.2 Our Favourite Example: The Free Scalar Field 102
4.7 Brief Comments on Conformal Field Theories with Boundaries 105
5. The Polyakov Path Integral and Ghosts 108
5.1 The Path Integral 108
5.1.1 The Faddeev-Popov Method 109
– 2 –
5.1.2 The Faddeev-Popov Determinant 112
5.1.3 Ghosts 113
5.2 The Ghost CFT 114
5.3 The Critical “Dimension” of String Theory 117
5.3.1 The Usual Nod to the Superstring 118
5.3.2 An Aside: Non-Critical Strings 119
5.4 States and Vertex Operators 120
5.4.1 An Example: Closed Strings in Flat Space 122
5.4.2 An Example: Open Strings in Flat Space 123
5.4.3 More General CFTs 124
6. String Interactions 125
6.1 What to Compute? 125
6.1.1 Summing Over Topologies 127
6.2 Closed String Amplitudes at Tree Level 130
6.2.1 Remnant Gauge Symmetry: SL(2,C) 130
6.2.2 The Virasoro-Shapiro Amplitude 132
6.2.3 Lessons to Learn 135
6.3 Open String Scattering 139
6.3.1 The Veneziano Amplitude 141
6.3.2 The Tension of D-Branes 142
6.4 One-Loop Amplitudes 143
6.4.1 The Moduli Space of the Torus 143
6.4.2 The One-Loop Partition Function 146
6.4.3 Interpreting the String Partition Function 149
6.4.4 So is String Theory Finite? 152
6.4.5 Beyond Perturbation Theory? 153
6.5 Appendix: Games with Integrals and Gamma Functions 154
7. Low Energy Effective Actions 157
7.1 Einstein’s Equations 158
7.1.1 The Beta Function 159
7.1.2 Ricci Flow 163
7.2 Other Couplings 163
7.2.1 Charged Strings and the B field 163
7.2.2 The Dilaton 165
7.2.3 Beta Functions 167
7.3 The Low-Energy Effective Action 167
– 3 –
7.3.1 String Frame and Einstein Frame 168
7.3.2 Corrections to Einstein’s Equations 170
7.3.3 Nodding Once More to the Superstring 171
7.4 Some Simple Solutions 173
7.4.1 Compactifications 174
7.4.2 The String Itself 175
7.4.3 Magnetic Branes 177
7.4.4 Moving Away from the Critical Dimension 180
7.4.5 The Elephant in the Room: The Tachyon 183
7.5 D-Branes Revisited: Background Gauge Fields 183
7.5.1 The Beta Function 184
7.5.2 The Born-Infeld Action 187
7.6 The DBI Action 188
7.6.1 Coupling to Closed String Fields 189
7.7 The Yang-Mills Action 191
7.7.1 D-Branes in Type II Superstring Theories 195
8. Compactification and T-Duality 197
8.1 The View from Spacetime 197
8.1.1 Moving around the Circle 199
8.2 The View from the Worldsheet 200
8.2.1 Massless States 202
8.2.2 Charged Fields 202
8.2.3 Enhanced Gauge Symmetry 203
8.3 Why Big Circles are the Same as Small Circles 204
8.3.1 A Path Integral Derivation of T-Duality 206
8.3.2 T-Duality for Open Strings 207
8.3.3 T-Duality for Superstrings 208
8.3.4 Mirror Symmetry 208
8.4 Epilogue 209
– 4 –
Acknowledgements
These lectures are aimed at beginning graduate students. They assume a working
knowledge of quantum field theory and general relativity. The lectures were given over
one semester and are based broadly on Volume one of the book by Joe Polchinski. I
inherited the course from Michael Green whose notes were extremely useful. I also
benefited enormously from the insightful and entertaining video lectures by Shiraz
Minwalla.
I’m grateful to Anirban Basu, Niklas Beisert, Joe Bhaseen, Diego Correa, Nick Dorey,
Michael Green, Anshuman Maharana, Malcolm Perry and Martin Schnabl for discus-
sions and help with various aspects of these notes. I’m also grateful to the students,
especially Carlos Guedes, for their excellent questions and superhuman typo-spotting
abilities. Finally, my thanks to Alex Considine for infinite patience and understanding
over the weeks these notes were written. I am supported by the Royal Society.
– 5 –
0. Introduction
String theory is an ambitious project. It purports to be an all-encompassing theory
of the universe, unifying the forces of Nature, including gravity, in a single quantum
mechanical framework.
The premise of string theory is that, at the fundamental level, matter does not consist
of point-particles but rather of tiny loops of string. From this slightly absurd beginning,
the laws of physics emerge. General relativity, electromagnetism and Yang-Mills gauge
theories all appear in a surprising fashion. However, they come with baggage. String
theory gives rise to a host of other ingredients, most strikingly extra spatial dimensions
of the universe beyond the three that we have observed. The purpose of this course is
to understand these statements in detail.
These lectures differ from most other courses that you will take in a physics degree.
String theory is speculative science. There is no experimental evidence that string
theory is the correct description of our world and scant hope that hard evidence will
arise in the near future. Moreover, string theory is very much a work in progress and
certain aspects of the theory are far from understood. Unresolved issues abound and
it seems likely that the final formulation has yet to be written. For these reasons, I’ll
begin this introduction by suggesting some answers to the question: Why study string
theory?
Reason 1. String theory is a theory of quantum gravity
String theory unifies Einstein’s theory of general relativity with quantum mechanics.
Moreover, it does so in a manner that retains the explicit connection with both quantum
theory and the low-energy description of spacetime.
But quantum gravity contains many puzzles, both technical and conceptual. What
does spacetime look like at the shortest distance scales? How can we understand
physics if the causal structure fluctuates quantum mechanically? Is the big bang truely
the beginning of time? Do singularities that arise in black holes really signify the end
of time? What is the microscopic origin of black hole entropy and what is it telling
us? What is the resolution to the information paradox? Some of these issues will be
reviewed later in this introduction.
Whether or not string theory is the true description of reality, it offers a framework
in which one can begin to explore these issues. For some questions, string theory
has given very impressive and compelling answers. For others, string theory has been
almost silent.
– 1 –
Reason 2. String theory may be the theory of quantum gravity
With broad brush, string theory looks like an extremely good candidate to describe the
real world. At low-energies it naturally gives rise to general relativity, gauge theories,
scalar fields and chiral fermions. In other words, it contains all the ingredients that
make up our universe. It also gives the only presently credible explanation for the value
of the cosmological constant although, in fairness, I should add that the explanation is
so distasteful to some that the community is rather amusingly split between whether
this is a good thing or a bad thing. Moreover, string theory incorporates several ideas
which do not yet have experimental evidence but which are considered to be likely
candidates for physics beyond the standard model. Prime examples are supersymmetry
and axions.
However, while the broad brush picture looks good, the finer details have yet to
be painted. String theory does not provide unique predictions for low-energy physics
but instead offers a bewildering array of possibilities, mostly dependent on what is
hidden in those extra dimensions. Partly, this problem is inherent to any theory of
quantum gravity: as we’ll review shortly, it’s a long way down from the Planck scale
to the domestic energy scales explored at the LHC. Using quantum gravity to extract
predictions for particle physics is akin to using QCD to extract predictions for how
coffee makers work. But the mere fact that it’s hard is little comfort if we’re looking
for convincing evidence that string theory describes the world in which we live.
While string theory cannot at present offer falsifiable predictions, it has nonetheless
inspired new and imaginative proposals for solving outstanding problems in particle
physics and cosmology. There are scenarios in which string theory might reveal itself
in forthcoming experiments. Perhaps we’ll find extra dimensions at the LHC, perhaps
we’ll see a network of fundamental strings stretched across the sky, or perhaps we’ll
detect some feature of non-Gaussianity in the CMB that is characteristic of D-branes
at work during inflation. My personal feeling however is that each of these is a long
shot and we may not know whether string theory is right or wrong within our lifetimes.
Of course, the history of physics is littered with naysayers, wrongly suggesting that
various theories will never be testable. With luck, I’ll be one of them.
Reason 3. String theory provides new perspectives on gauge theories
String theory was born from attempts to understand the strong force. Almost forty
years later, this remains one of the prime motivations for the subject. String theory
provides tools with which to analyze down-to-earth aspects of quantum field theory
that are far removed from high-falutin’ ideas about gravity and black holes.
– 2 –
Of immediate relevance to this course are the pedagogical reasons to invest time in
string theory. At heart, it is the study of conformal field theory and gauge symmetry.
The techniques that we’ll learn are not isolated to string theory, but apply to countless
systems which have direct application to real world physics.
On a deeper level, string theory provides new and very surprising methods to under-
stand aspects of quantum gauge theories. Of these, the most startling is the AdS/CFT
correspondence, first conjectured by Juan Maldacena, which gives a relationship be-
tween strongly coupled quantum field theories and gravity in higher dimensions. These
ideas have been applied in areas ranging from nuclear physics to condensed matter
physics and have provided qualitative (and arguably quantitative) insights into strongly
coupled phenomena.
Reason 4. String theory provides new results in mathematics
For the past 250 years, the close relationship between mathematics and physics has
been almost a one-way street: physicists borrowed many things from mathematicians
but, with a few noticeable exceptions, gave little back. In recent times, that has
changed. Ideas and techniques from string theory and quantum field theory have been
employed to give new “proofs” and, perhaps more importantly, suggest new directions
and insights in mathematics. The most well known of these is mirror symmetry, a
relationship between topologically different Calabi-Yau manifolds.
The four reasons described above also crudely characterize the string theory commu-
nity: there are “relativists” and “phenomenologists” and “field theorists” and “math-
ematicians”. Of course, the lines between these different sub-disciplines are not fixed
and one of the great attractions of string theory is its ability to bring together people
working in different areas — from cosmology to condensed matter to pure mathematics
— and provide a framework in which they can profitably communicate. In my opinion,
it is this cross-fertilization between fields which is the greatest strength of string theory.
0.1 Quantum Gravity
This is a starter course in string theory. Our focus will be on the perturbative approach
to the bosonic string and, in particular, why this gives a consistent theory of quantum
gravity. Before we leap into this, it is probably best to say a few words about quantum
gravity itself. Like why it’s hard. And why it’s important. (And why it’s not).
The Einstein Hilbert action is given by
SEH =
1
16πGN
d4
x
√
−gR
– 3 –
Newton’s constant GN can be written as
8πGN =
c
M2
pl
Throughout these lectures we work in units with = c = 1. The Planck mass Mpl
defines an energy scale
Mpl ≈ 2 × 1018
GeV .
(This is sometimes referred to as the reduced Planck mass, to distinguish it from the
scale without the factor of 8π, namely 1/GN ≈ 1 × 1019
GeV).
There are a couple of simple lessons that we can already take from this. The first is
that the relevant coupling in the quantum theory is 1/Mpl. To see that this is indeed
the case from the perspective of the action, we consider small perturbations around flat
Minkowski space,
gµν = ηµν +
1
Mpl
hµν
The factor of 1/Mpl is there to ensure that when we expand out the Einstein-Hilbert
action, the kinetic term for h is canonically normalized, meaning that it comes with no
powers of Mpl. This then gives the kind of theory that you met in your first course on
quantum field theory, albeit with an infinite series of interaction terms,
SEH = d4
x (∂h)2
+
1
Mpl
h (∂h)2
+
1
M2
pl
h2
(∂h)2
+ . . .
Each of these terms is schematic: if you were to do this explicitly, you would find a
mess of indices contracted in different ways. We see that the interactions are suppressed
by powers of Mpl. This means that quantum perturbation theory is an expansion in
the dimensionless ratio E2
/M2
pl, where E is the energy associated to the process of
interest. We learn that gravity is weak, and therefore under control, at low-energies.
But gravitational interactions become strong as the energy involved approaches the
Planck scale. In the language of the renormalization group, couplings of this type are
known as irrelevant.
The second lesson to take away is that the Planck scale Mpl is very very large. The
LHC will probe the electroweak scale, MEW ∼ 103
GeV. The ratio is MEW /Mpl ∼ 10−15
.
For this reason, quantum gravity will not affect your daily life, even if your daily life
involves the study of the most extreme observable conditions in the universe.
– 4 –
Gravity is Non-Renormalizable
Quantum field theories with irrelevant couplings are typically ill-behaved at high-
energies, rendering the theory ill-defined. Gravity is no exception. Theories of this
type are called non-renormalizable, which means that the divergences that appear in
the Feynman diagram expansion cannot be absorbed by a finite number of countert-
erms. In pure Einstein gravity, the symmetries of the theory are enough to ensure that
the one-loop S-matrix is finite. The first divergence occurs at two-loops and requires
the introduction of a counterterm of the form,
Γ ∼
1 1
M4
pl
d4
x
√
−g Rµν
ρσRρσ
λκRλκ
µν
with = 4−D. All indications point towards the fact that this is the first in an infinite
number of necessary counterterms.
Coupling gravity to matter requires an interaction term of the form,
Sint = d4
x
1
Mpl
hµνTµν
+ O(h2
)
This makes the situation marginally worse, with the first diver-
Figure 1:
gence now appearing at one-loop. The Feynman diagram in the
figure shows particle scattering through the exchange of two gravi-
tons. When the momentum k running in the loop is large, the
diagram is badly divergent: it scales as
1
M4
pl
∞
d4
k
Non-renormalizable theories are commonplace in the history of physics, the most com-
monly cited example being Fermi’s theory of the weak interaction. The first thing to say
about them is that they are far from useless! Non-renormalizable theories are typically
viewed as effective field theories, valid only up to some energy scale Λ. One deals with
the divergences by simply admitting ignorance beyond this scale and treating Λ as a
UV cut-off on any momentum integral. In this way, we get results which are valid to an
accuracy of E/Λ (perhaps raised to some power). In the case of the weak interaction,
Fermi’s theory accurately predicts physics up to an energy scale of 1/GF ∼ 100 GeV.
In the case of quantum gravity, Einstein’s theory works to an accuracy of (E/Mpl)2
.
– 5 –
However, non-renormalizable theories are typically unable to describe physics at their
cut-off scale Λ or beyond. This is because they are missing the true ultra-violet degrees
of freedom which tame the high-energy behaviour. In the case of the weak force, these
new degrees of freedom are the W and Z bosons. We would like to know what missing
degrees of freedom are needed to complete gravity.
Singularities
Only a particle physicist would phrase all questions about the universe in terms of
scattering amplitudes. In general relativity we typically think about the geometry as
a whole, rather than bastardizing the Einstein-Hilbert action and discussing perturba-
tions around flat space. In this language, the question of high-energy physics turns into
one of short distance physics. Classical general relativity is not to be trusted in regions
where the curvature of spacetime approaches the Planck scale and ultimately becomes
singular. A quantum theory of gravity should resolve these singularities.
The question of spacetime singularities is morally equivalent to that of high-energy
scattering. Both probe the ultra-violet nature of gravity. A spacetime geometry is
made of a coherent collection of gravitons, just as the electric and magnetic fields in a
laser are made from a collection of photons. The short distance structure of spacetime
is governed – after Fourier transform – by high momentum gravitons. Understanding
spacetime singularities and high-energy scattering are different sides of the same coin.
There are two situations in general relativity where singularity theorems tell us that
the curvature of spacetime gets large: at the big bang and in the center of a black hole.
These provide two of the biggest challenges to any putative theory of quantum gravity.
Gravity is Subtle
It is often said that general relativity contains the seeds of its own destruction. The
theory is unable to predict physics at the Planck scale and freely admits to it. Problems
such as non-renormalizability and singularities are, in a Rumsfeldian sense, known
unknowns. However, the full story is more complicated and subtle. On the one hand,
the issue of non-renormalizability may not quite be the crisis that it first appears. On
the other hand, some aspects of quantum gravity suggest that general relativity isn’t
as honest about its own failings as is usually advertised. The theory hosts a number of
unknown unknowns, things that we didn’t even know that we didn’t know. We won’t
have a whole lot to say about these issues in this course, but you should be aware of
them. Here I mention only a few salient points.
– 6 –
Firstly, there is a key difference between Fermi’s theory of the weak interaction and
gravity. Fermi’s theory was unable to provide predictions for any scattering process
at energies above 1/GF . In contrast, if we scatter two objects at extremely high-
energies in gravity — say, at energies E Mpl — then we know exactly what will
happen: we form a big black hole. We don’t need quantum gravity to tell us this.
Classical general relativity is sufficient. If we restrict attention to scattering, the crisis
of non-renormalizability is not problematic at ultra-high energies. It’s troublesome only
within a window of energies around the Planck scale.
Similar caveats hold for singularities. If you are foolish enough to jump into a black
hole, then you’re on your own: without a theory of quantum gravity, no one can tell you
what fate lies in store at the singularity. Yet, if you are smart and stay outside of the
black hole, you’ll be hard pushed to see any effects of quantum gravity. This is because
Nature has conspired to hide Planck scale curvatures from our inquisitive eyes. In the
case of black holes this is achieved through cosmic censorship which is a conjecture in
classical general relativity that says singularities are hidden behind horizons. In the
case of the big bang, it is achieved through inflation, washing away any traces from the
very early universe. Nature appears to shield us from the effects of quantum gravity,
whether in high-energy scattering or in singularities. I think it’s fair to say that no one
knows if this conspiracy is pointing at something deep, or is merely inconvenient for
scientists trying to probe the Planck scale.
While horizons may protect us from the worst excesses of singularities, they come
with problems of their own. These are the unknown unknowns: difficulties that arise
when curvatures are small and general relativity says “trust me”. The entropy of black
holes and the associated paradox of information loss strongly suggest that local quan-
tum field theory breaks down at macroscopic distance scales. Attempts to formulate
quantum gravity in de Sitter space, or in the presence of eternal inflation, hint at similar
difficulties. Ideas of holography, black hole complimentarity and the AdS/CFT corre-
spondence all point towards non-local effects and the emergence of spacetime. These are
the deep puzzles of quantum gravity and their relationship to the ultra-violet properties
of gravity is unclear.
As a final thought, let me mention the one observation that has an outside chance of
being related to quantum gravity: the cosmological constant. With an energy scale of
Λ ∼ 10−3
eV it appears to have little to do with ultra-violet physics. If it does have its
origins in a theory of quantum gravity, it must either be due to some subtle “unknown
unknown”, or because it is explained away as an environmental quantity as in string
theory.
– 7 –
Is the Time Ripe?
Our current understanding of physics, embodied in the standard model, is valid up to
energy scales of 103
GeV. This is 15 orders of magnitude away from the Planck scale.
Why do we think the time is now ripe to tackle quantum gravity? Surely we are like
the ancient Greeks arguing about atomism. Why on earth do we believe that we’ve
developed the right tools to even address the question?
The honest answer, I think, is hubris.
Figure 2:
However, there is mild circumstantial evidence
that the framework of quantum field theory might
hold all the way to the Planck scale without any-
thing very dramatic happening in between. The
main argument is unification. The three coupling
constants of Nature run logarithmically, meeting
miraculously at the GUT energy scale of 1015
GeV.
Just slightly later, the fourth force of Nature, grav-
ity, joins them. While not overwhelming, this does
provide a hint that perhaps quantum field theory
can be taken seriously at these ridiculous scales.
Historically I suspect this was what convinced large parts of the community that it was
ok to speak about processes at 1018
GeV.
Finally, perhaps the most compelling argument for studying physics at the Planck
scale is that string theory does provide a consistent unified quantum theory of gravity
and the other forces. Given that we have this theory sitting in our laps, it would be
foolish not to explore its consequences. The purpose of these lecture notes is to begin
this journey.
– 8 –
1. The Relativistic String
All lecture courses on string theory start with a discussion of the point particle. Ours
is no exception. We’ll take a flying tour through the physics of the relativistic point
particle and extract a couple of important lessons that we’ll take with us as we move
onto string theory.
1.1 The Relativistic Point Particle
We want to write down the Lagrangian describing a relativistic particle of mass m.
In anticipation of string theory, we’ll consider D-dimensional Minkowski space R1,D−1
.
Throughout these notes, we work with signature
ηµν = diag(−1, +1, +1, . . . , +1)
Note that this is the opposite signature to my quantum field theory notes.
If we fix a frame with coordinates Xµ
= (t, x) the action is simple:
S = −m dt 1 − ˙x · ˙x . (1.1)
To see that this is correct we can compute the momentum p, conjugate to x, and the
energy E which is equal to the Hamiltonian,
p =
m˙x
1 − ˙x · ˙x
, E = m2 + p2 ,
both of which should be familiar from courses on special relativity.
Although the Lagrangian (1.1) is correct, it’s not fully satisfactory. The reason is
that time t and space x play very different roles in this Lagrangian. The position x is
a dynamical degree of freedom. In contrast, time t is merely a parameter providing a
label for the position. Yet Lorentz transformations are supposed to mix up t and x and
such symmetries are not completely obvious in (1.1). Can we find a new Lagrangian
in which time and space are on equal footing?
One possibility is to treat both time and space as labels. This leads us to the
concept of field theory. However, in this course we will be more interested in the other
possibility: we will promote time to a dynamical degree of freedom. At first glance,
this may appear odd: the number of degrees of freedom is one of the crudest ways we
have to characterize a system. We shouldn’t be able to add more degrees of freedom
– 9 –
at will without fundamentally changing the system that we’re talking about. Another
way of saying this is that the particle has the option to move in space, but it doesn’t
have the option to move in time. It has to move in time. So we somehow need a way
to promote time to a degree of freedom without it really being a true dynamical degree
of freedom! How do we do this? The answer, as we will now show, is gauge symmetry.
Consider the action, X
X
0
Figure 3:
S = −m dτ − ˙Xµ ˙Xνηµν , (1.2)
where µ = 0, . . . , D − 1 and ˙Xµ
= dXµ
/dτ. We’ve introduced a
new parameter τ which labels the position along the worldline of
the particle as shown by the dashed lines in the figure. This action
has a simple interpretation: it is just the proper time ds along the
worldline.
Naively it looks as if we now have D physical degrees of freedom rather than D − 1
because, as promised, the time direction X0
≡ t is among our dynamical variables:
X0
= X0
(τ). However, this is an illusion. To see why, we need to note that the action
(1.2) has a very important property: reparameterization invariance. This means that
we can pick a different parameter ˜τ on the worldline, related to τ by any monotonic
function
˜τ = ˜τ(τ) .
Let’s check that the action is invariant under transformations of this type. The inte-
gration measure in the action changes as dτ = d˜τ |dτ/d˜τ|. Meanwhile, the velocities
change as dXµ
/dτ = (dXµ
/d˜τ) (d˜τ/dτ). Putting this together, we see that the action
can just as well be written in the ˜τ reparameterization,
S = −m d˜τ −
dXµ
d˜τ
dXν
d˜τ
ηµν .
The upshot of this is that not all D degrees of freedom Xµ
are physical. For example,
suppose you find a solution to this system, so that you know how X0
changes with
τ and how X1
changes with τ and so on. Not all of that information is meaningful
because τ itself is not meaningful. In particular, we could use our reparameterization
invariance to simply set
τ = X0
(τ) ≡ t (1.3)
– 10 –
If we plug this choice into the action (1.2) then we recover our initial action (1.1). The
reparameterization invariance is a gauge symmetry of the system. Like all gauge sym-
metries, it’s not really a symmetry at all. Rather, it is a redundancy in our description.
In the present case, it means that although we seem to have D degrees of freedom Xµ
,
one of them is fake.
The fact that one of the degrees of freedom is a fake also shows up if we look at the
momenta,
pµ =
∂L
∂ ˙Xµ
=
m ˙Xν
ηµν
− ˙Xλ ˙Xρ ηλρ
(1.4)
These momenta aren’t all independent. They satisfy
pµpµ
+ m2
= 0 (1.5)
This is a constraint on the system. It is, of course, the mass-shell constraint for a
relativistic particle of mass m. From the worldline perspective, it tells us that the
particle isn’t allowed to sit still in Minkowski space: at the very least, it had better
keep moving in a timelike direction with (p0
)2
≥ m2
.
One advantage of the action (1.2) is that the Poincar´e symmetry of the particle is
now manifest, appearing as a global symmetry on the worldline
Xµ
→ Λµ
νXν
+ cµ
(1.6)
where Λ is a Lorentz transformation satisfying Λµ
νηνρ
Λσ
ρ = ηµσ
, while cµ
corresponds
to a constant translation. We have made all the symmetries manifest at the price of
introducing a gauge symmetry into our system. A similar gauge symmetry will arise
in the relativistic string and much of this course will be devoted to understanding its
consequences.
1.1.1 Quantization
It’s a trivial matter to quantize this action. We introduce a wavefunction Ψ(X). This
satisfies the usual Schr¨odinger equation,
i
∂Ψ
∂τ
= HΨ .
But, computing the Hamiltonian H = ˙Xµ
pµ −L, we find that it vanishes: H = 0. This
shouldn’t be surprising. It is simply telling us that the wavefunction doesn’t depend on
– 11 –
τ. Since the wavefunction is something physical while, as we have seen, τ is not, this is
to be expected. Note that this doesn’t mean that time has dropped out of the problem.
On the contrary, in this relativistic context, time X0
is an operator, just like the spatial
coordinates x. This means that the wavefunction Ψ is immediately a function of space
and time. It is not like a static state in quantum mechanics, but more akin to the fully
integrated solution to the non-relativistic Schr¨odinger equation.
The classical system has a constraint given by (1.5). In the quantum theory, we
impose this constraint as an operator equation on the wavefunction, namely (pµ
pµ +
m2
)Ψ = 0. Using the usual representation of the momentum operator pµ = −i∂/∂Xµ
,
we recognize this constraint as the Klein-Gordon equation
−
∂
∂Xµ
∂
∂Xν
ηµν
+ m2
Ψ(X) = 0 (1.7)
Although this equation is familiar from field theory, it’s important to realize that the
interpretation is somewhat different. In relativistic field theory, the Klein-Gordon equa-
tion is the equation of motion obeyed by a scalar field. In relativistic quantum mechan-
ics, it is the equation obeyed by the wavefunction. In the early days of field theory,
the fact that these two equations are the same led people to think one should view
the wavefunction as a classical field and quantize it a second time. This isn’t cor-
rect, but nonetheless the language has stuck and it is common to talk about the point
particle perspective as “first quantization” and the field theory perspective as “second
quantization”.
So far we’ve considered only a free point particle. How can we
Figure 4:
introduce interactions into this framework? We would have to first
decide which interactions are allowed: perhaps the particle can split
into two; perhaps it can fuse with other particles? Obviously, there is
a huge range of options for us to choose from. We would then assign
amplitudes for these processes to happen. There would be certain
restrictions coming from the requirement of unitarity which, among
other things, would lead to the necessity of anti-particles. We could draw diagrams
associated to the different interactions — an example is given in the figure — and in
this manner we would slowly build up the Feynman diagram expansion that is familiar
from field theory. In fact, this was pretty much the way Feynman himself approached
the topic of QED. However, in practice we rarely construct particle interactions in
this way because the field theory framework provides a much better way of looking at
things. In contrast, this way of building up interactions is exactly what we will later
do for strings.
– 12 –
1.1.2 Ein Einbein
There is another action that describes the relativistic point particle. We introduce yet
another field on the worldline, e(τ), and write
S =
1
2
dτ e−1 ˙X2
− em2
, (1.8)
where we’ve used the notation ˙X2
= ˙Xµ ˙Xν
ηµν. For the rest of these lectures, terms
like X2
will always mean an implicit contraction with the spacetime Minkowski metric.
This form of the action makes it look as if we have coupled the worldline theory to
1d gravity, with the field e(τ) acting as an einbein (in the sense of vierbeins that are
introduced in general relativity). To see this, note that we could change notation and
write this action in the more suggestive form
S = −
1
2
dτ
√
−gττ gττ ˙X2
+ m2
. (1.9)
where gττ = (gττ
)−1
is the metric on the worldline and e =
√
−gττ
Although our action appears to have one more degree of freedom, e, it can be easily
checked that it has the same equations of motion as (1.2). The reason for this is that
e is completely fixed by its equation of motion, ˙X2
+ e2
m2
= 0. Substituting this into
the action (1.8) recovers (1.2)
The action (1.8) has a couple of advantages over (1.2). Firstly, it works for massless
particles with m = 0. Secondly, the absence of the annoying square root means that
it’s easier to quantize in a path integral framework.
The action (1.8) retains invariance under reparameterizations which are now written
in a form that looks more like general relativity. For transformations parameterized by
an infinitesimal η, we have
τ → ˜τ = τ − η(τ) , δe =
d
dτ
(η(τ)e) , δXµ
=
dXµ
dτ
η(τ) (1.10)
The einbein e transforms as a density on the worldline, while each of the coordinates
Xµ
transforms as a worldline scalar.
– 13 –
1.2 The Nambu-Goto Action
A particle sweeps out a worldline in Minkowski space. A string
σ
τ
Figure 5:
sweeps out a worldsheet. We’ll parameterize this worldsheet by
one timelike coordinate τ, and one spacelike coordinate σ. In this
section we’ll focus on closed strings and take σ to be periodic,
with range
σ ∈ [0, 2π) . (1.11)
We will sometimes package the two worldsheet coordinates to-
gether as σα
= (τ, σ), α = 0, 1. Then the string sweeps out a
surface in spacetime which defines a map from the worldsheet to
Minkowski space, Xµ
(σ, τ) with µ = 0, . . . , D − 1. For closed strings, we require
Xµ
(σ, τ) = Xµ
(σ + 2π, τ) .
In this context, spacetime is sometimes referred to as the target space to distinguish it
from the worldsheet.
We need an action that describes the dynamics of this string. The key property
that we will ask for is that nothing depends on the coordinates σα
that we choose
on the worldsheet. In other words, the string action should be reparameterization
invariant. What kind of action does the trick? Well, for the point particle the action
was proportional to the length of the worldline. The obvious generalization is that the
action for the string should be proportional to the area, A, of the worldsheet. This
is certainly a property that is characteristic of the worldsheet itself, rather than any
choice of parameterization.
How do we find the area A in terms of the coordinates Xµ
(σ, τ)? The worldsheet is
a curved surface embedded in spacetime. The induced metric, γαβ, on this surface is
the pull-back of the flat metric on Minkowski space,
γαβ =
∂Xµ
∂σα
∂Xν
∂σβ
ηµν . (1.12)
Then the action which is proportional to the area of the worldsheet is given by,
S = −T d2
σ − det γ . (1.13)
Here T is a constant of proportionality. We will see shortly that it is the tension of the
string, meaning the mass per unit length.
– 14 –
We can write this action a little more explicitly. The pull-back of the metric is given
by,
γαβ =
˙X2 ˙X · X
˙X · X X 2
.
where ˙Xµ
= ∂Xµ
/∂τ and Xµ
= ∂Xµ
/∂σ. The action then takes the form,
S = −T d2
σ −( ˙X)2 (X )2 + ( ˙X · X )2 . (1.14)
This is the Nambu-Goto action for a relativistic string.
Action = Area: A Check
If you’re unfamiliar with differential geometry, the argu-
dl
dl
1
2
τ
σ
Figure 6:
ment about the pull-back of the metric may be a bit slick.
Thankfully, there’s a more pedestrian way to see that the
action (1.14) is equal to the area swept out by the world-
sheet. It’s slightly simpler to make this argument for a sur-
face embedded in Euclidean space rather than Minkowski
space. We choose some parameterization of the sheet in
terms of τ and σ, as drawn in the figure, and we write the
coordinates of Euclidean space as X(σ, τ). We’ll compute the area of the infinitesimal
shaded region. The vectors tangent to the boundary are,
dl1 =
∂X
∂σ
, dl2 =
∂X
∂τ
.
If the angle between these two vectors is θ, then the area is then given by
ds2
= |dl1||dl2| sin θ = dl2
1 dl2
2(1 − cos2 θ) = dl2
1 dl2
2 − (dl1 · dl2)2 (1.15)
which indeed takes the form of the integrand of (1.14).
Tension and Dimension
Let’s now see that T has the physical interpretation of tension. We write Minkowski
coordinates as Xµ
= (t, x). We work in a gauge with X0
≡ t = Rτ, where R is a
constant that is needed to balance up dimensions (see below) and will drop out at the
end of the argument. Consider a snapshot of a string configuration at a time when
– 15 –
dx/dτ = 0 so that the instantaneous kinetic energy vanishes. Evaluating the action for
a time dt gives
S = −T dτdσR (dx/dσ)2 = −T dt (spatial length of string) . (1.16)
But, when the kinetic energy vanishes, the action is proportional to the time integral
of the potential energy,
potential energy = T × (spatial length of string) .
So T is indeed the energy per unit length as claimed. We learn that the string acts
rather like an elastic band and its energy increases linearly with length. (This is different
from the elastic bands you’re used to which obey Hooke’s law where energy increased
quadratically with length). To minimize its potential energy, the string will want to
shrink to zero size. We’ll see that when we include quantum effects this can’t happen
because of the usual zero point energies.
There is a slightly annoying way of writing the tension that has its origin in ancient
history, but is commonly used today
T =
1
2πα
(1.17)
where α is pronounced “alpha-prime”. In the language of our ancestors, α is referred
to as the “universal Regge slope”. We’ll explain why later in this course.
At this point, it’s worth pointing out some conventions that we have, until now,
left implicit. The spacetime coordinates have dimension [X] = −1. In contrast, the
worldsheet coordinates are taken to be dimensionless, [σ] = 0. (This can be seen in our
identification σ ≡ σ + 2π). The tension is equal to the mass per unit length and has
dimension [T] = 2. Obviously this means that [α ] = −2. We can therefore associate a
length scale, ls, by
α = l2
s (1.18)
The string scale ls is the natural length that appears in string theory. In fact, in a
certain sense (that we will make more precise later in the course) this length scale is
the only parameter of the theory.
– 16 –
Actual Strings vs. Fundamental Strings
There are several situations in Nature where string-like objects arise. Prime examples
include magnetic flux tubes in superconductors and chromo-electric flux tubes in QCD.
Cosmic strings, a popular speculation in cosmology, are similar objects, stretched across
the sky. In each of these situations, there are typically two length scales associated to
the string: the tension, T and the width of the string, L. For all these objects, the
dynamics is governed by the Nambu-Goto action as long as the curvature of the string is
much greater than L. (In the case of superconductors, one should work with a suitable
non-relativistic version of the Nambu-Goto action).
However, in each of these other cases, the Nambu-Goto action is not the end of the
story. There will typically be additional terms in the action that depend on the width
of the string. The form of these terms is not universal, but often includes a rigidity
piece of form L K2
, where K is the extrinsic curvature of the worldsheet. Other
terms could be added to describe fluctuations in the width of the string.
The string scale, ls, or equivalently the tension, T, depends on the kind of string that
we’re considering. For example, if we’re interested in QCD flux tubes then we would
take
T ∼ (1 GeV)2
(1.19)
In this course we will consider fundamental strings which have zero width. What this
means in practice is that we take the Nambu-Goto action as the complete description
for all configurations of the string. These strings will have relevance to quantum gravity
and the tension of the string is taken to be much larger, typically an order of magnitude
or so below the Planck scale.
T M2
pl = (1018
GeV)2
(1.20)
However, I should point out that when we try to view string theory as a fundamental
theory of quantum gravity, we don’t really know what value T should take. As we
will see later in this course, it depends on many other aspects, most notably the string
coupling and the volume of the extra dimensions.
1.2.1 Symmetries of the Nambu-Goto Action
The Nambu-Goto action has two types of symmetry, each of a different nature.
• Poincar´e invariance of the spacetime (1.6). This is a global symmetry from the
perspective of the worldsheet, meaning that the parameters Λµ
ν and cµ
which label
– 17 –
the symmetry transformation are constants and do not depend on worldsheet
coordinates σα
.
• Reparameterization invariance, σα
→ ˜σα
(σ). As for the point particle, this is a
gauge symmetry. It reflects the fact that we have a redundancy in our description
because the worldsheet coordinates σα
have no physical meaning.
1.2.2 Equations of Motion
To derive the equations of motion for the Nambu-Goto string, we first introduce the
momenta which we call Π because there will be countless other quantities that we want
to call p later,
Πτ
µ =
∂L
∂ ˙Xµ
= −T
( ˙X · X )Xµ − (X 2
) ˙Xµ
( ˙X · X )2 − ˙X2 X 2
.
Πσ
µ =
∂L
∂X µ
= −T
( ˙X · X ) ˙Xµ − ( ˙X2
)Xµ
( ˙X · X )2 − ˙X2 X 2
.
The equations of motion are then given by,
∂Πτ
µ
∂τ
+
∂Πσ
µ
∂σ
= 0
These look like nasty, non-linear equations. In fact, there’s a slightly nicer way to write
these equations, starting from the earlier action (1.13). Recall that the variation of a
determinant is δ
√
−γ = 1
2
√
−γ γαβ
δγαβ. Using the definition of the pull-back metric
γαβ, this gives rise to the equations of motion
∂α( − det γ γαβ
∂βXµ
) = 0 , (1.21)
Although this notation makes the equations look a little nicer, we’re kidding ourselves.
Written in terms of Xµ
, they are still the same equations. Still nasty.
1.3 The Polyakov Action
The square-root in the Nambu-Goto action means that it’s rather difficult to quantize
using path integral techniques. However, there is another form of the string action
which is classically equivalent to the Nambu-Goto action. It eliminates the square root
at the expense of introducing another field,
S = −
1
4πα
d2
σ
√
−ggαβ
∂αXµ
∂βXν
ηµν (1.22)
where g ≡ det g. This is the Polyakov action. (Polyakov didn’t discover the action, but
he understood how to work with it in the path integral and for this reason it carries
his name. The path integral treatment of this action will be the subject of Chapter 5).
– 18 –
The new field is gαβ. It is a dynamical metric on the worldsheet. From the perspective
of the worldsheet, the Polyakov action is a bunch of scalar fields X coupled to 2d gravity.
The equation of motion for Xµ
is
∂α(
√
−ggαβ
∂βXµ
) = 0 , (1.23)
which coincides with the equation of motion (1.21) from the Nambu-Goto action, except
that gαβ is now an independent variable which is fixed by its own equation of motion. To
determine this, we vary the action (remembering again that δ
√
−g = −1
2
√
−ggαβδgαβ
=
+1
2
√
−ggαβ
δgαβ),
δS = −
T
2
d2
σ δgαβ
√
−g ∂αXµ
∂βXν
− 1
2
√
−g gαβgρσ
∂ρXµ
∂σXν
ηµν = 0 .(1.24)
The worldsheet metric is therefore given by,
gαβ = 2f(σ) ∂αX · ∂βX , (1.25)
where the function f(σ) is given by,
f−1
= gρσ
∂ρX · ∂σX
A comment on the potentially ambiguous notation: here, and below, any function f(σ)
is always short-hand for f(σ, τ): it in no way implies that f depends only on the spatial
worldsheet coordinate.
We see that gαβ isn’t quite the same as the pull-back metric γαβ defined in equation
(1.12); the two differ by the conformal factor f. However, this doesn’t matter because,
rather remarkably, f drops out of the equation of motion (1.23). This is because the
√
−g term scales as f, while the inverse metric gαβ
scales as f−1
and the two pieces
cancel. We therefore see that Nambu-Goto and the Polyakov actions result in the same
equation of motion for X.
In fact, we can see more directly that the Nambu-Goto and Polyakov actions coincide.
We may replace gαβ in the Polyakov action (1.22) with its equation of motion gαβ =
2f γαβ. The factor of f also drops out of the action for the same reason that it dropped
out of the equation of motion. In this manner, we recover the Nambu-Goto action
(1.13).
– 19 –
1.3.1 Symmetries of the Polyakov Action
The fact that the presence of the factor f(σ, τ) in (1.25) didn’t actually affect the
equations of motion for Xµ
reflects the existence of an extra symmetry which the
Polyakov action enjoys. Let’s look more closely at this. Firstly, the Polyakov action
still has the two symmetries of the Nambu-Goto action,
• Poincar´e invariance. This is a global symmetry on the worldsheet.
Xµ
→ Λµ
νXν
+ cµ
.
• Reparameterization invariance, also known as diffeomorphisms. This is a gauge
symmetry on the worldsheet. We may redefine the worldsheet coordinates as
σα
→ ˜σα
(σ). The fields Xµ
transform as worldsheet scalars, while gαβ transforms
in the manner appropriate for a 2d metric.
Xµ
(σ) → ˜Xµ
(˜σ) = Xµ
(σ)
gαβ(σ) → ˜gαβ(˜σ) =
∂σγ
∂˜σα
∂σδ
∂˜σβ
gγδ(σ)
It will sometimes be useful to work infinitesimally. If we make the coordinate
change σα
→ ˜σα
= σα
− ηα
(σ), for some small η. The transformations of the
fields then become,
δXµ
(σ) = ηα
∂αXµ
δgαβ(σ) = αηβ + βηα
where the covariant derivative is defined by αηβ = ∂αηβ − Γσ
αβησ with the Levi-
Civita connection associated to the worldsheet metric given by the usual expres-
sion,
Γσ
αβ = 1
2
gσρ
(∂αgβρ + ∂βgρα − ∂ρgαβ)
Together with these familiar symmetries, there is also a new symmetry which is novel
to the Polyakov action. It is called Weyl invariance.
• Weyl Invariance. Under this symmetry, Xµ
(σ) → Xµ
(σ), while the metric
changes as
gαβ(σ) → Ω2
(σ) gαβ(σ) . (1.26)
Or, infinitesimally, we can write Ω2
(σ) = e2φ(σ)
for small φ so that
δgαβ(σ) = 2φ(σ) gαβ(σ) .
– 20 –
It is simple to see that the Polyakov action is invariant under this transformation:
the factor of Ω2
drops out just as the factor of f did in equation (1.25), canceling
between
√
−g and the inverse metric gαβ
. This is a gauge symmetry of the string,
as seen by the fact that the parameter Ω depends on the worldsheet coordinates
σ. This means that two metrics which are related by a Weyl transformation (1.26)
are to be considered as the same physical state.
Figure 7: An example of a Weyl transformation
How should we think of Weyl invariance? It is not a coordinate change. Instead it is
the invariance of the theory under a local change of scale which preserves the angles
between all lines. For example the two worldsheet metrics shown in the figure are
viewed by the Polyakov string as equivalent. This is rather surprising! And, as you
might imagine, theories with this property are extremely rare. It should be clear from
the discussion above that the property of Weyl invariance is special to two dimensions,
for only there does the scaling factor coming from the determinant
√
−g cancel that
coming from the inverse metric. But even in two dimensions, if we wish to keep Weyl
invariance then we are strictly limited in the kind of interactions that can be added to
the action. For example, we would not be allowed a potential term for the worldsheet
scalars of the form,
d2
σ
√
−g V (X) .
These break Weyl invariance. Nor can we add a worldsheet cosmological constant term,
µ d2
σ
√
−g .
This too breaks Weyl invariance. We will see later in this course that the requirement
of Weyl invariance becomes even more stringent in the quantum theory. We will also
see what kind of interactions terms can be added to the worldsheet. Indeed, much of
this course can be thought of as the study of theories with Weyl invariance.
– 21 –
1.3.2 Fixing a Gauge
As we have seen, the equation of motion (1.23) looks pretty nasty. However, we can use
the redundancy inherent in the gauge symmetry to choose coordinates in which they
simplify. Let’s think about what we can do with the gauge symmetry.
Firstly, we have two reparameterizations to play with. The worldsheet metric has
three independent components. This means that we expect to be able to set any two of
the metric components to a value of our choosing. We will choose to make the metric
locally conformally flat, meaning
gαβ = e2φ
ηαβ , (1.27)
where φ(σ, τ) is some function on the worldsheet. You can check that this is possible
by writing down the change of the metric under a coordinate transformation and seeing
that the differential equations which result from the condition (1.27) have solutions, at
least locally. Choosing a metric of the form (1.27) is known as conformal gauge.
We have only used reparameterization invariance to get to the metric (1.27). We still
have Weyl transformations to play with. Clearly, we can use these to remove the last
independent component of the metric and set φ = 0 such that,
gαβ = ηαβ . (1.28)
We end up with the flat metric on the worldsheet in Minkowski coordinates.
A Diversion: How to make a metric flat
The fact that we can use Weyl invariance to make any two-dimensional metric flat is
an important result. Let’s take a quick diversion from our main discussion to see a
different proof that isn’t tied to the choice of Minkowski coordinates on the worldsheet.
We’ll work in 2d Euclidean space to avoid annoying minus signs. Consider two metrics
related by a Weyl transformation, gαβ = e2φ
gαβ. One can check that the Ricci scalars
of the two metrics are related by,
g R =
√
g(R − 2 2
φ) . (1.29)
We can therefore pick a φ such that the new metric has vanishing Ricci scalar, R = 0,
simply by solving this differential equation for φ. However, in two dimensions (but
not in higher dimensions) a vanishing Ricci scalar implies a flat metric. The reason is
simply that there aren’t too many indices to play with. In particular, symmetry of the
Riemann tensor in two dimensions means that it must take the form,
Rαβγδ =
R
2
(gαγgβδ − gαδgβγ) .
– 22 –
So R = 0 is enough to ensure that Rαβγδ = 0, which means that the manifold is flat. In
equation (1.28), we’ve further used reparameterization invariance to pick coordinates
in which the flat metric is the Minkowski metric.
The equations of motion and the stress-energy tensor
With the choice of the flat metric (1.28), the Polyakov action simplifies tremendously
and becomes the theory of D free scalar fields. (In fact, this simplification happens in
any conformal gauge).
S = −
1
4πα
d2
σ ∂αX · ∂α
X , (1.30)
and the equations of motion for Xµ
reduce to the free wave equation,
∂α∂α
Xµ
= 0 . (1.31)
Now that looks too good to be true! Are the horrible equations (1.23) really equivalent
to a free wave equation? Well, not quite. There is something that we’ve forgotten:
we picked a choice of gauge for the metric gαβ. But we must still make sure that the
equation of motion for gαβ is satisfied. In fact, the variation of the action with respect
to the metric gives rise to a rather special quantity: it is the stress-energy tensor, Tαβ.
With a particular choice of normalization convention, we define the stress-energy tensor
to be
Tαβ = −
2
T
1
√
−g
∂S
∂gαβ
.
We varied the Polyakov action with respect to gαβ in (1.24). When we set gαβ = ηαβ
we get
Tαβ = ∂αX · ∂βX − 1
2
ηαβ ηρσ
∂ρX · ∂σX . (1.32)
The equation of motion associated to the metric gαβ is simply Tαβ = 0. Or, more
explicitly,
T01 = ˙X · X = 0
T00 = T11 = 1
2
( ˙X2
+ X 2
) = 0 . (1.33)
We therefore learn that the equations of motion of the string are the free wave equations
(1.31) subject to the two constraints (1.33) arising from the equation of motion Tαβ = 0.
– 23 –
Getting a feel for the constraints
Let’s try to get some intuition for these constraints. There is a simple
Figure 8:
meaning of the first constraint in (1.33): we must choose our parame-
terization such that lines of constant σ are perpendicular to the lines
of constant τ, as shown in the figure.
But we can do better. To gain more physical insight, we need to make
use of the fact that we haven’t quite exhausted our gauge symmetry.
We will discuss this more in Section 2.2, but for now one can check that
there is enough remnant gauge symmetry to allow us to go to static
gauge,
X0
≡ t = Rτ ,
so that (X0
) = 0 and ˙X0
= R, where R is a constant that is needed on dimensional
grounds. The interpretation of this constant will become clear shortly. Then, writing
Xµ
= (t, x), the equation of motion for spatial components is the free wave equation,
¨x − x = 0
while the constraints become
˙x · x = 0
˙x 2
+ x 2
= R2
(1.34)
The first constraint tells us that the motion of the string must be perpendicular to the
string itself. In other words, the physical modes of the string are transverse oscillations.
There is no longitudinal mode. We’ll also see this again in Section 2.2.
From the second constraint, we can understand the meaning of the constant R: it is
related to the length of the string when ˙x = 0,
dσ (dx/dσ)2 = 2πR .
Of course, if we have a stretched string with ˙x = 0 at one moment of time, then it won’t
stay like that for long. It will contract under its own tension. As this happens, the
second constraint equation relates the length of the string to the instantaneous velocity
of the string.
– 24 –
1.4 Mode Expansions
Let’s look at the equations of motion and constraints more closely. The equations of
motion (1.31) are easily solved. We introduce lightcone coordinates on the worldsheet,
σ±
= τ ± σ ,
in terms of which the equations of motion simply read
∂+∂−Xµ
= 0
The most general solution is,
Xµ
(σ, τ) = Xµ
L(σ+
) + Xµ
R(σ−
)
for arbitrary functions Xµ
L and Xµ
R. These describe left-moving and right-moving waves
respectively. Of course the solution must still obey both the constraints (1.33) as well
as the periodicity condition,
Xµ
(σ, τ) = Xµ
(σ + 2π, τ) . (1.35)
The most general, periodic solution can be expanded in Fourier modes,
Xµ
L(σ+
) = 1
2
xµ
+ 1
2
α pµ
σ+
+ i
α
2 n=0
1
n
˜αµ
n e−inσ+
,
Xµ
R(σ−
) = 1
2
xµ
+ 1
2
α pµ
σ−
+ i
α
2 n=0
1
n
αµ
n e−inσ−
. (1.36)
This mode expansion will be very important when we come to the quantum theory.
Let’s make a few simple comments here.
• Various normalizations in this expression, such as the α and factor of 1/n have
been chosen for later convenience.
• XL and XR do not individually satisfy the periodicity condition (1.35) due to the
terms linear in σ±
. However, the sum of them is invariant under σ → σ + 2π as
required.
• The variables xµ
and pµ
are the position and momentum of the center of mass of
the string. This can be checked, for example, by studying the Noether currents
arising from the spacetime translation symmetry Xµ
→ Xµ
+ cµ
. One finds that
the conserved charge is indeed pµ
.
• Reality of Xµ
requires that the coefficients of the Fourier modes, αµ
n and ˜αµ
n, obey
αµ
n = (αµ
−n) , ˜αµ
n = (˜αµ
−n) . (1.37)
– 25 –
1.4.1 The Constraints Revisited
We still have to impose the two constraints (1.33). In the worldsheet lightcone coordi-
nates σ±
, these become,
(∂+X)2
= (∂−X)2
= 0 . (1.38)
These equations give constraints on the momenta pµ
and the Fourier modes αµ
n and ˜αµ
n.
To see what these are, let’s look at
∂−Xµ
= ∂−Xµ
R =
α
2
pµ
+
α
2 n=0
αµ
n e−inσ−
=
α
2 n
αµ
ne−inσ−
where in the second line the sum is over all n ∈ Z and we have defined αµ
0 to be
αµ
0 ≡
α
2
pµ
.
The constraint (1.38) can then be written as
(∂−X)2
=
α
2 m,p
αm · αp e−i(m+p)σ−
=
α
2 m,n
αm · αn−m e−inσ−
≡ α
n
Ln e−inσ−
= 0 .
where we have defined the sum of oscillator modes,
Ln =
1
2 m
αn−m · αm . (1.39)
We can also do the same for the left-moving modes, where we again define an analogous
sum of operator modes,
˜Ln =
1
2 m
˜αn−m · ˜αm . (1.40)
with the zero mode defined to be,
˜αµ
0 ≡
α
2
pµ
.
– 26 –
The fact that ˜αµ
0 = αµ
0 looks innocuous but is a key point to remember when we come
to quantize the string. The Ln and ˜Ln are the Fourier modes of the constraints. Any
classical solution of the string of the form (1.36) must further obey the infinite number
of constraints,
Ln = ˜Ln = 0 n ∈ Z .
We’ll meet these objects Ln and ˜Ln again in a more general context when we come to
discuss conformal field theory.
The constraints arising from L0 and ˜L0 have a rather special interpretation. This is
because they include the square of the spacetime momentum pµ
. But, the square of the
spacetime momentum is an important quantity in Minkowski space: it is the square of
the rest mass of a particle,
pµpµ
= −M2
.
So the L0 and ˜L0 constraints tell us the effective mass of a string in terms of the excited
oscillator modes, namely
M2
=
4
α n>0
αn · α−n =
4
α n>0
˜αn · ˜α−n (1.41)
Because both αµ
0 and ˜αµ
0 are equal to α /2 pµ
, we have two expressions for the invariant
mass: one in terms of right-moving oscillators αµ
n and one in terms of left-moving
oscillators ˜αµ
n. And these two terms must be equal to each other. This is known as
level matching. It will play an important role in the next section where we turn to the
quantum theory.
– 27 –
2. The Quantum String
Our goal in this section is to quantize the string. We have seen that the string action
involves a gauge symmetry and whenever we wish to quantize a gauge theory we’re
presented with a number of different ways in which we can proceed. If we’re working
in the canonical formalism, this usually boils down to one of two choices:
• We could first quantize the system and then subsequently impose the constraints
that arise from gauge fixing as operator equations on the physical states of the
system. For example, in QED this is the Gupta-Bleuler method of quantization
that we use in Lorentz gauge. In string theory it consists of treating all fields Xµ
,
including time X0
, as operators and imposing the constraint equations (1.33) on
the states. This is usually called covariant quantization.
• The alternative method is to first solve all of the constraints of the system to
determine the space of physically distinct classical solutions. We then quantize
these physical solutions. For example, in QED, this is the way we proceed in
Coulomb gauge. Later in this chapter, we will see a simple way to solve the
constraints of the free string.
Of course, if we do everything correctly, the two methods should agree. Usually, each
presents a slightly different challenge and offers a different viewpoint.
In these lectures, we’ll take a brief look at the first method of covariant quantization.
However, at the slightest sign of difficulties, we’ll bail! It will be useful enough to
see where the problems lie. We’ll then push forward with the second method described
above which is known as lightcone quantization in string theory. Although we’ll succeed
in pushing quantization through to the end, our derivations will be a little cheap and
unsatisfactory in places. In Section 5 we’ll return to all these issues, armed with more
sophisticated techniques from conformal field theory.
2.1 A Lightning Look at Covariant Quantization
We wish to quantize D free scalar fields Xµ
whose dynamics is governed by the action
(1.30). We subsequently wish to impose the constraints
˙X · X = ˙X2
+ X 2
= 0 . (2.1)
The first step is easy. We promote Xµ
and their conjugate momenta Πµ = (1/2πα ) ˙Xµ
to operator valued fields obeying the canonical equal-time commutation relations,
[Xµ
(σ, τ), Πν(σ , τ)] = iδ(σ − σ ) δµ
ν ,
[Xµ
(σ, τ), Xν
(σ , τ)] = [Πµ(σ, τ), Πν(σ , τ)] = 0 .
– 28 –
We translate these into commutation relations for the Fourier modes xµ
, pµ
, αµ
n and
˜αµ
n. Using the mode expansion (1.36) we find
[xµ
, pν] = iδµ
ν and [αµ
n, αν
m] = [˜αµ
n, ˜αν
m] = n ηµν
δn+m, 0 , (2.2)
with all others zero. The commutation relations for xµ
and pµ
are expected for oper-
ators governing the position and momentum of the center of mass of the string. The
commutation relations of αµ
n and ˜αµ
n are those of harmonic oscillator creation and anni-
hilation operators in disguise. And the disguise isn’t that good. We just need to define
(ignoring the µ index for now)
an =
αn
√
n
, a†
n =
α−n
√
n
n > 0 (2.3)
Then (2.2) gives the familiar [an, a†
m] = δmn. So each scalar field gives rise to two infinite
towers of creation and annihilation operators, with αn acting as a rescaled annihilation
operator for n > 0 and as a creation operator for n < 0. There are two towers because
we have right-moving modes αn and left-moving modes ˜αn.
With these commutation relations in hand we can now start building the Fock space
of our theory. We introduce a vacuum state of the string |0 , defined to obey
αµ
n |0 = ˜αµ
n |0 = 0 for n > 0 (2.4)
The vacuum state of string theory has a different interpretation from the analogous
object in field theory. This is not the vacuum state of spacetime. It is instead the
vacuum state of a single string. This is reflected in the fact that the operators xµ
and pµ
give extra structure to the vacuum. The true ground state of the string is |0 ,
tensored with a spatial wavefunction Ψ(x). Alternatively, if we work in momentum
space, the vacuum carries another quantum number, pµ
, which is the eigenvalue of the
momentum operator. We should therefore write the vacuum as |0; p , which still obeys
(2.4), but now also
ˆpµ
|0; p = pµ
|0; p (2.5)
where (for the only time in these lecture notes) we’ve put a hat on the momentum
operator ˆpµ
on the left-hand side of this equation to distinguish it from the eigenvalue
pµ
on the right-hand side.
We can now start to build up the Fock space by acting with creation operators αµ
n
and ˜αµ
n with n < 0. A generic state comes from acting with any number of these
creation operators on the vacuum,
(αµ1
−1)nµ1 (αµ2
−2)nµ2 . . . (˜αν1
−1)nν1 (˜αν2
−2)nν2 . . . |0; p
– 29 –
Each state in the Fock space is a different excited state of the string. Each has the
interpretation of a different species of particle in spacetime. We’ll see exactly what
particles they are shortly. But for now, notice that because there’s an infinite number
of ways to excite a string there are an infinite number of different species of particles
in this theory.
2.1.1 Ghosts
There’s a problem with the Fock space that we’ve constructed: it doesn’t have positive
norm. The reason for this is that one of the scalar fields, X0
, comes with the wrong sign
kinetic term in the action (1.30). From the perspective of the commutation relations,
this issue raises its head in presence of the spacetime Minkowski metric in the expression
[αµ
n, αν †
m ] = n ηµν
δn,m .
This gives rise to the offending negative norm states, which come with an odd number
of timelike oscillators excited, for example
p ; 0|α0
1α0
−1|0; p ∼ −δD
(p − p )
This is the first problem that arises in the covariant approach to quantization. States
with negative norm are referred to as ghosts. To make sense of the theory, we have
to make sure that they can’t be produced in any physical processes. Of course, this
problem is familiar from attempts to quantize QED in Lorentz gauge. In that case,
gauge symmetry rides to the rescue since the ghosts are removed by imposing the gauge
fixing constraint. We must hope that the same happens in string theory.
2.1.2 Constraints
Although we won’t push through with this programme at the present time, let us briefly
look at what kind of constraints we have in string theory. In terms of Fourier modes,
the classical constraints can be written as Ln = ˜Ln = 0, where
Ln =
1
2 m
αn−m · αm
and similar for ˜Ln. As in the Gupta-Bleuler quantization of QED, we don’t impose all
of these as operator equations on the Hilbert space. Instead we only require that the
operators Ln and ˜Ln have vanishing matrix elements when sandwiched between two
physical states |phys and |phys ,
phys |Ln|phys = phys |˜Ln|phys = 0
– 30 –
Because L†
n = L−n, it is therefore sufficient to require
Ln|phys = ˜Ln|phys = 0 for n > 0 (2.6)
However, we still haven’t explained how to impose the constraints L0 and ˜L0. And
these present a problem that doesn’t arise in the case of QED. The problem is that,
unlike for Ln with n = 0, the operator L0 is not uniquely defined when we pass to the
quantum theory. There is an operator ordering ambiguity arising from the commuta-
tion relations (2.2). Commuting the αµ
n operators past each other in L0 gives rise to
extra constant terms.
Question: How do we know what order to put the αµ
n operators in the quantum
operator L0? Or the ˜αµ
n operators in ˜L0?
Answer: We don’t! Yet. Naively it looks as if each different choice will define a
different theory when we impose the constraints. To make this ambiguity manifest, for
now let’s just pick a choice of ordering. We define the quantum operators to be normal
ordered, with the annihilation operators αi
n, n > 0, moved to the right,
L0 =
∞
m=1
α−m · αm +
1
2
α2
0 , ˜L0 =
∞
m=1
˜α−m · ˜αm +
1
2
˜α2
0
Then the ambiguity rears its head in the different constraint equations that we could
impose, namely
(L0 − a)|phys = (˜L0 − a)|phys = 0 (2.7)
for some constant a.
As we saw classically, the operators L0 and ˜L0 play an important role in determining
the spectrum of the string because they include a term quadratic in the momentum
αµ
0 = ˜αµ
0 = α /2 pµ
. Combining the expression (1.41) with our constraint equation
for L0 and ˜L0, we find the spectrum of the string is given by,
M2
=
4
α
−a +
∞
m=1
α−m · αm =
4
α
−a +
∞
m=1
˜α−m · ˜αm
We learn therefore that the undetermined constant a has a direct physical effect: it
changes the mass spectrum of the string. In the quantum theory, the sums over αµ
n
modes are related to the number operators for the harmonic oscillator: they count the
number of excited modes of the string. The level matching in the quantum theory
tells us that the number of left-moving modes must equal the number of right-moving
modes.
– 31 –
Ultimately, we will find that the need to decouple the ghosts forces us to make a
unique choice for the constant a. (Spoiler alert: it turns out to be a = 1). In fact, the
requirement that there are no ghosts is much stronger than this. It also restricts the
number of scalar fields that we have in the theory. (Another spoiler: D = 26). If you’re
interested in how this works in covariant formulation then you can read about it in the
book by Green, Schwarz and Witten. Instead, we’ll show how to quantize the string
and derive these values for a and D in lightcone gauge. However, after a trip through
the world of conformal field theory, we’ll come back to these ideas in a context which
is closer to the covariant approach.
2.2 Lightcone Quantization
We will now take the second path described at the beginning of this section. We will
try to find a parameterization of all classical solutions of the string. This is equivalent
to finding the classical phase space of the theory. We do this by solving the constraints
(2.1) in the classical theory, leaving behind only the physical degrees of freedom.
Recall that we fixed the gauge to set the worldsheet metric to
gαβ = ηαβ .
However, this isn’t the end of our gauge freedom. There still remain gauge transforma-
tions which preserve this choice of metric. In particular, any coordinate transformation
σ → ˜σ(σ) which changes the metric by
ηαβ → Ω2
(σ)ηαβ , (2.8)
can be undone by a Weyl transformation. What are these coordinate transformations?
It’s simplest to answer this using lightcone coordinates on the worldsheet,
σ±
= τ ± σ , (2.9)
where the flat metric on the worldsheet takes the form,
ds2
= −dσ+
dσ−
In these coordinates, it’s clear that any transformation of the form
σ+
→ ˜σ+
(σ+
) , σ−
→ ˜σ−
(σ−
) , (2.10)
simply multiplies the flat metric by an overall factor (2.8) and so can be undone by
a compensating Weyl transformation. Some quick comments on this surviving gauge
symmetry:
– 32 –
• Recall that in Section 1.3.2 we used the argument that 3 gauge invariances (2
reparameterizations + 1 Weyl) could be used to fix 3 components of the world-
sheet metric gαβ. What happened to this argument? Why do we still have some
gauge symmetry left? The reason is that ˜σ±
are functions of just a single variable,
not two. So we did fix nearly all our gauge symmetries. What is left is a set of
measure zero amongst the full gauge symmetry that we started with.
• The remaining reparameterization invariance (2.10) has an important physical
implication. Recall that the solutions to the equations of motion are of the form
Xµ
L(σ+
) + Xµ
R(σ−
) which looks like 2D functions worth of solutions. Of course,
we still have the constraints which, in terms of σ±
, read
(∂+X)2
= (∂−X)2
= 0 , (2.11)
which seems to bring the number down to 2(D − 1) functions. But the reparam-
eterization invariance (2.10) tells us that even some of these are fake since we
can always change what we mean by σ±
. The physical solutions of the string are
therefore actually described by 2(D − 2) functions. But this counting has a nice
interpretation: the degrees of freedom describe the transverse fluctuations of the
string.
• The above comment reaches the same conclusion as the discussion in Section
1.3.2. There, in an attempt to get some feel for the constraints, we claimed that
we could go to static gauge X0
= Rτ for some dimensionful parameter R. It
is easy to check that this is simple to do using reparameterizations of the form
(2.10). However, to solve the string constraints in full, it turns out that static
gauge is not that useful. Rather we will use something called “lightcone gauge”.
2.2.1 Lightcone Gauge
We would like to gauge fix the remaining reparameterization invariance (2.10). The best
way to do this is called lightcone gauge. In counterpoint to the worldsheet lightcone
coordinates (2.9), we introduce the spacetime lightcone coordinates,
X±
=
1
2
(X0
± XD−1
) . (2.12)
Note that this choice picks out a particular time direction and a particular spatial
direction. It means that any calculations that we do involving X±
will not be manifestly
Lorentz invariant. You might think that we needn’t really worry about this. We could
try to make the following argument: “The equations may not look Lorentz invariant
– 33 –
but, since we started from a Lorentz invariant theory, at the end of the day any physical
process is guaranteed to obey this symmetry”. Right?! Well, unfortunately not. One
of the more interesting and subtle aspects of quantum field theory is the possibility of
anomalies: these are symmetries of the classical theory that do not survive the journey
of quantization. When we come to the quantum theory, if our equations don’t look
Lorentz invariant then there’s a real possibility that it’s because the underlying physics
actually isn’t Lorentz invariant. Later we will need to spend some time figuring out
under what circumstances our quantum theory keeps the classical Lorentz symmetry.
In lightcone coordinates, the spacetime Minkowski metric reads
ds2
= −2dX+
dX−
+
D−2
i=1
dXi
dXi
This means that indices are raised and lowered with A+ = −A−
and A− = −A+
and
Ai = Ai
. The product of spacetime vectors reads A · B = −A+
B−
− A−
B+
+ Ai
Bi
.
Let’s look at the solution to the equation of motion for X+
. It reads,
X+
= X+
L (σ+
) + X+
R (σ−
) .
We now gauge fix. We use our freedom of reparameterization invariance to choose
coordinates such that
X+
L = 1
2
x+
+ 1
2
α p+
σ+
, X+
R = 1
2
x+
+ 1
2
α p+
σ−
.
You might think that we could go further and eliminate p+
X+
Figure 9:
and x+
but this isn’t possible because we don’t quite have
the full freedom of reparameterization invariance since all
functions should remain periodic in σ. The upshot of this
choice of gauge is that
X+
= x+
+ α p+
τ . (2.13)
This is lightcone gauge. Notice that, as long as p+
= 0, we
can always shift x+
by a shift in τ.
There’s something a little disconcerting about the choice
(2.13). We’ve identified a timelike worldsheet coordinate with a null spacetime coor-
dinate. Nonetheless, as you can see from the figure, it seems to be a good parameter-
ization of the worldsheet. One could imagine that the parameterization might break
if the string is actually massless and travels in the X−
direction, with p+
= 0. But
otherwise, all should be fine.
– 34 –
Solving for X−
The choice (2.13) does the job of fixing the reparameterization invariance (2.10). As
we will now see, it also renders the constraint equations trivial. The first thing that we
have to worry about is the possibility of extra constraints arising from this new choice
of gauge fixing. This can be checked by looking at the equation of motion for X+
,
∂+∂−X−
= 0
But we can solve this by the usual ansatz,
X−
= X−
L (σ+
) + X−
R (σ−
) .
We’re still left with all the other constraints (2.11). Here we see the real benefit of
working in lightcone gauge (which is actually what makes quantization possible at all):
X−
is completely determined by these constraints. For example, the first of these reads
2∂+X−
∂+X+
=
D−2
i=1
∂+Xi
∂+Xi
(2.14)
which, using (2.13), simply becomes
∂+X−
L =
1
α p+
D−2
i=1
∂+Xi
∂+Xi
. (2.15)
Similarly,
∂−X−
R =
1
α p+
D−2
i=1
∂−Xi
∂−Xi
. (2.16)
So, up to an integration constant, the function X−
(σ+
, σ−
) is completely determined
in terms of the other fields. If we write the usual mode expansion for X−
L/R
X−
L (σ+
) = 1
2
x−
+ 1
2
α p−
σ+
+ i
α
2 n=0
1
n
˜α−
n e−inσ+
,
X−
R (σ−
) = 1
2
x−
+ 1
2
α p−
σ−
+ i
α
2 n=0
1
n
α−
n e−inσ−
.
then x−
is the undetermined integration constant, while p−
, α−
n and ˜α−
n are all fixed by
the constraints (2.15) and (2.16). For example, the oscillator modes α−
n are given by,
α−
n =
1
2α
1
p+
+∞
m=−∞
D−2
i=1
αi
n−mαi
m , (2.17)
– 35 –
A special case of this is the α−
0 = α /2 p−
equation, which reads
α p−
2
=
1
2p+
D−2
i=1
1
2
α pi
pi
+
n=0
αi
nαi
−n . (2.18)
We also get another equation for p−
from the ˜α−
0 equation arising from (2.15)
α p−
2
=
1
2p+
D−2
i=1
1
2
α pi
pi
+
n=0
˜αi
n ˜αi
−n . (2.19)
From these two equations, we can reconstruct the old, classical, level matching condi-
tions (1.41). But now with a difference:
M2
= 2p+
p−
−
D−2
i=1
pi
pi
=
4
α
D−2
i=1 n>0
αi
−nαi
n =
4
α
D−2
i=1 n>0
˜αi
−n ˜αi
n . (2.20)
The difference is that now the sum is over oscillators αi
and ˜αi
only, with i = 1, . . . , D−
2. We’ll refer to these as transverse oscillators. Note that the string isn’t necessarily
living in the X0
-XD−1
plane, so these aren’t literally the transverse excitations of the
string. Nonetheless, if we specify the αi
then all other oscillator modes are determined.
In this sense, they are the physical excitation of the string.
Let’s summarize the state of play so far. The most general classical solution is
described in terms of 2(D − 2) transverse oscillator modes αi
n and ˜αi
n, together with
a number of zero modes describing the center of mass and momentum of the string:
xi
, pi
, p+
and x−
. But x+
can be absorbed by a shift of τ in (2.13) and p−
is constrained
to obey (2.18) and (2.19). In fact, p−
can be thought of as (proportional to) the
lightcone Hamiltonian. Indeed, we know that p−
generates translations in x+
, but this
is equivalent to shifts in τ.
2.2.2 Quantization
Having identified the physical degrees of freedom, let’s now quantize. We want to
impose commutation relations. Some of these are easy:
[xi
, pj
] = iδij
, [x−
, p+
] = −i
[αi
n, αj
m] = [˜αi
n, ˜αj
m] = nδij
δn+m,0 . (2.21)
– 36 –
all of which follow from the commutation relations (2.2) that we saw in covariant
quantization1
.
What to do with x+
and p−
? We could implement p−
as the Hamiltonian acting on
states. In fact, it will prove slightly more elegant (but equivalent) if we promote both
x+
and p−
to operators with the expected commutation relation,
[x+
, p−
] = −i . (2.22)
This is morally equivalent to writing [t, H] = −i in non-relativistic quantum mechanics,
which is true on a formal level. In the present context, it means that we can once again
choose states to be eigenstates of pµ
, with µ = 0, . . . , D, but the constraints (2.18) and
(2.19) must still be imposed as operator equations on the physical states. We’ll come
to this shortly.
The Hilbert space of states is very similar to that described in covariant quantization:
we define a vacuum state, |0; p such that
ˆpµ
|0; p = pµ
|0; p , αi
n|0; p = ˜αi
n|0; p = 0 for n > 0 (2.23)
and we build a Fock space by acting with the creation operators αi
−n and ˜αi
−n with
n > 0. The difference with the covariant quantization is that we only act with transverse
oscillators which carry a spatial index i = 1, . . . , D − 2. For this reason, the Hilbert
space is, by construction, positive definite. We don’t have to worry about ghosts.
1
Mea Culpa: We’re not really supposed to do this. The whole point of the approach that we’re
taking is to quantize just the physical degrees of freedom. The resulting commutation relations are
not, in general, inherited from the larger theory that we started with simply by closing our eyes
and forgetting about all the other fields that we’ve gauge fixed. We can see the problem by looking
at (2.17), where α−
n is determined in terms of αi
n. This means that the commutation relations for
αi
n might be infected by those of α−
n which could potentially give rise to extra terms. The correct
procedure to deal with this is to figure out the Poisson bracket structure of the physical degrees of
freedom in the classical theory. Or, in fancier language, the symplectic form on the phase space which
schematically looks like
ω ∼ dσ − d ˙X+
∧ dX−
− d ˙X−
∧ dX+
+ 2d ˙Xi
∧ dXi
,
The reason that the commutation relations (2.21) do not get infected is because the α−
terms in the
symplectic form come multiplying X+
. Yet X+
is given in (2.13). It has no oscillator modes. That
means that the symplectic form doesn’t pick up the Fourier modes of X−
and so doesn’t receive any
corrections from α−
n . The upshot of this is that the naive commutation relations (2.21) are actually
right.
– 37 –
The Constraints
Because p−
is not an independent variable in our theory, we must impose the constraints
(2.18) and (2.19) by hand as operator equations which define the physical states. In the
classical theory, we saw that these constraints are equivalent to mass-shell conditions
(2.20).
But there’s a problem when we go to the quantum theory. It’s the same problem
that we saw in covariant quantization: there’s an ordering ambiguity in the sum over
oscillator modes on the right-hand side of (2.20). If we choose all operators to be normal
ordered then this ambiguity reveals itself in an overall constant, a, which we have not
yet determined. The final result for the mass of states in lightcone gauge is:
M2
=
4
α
D−2
i=1 n>0
αi
−nαi
n − a =
4
α
D−2
i=1 n>0
˜αi
−n ˜αi
n − a
Since we’ll use this formula quite a lot in what follows, it’s useful to introduce quantities
related to the number operators of the harmonic oscillator,
N =
D−2
i=1 n>0
αi
−nαi
n , ˜N =
D−2
i=1 n>0
˜αi
−n ˜αi
n . (2.24)
These are not quite number operators because of the factor of 1/
√
n in (2.3). The value
of N and ˜N is often called the level. Which, if nothing else, means that the name “level
matching” makes sense. We now have
M2
=
4
α
(N − a) =
4
α
( ˜N − a) . (2.25)
How are we going to fix a? Later in the course we’ll see the correct way to do it. For
now, I’m just going to give you a quick and dirty derivation.
The Casimir Energy
“I told him that the sum of an infinite no. of terms of the series: 1 + 2 +
3 + 4 + . . . = − 1
12
under my theory. If I tell you this you will at once point
out to me the lunatic asylum as my goal. ”
Ramanujan, in a letter to G.H.Hardy.
What follows is a heuristic derivation of the normal ordering constant a. Suppose
that we didn’t notice that there was any ordering ambiguity and instead took the naive
classical result directly over to the quantum theory, that is
1
2 n=0
αi
−nαi
n =
1
2 n<0
αi
−nαi
n +
1
2 n>0
αi
−nαi
n .
– 38 –
where we’ve left the sum over i = 1, . . . , D − 2 implicit. We’ll now try to put this in
normal ordered form, with the annihilation operators αi
n with n > 0 on the right-hand
side. It’s the first term that needs changing. We get
1
2 n<0
αi
nαi
−n − n(D − 2) +
1
2 n>0
αi
−nαi
n =
n>0
αi
−nαi
n +
D − 2
2 n>0
n .
The final term clearly diverges. But it at least seems to have a physical interpretation:
it is the sum of zero point energies of an infinite number of harmonic oscillators. In
fact, we came across exactly the same type of term in the course on quantum field
theory where we learnt that, despite the divergence, one can still extract interesting
physics from this. This is the physics of the Casimir force.
Let’s recall the steps that we took to derive the Casimir force. Firstly, we introduced
an ultra-violet cut-off 1, probably muttering some words about no physical plates
being able to withstand very high energy quanta. Unfortunately, those words are no
longer available to us in string theory, but let’s proceed regardless. We replace the
divergent sum over integers by the expression,
∞
n=1
n −→
∞
n=1
ne− n
= −
∂
∂
∞
n=1
e− n
= −
∂
∂
(1 − e−
)−1
=
1
2
−
1
12
+ O( )
Obviously the 1/ 2
piece diverges as → 0. This term should be renormalized away.
In fact, this is necessary to preserve the Weyl invariance of the Polyakov action since it
contributes to a cosmological constant on the worldsheet. After this renormalization,
we’re left with the wonderful answer, first intuited by Ramanujan
∞
n=1
n = −
1
12
.
While heuristic, this argument does predict the correct physical Casimir energy mea-
sured in one-dimensional systems. For example, this effect is seen in simulations of
quantum spin chains.
What does this mean for our string? It means that we should take the unknown
constant a in the mass formula (2.25) to be,
M2
=
4
α
N −
D − 2
24
=
4
α
˜N −
D − 2
24
. (2.26)
This is the formula that we will use to determine the spectrum of the string.
– 39 –
Zeta Function Regularization
I appreciate that the preceding argument is not totally convincing. We could spend
some time making it more robust at this stage, but it’s best if we wait until later in the
course when we will have the tools of conformal field theory at our disposal. We will
eventually revisit this issue and provide a respectable derivation of the Casimir energy
in Section 4.4.1. For now I merely offer an even less convincing argument, known as
zeta-function regularization.
The zeta-function is defined, for Re(s) > 1, by the sum
ζ(s) =
∞
n=1
n−s
.
But ζ(s) has a unique analytic continuation to all values of s. In particular,
ζ(−1) = −
1
12
.
Good? Good. This argument is famously unconvincing the first time you meet it! But
it’s actually a very useful trick for getting the right answer.
2.3 The String Spectrum
Finally, we’re in a position to analyze the spectrum of a single, free string.
2.3.1 The Tachyon
Let’s start with the ground state |0; p defined in (2.23). With no oscillators excited,
the mass formula (2.26) gives
M2
= −
1
α
D − 2
6
. (2.27)
But that’s a little odd. It’s a negative mass-squared. Such particles are called tachyons.
In fact, tachyons aren’t quite as pathological as you might think. If you’ve heard
of these objects before, it’s probably in the context of special relativity where they’re
strange beasts which always travel faster than the speed of light. But that’s not the
right interpretation. Rather we should think more in the language of quantum field
theory. Suppose that we have a field in spacetime — let’s call it T(X) — whose quanta
will give rise to this particle. The mass-squared of the particle is simply the quadratic
term in the action, or
M2
=
∂2
V (T)
∂T2
T=0
– 40 –
So the negative mass-squared in (2.27) is telling us that we’re expanding around a
maximum of the potential for the tachyon field as shown in the figure. Note that from
this perspective, the Higgs field in the standard model at H = 0 is also a tachyon.
The fact that string theory turns out to sit at an unstable
T
V(T)
?
Figure 11:
point in the tachyon field is unfortunate. The natural ques-
tion is whether the potential has a good minimum elsewhere,
as shown in the figure to the right. No one knows the answer
to this! Naive attempts to understand this don’t work. We
know that around T = 0, the leading order contribution to
the potential is negative and quadratic. But there are fur-
ther terms that we can compute using techniques that we’ll
describe in Section 6. An expansion of the tachyon potential
around T = 0 looks like
V (T) =
1
2
M2
T2
+ c3T3
+ c4T4
+ . . .
It turns out that the T3
term in the potential does give rise to a minimum. But the T4
term destabilizes it again. Moreover, the T field starts to mix with other scalar fields
in the theory that we will come across soon. The ultimate fate of the tachyon in the
bosonic string is not yet understood.
The tachyon is a problem for the bosonic string. It may well be that this theory
makes no sense — or, at the very least, has no time-independent stable solutions. Or
perhaps we just haven’t worked out how to correctly deal with the tachyon. Either
way, the problem does not arise when we introduce fermions on the worldsheet and
study the superstring. This will involve several further technicalities which we won’t
get into in this course. Instead, our time will be put to better use if we continue to
study the bosonic string since all the lessons that we learn will carry over directly to the
superstring. However, one should be aware that the problem of the unstable vacuum
will continue to haunt us throughout this course.
Although we won’t describe it in detail, at several times along our journey we’ll make
an aside about how calculations work out for the superstring.
2.3.2 The First Excited States
We now look at the first excited states. If we act with a creation operator αj
−1, then the
level matching condition (2.25) tells us that we also need to act with a ˜αi
−1 operator.
This gives us (D − 2)2
particle states,
˜αi
−1αj
−1 |0; p , (2.28)
– 41 –
each of which has mass
M2
=
4
α
1 −
D − 2
24
.
But now we seem to have a problem. Our states have space indices i, j = 1, . . . , D − 2.
The operators αi
and ˜αi
each transform in the vector representation of SO(D − 2) ⊂
SO(1, D−1) which is manifest in lightcone gauge. But ultimately we want these states
to fit into some representation of the full Lorentz SO(1, D − 1) group. That looks as
if it’s going to be hard to arrange. This is the first manifestation of the comment that
we made after equation (2.12): it’s tricky to see Lorentz invariance in lightcone gauge.
To proceed, let’s recall Wigner’s classification of representations of the Poincar´e
group. We start by looking at massive particles in R1,D−1
. After going to the rest
frame of the particle by setting pµ
= (p, 0, . . . , 0), we can watch how any internal
indices transform under the little group SO(D − 1) of spatial rotations. The upshot
of this is that any massive particle must form a representation of SO(D − 1). But the
particles described by (2.28) have (D − 2)2
states. There’s no way to package these
states into a representation of SO(D − 1) and this means that there’s no way that the
first excited states of the string can form a massive representation of the D-dimensional
Poincar´e group.
It looks like we’re in trouble. Thankfully, there’s a way out. If the states are massless,
then we can’t go to the rest frame. The best that we can do is choose a spacetime
momentum for the particle of the form pµ
= (p, 0, . . . , 0, p). In this case, the particles
fill out a representation of the little group SO(D−2). This means that massless particles
get away with having fewer internal states than massive particles. For example, in four
dimensions the photon has two polarization states, but a massive spin-1 particle must
have three.
The first excited states (2.28) happily sit in a representation of SO(D −2). We learn
that if we want the quantum theory to preserve the SO(1, D − 1) Lorentz symmetry
that we started with, then these states will have to be massless. And this is only the
case if the dimension of spacetime is
D = 26 .
This is our first derivation of the critical dimension of the bosonic string.
Moreover, we’ve found that our theory contains a bunch of massless particles. And
massless particles are interesting because they give rise to long range forces. Let’s look
– 42 –
more closely at what massless particles the string has given us. The states (2.28) trans-
form in the 24 ⊗ 24 representation of SO(24). These decompose into three irreducible
representations:
traceless symmetric ⊕ anti-symmetric ⊕ singlet (=trace)
To each of these modes, we associate a massless field in spacetime such that the string
oscillation can be identified with a quantum of these fields. The fields are:
Gµν(X) , Bµν(X) , Φ(X) (2.29)
Of these, the first is the most interesting and we shall have more to say momentarily.
The second is an anti-symmetric tensor field which is usually called the anti-symmetric
tensor field. It also goes by the names of the “Kalb-Ramond field” or, in the language
of differential geometry, the “2-form”. The scalar field is called the dilaton. These three
massless fields are common to all string theories. We’ll learn more about the role these
fields play later in the course.
The particle in the symmetric traceless representation of SO(24) is particularly in-
teresting. This is a massless spin 2 particle. However, there are general arguments,
due originally to Feynman and Weinberg, that any theory of interacting massless spin
two particles must be equivalent to general relativity2
. We should therefore identify
the field Gµν(X) with the metric of spacetime. Let’s pause briefly to review the thrust
of these arguments.
Why Massless Spin 2 = General Relativity
Let’s call the spacetime metric Gµν(X). We can expand around flat space by writing
Gµν = ηµν + hµν(X) .
Then the Einstein-Hilbert action has an expansion in powers of h. If we truncate to
quadratic order, we simply have a free theory which we may merrily quantize in the
usual canonical fashion: we promote hµν to an operator and introduce the associated
creation and annihilation operators aµν and a†
µν. This way of looking at gravity is
anathema to those raised in the geometrical world of general relativity. But from a
particle physics language it is very standard: it is simply the quantization of a massless
spin 2 field, hµν.
2
A very readable description of this can be found in the first few chapters of the Feynman Lectures
on Gravitation.
– 43 –
However, even on this simple level, there is a problem due to the indefinite signature
of the spacetime Minkowski metric. The canonical quantization relations of the creation
and annihilation operators are schematically of the form,
[aµν, a†
ρσ] ∼ ηµρηνσ + ηµσηνρ
But this will lead to a Hilbert space with negative norm states coming from acting with
time-like creation operators. For example, the one-graviton state of the form,
a†
0i|0 (2.30)
suffers from a negative norm. This should be becoming familiar by now: it is the usual
problem that we run into if we try to covariantly quantize a gauge theory. And, indeed,
general relativity is a gauge theory. The gauge transformations are diffeomorphisms.
We would hope that this saves the theory of quantum gravity from these negative norm
states.
Let’s look a little more closely at what the gauge symmetry looks like for small fluctu-
ations hµν. We’ve butchered the Einstein-Hilbert action and left only terms quadratic
in h. Including all the index contractions, we find
SEH =
M2
pl
2
d4
x ∂µhρ
ρ∂νhµν
− ∂ρ
hµν
∂µhρν +
1
2
∂ρhµν∂ρ
hµν
−
1
2
∂µhν
ν∂µ
hρ
ρ + . . .
One can check that this truncated action is invariant under the gauge symmetry,
hµν −→ hµν + ∂µξν + ∂νξµ (2.31)
for any function ξµ(X). The gauge symmetry is the remnant of diffeomorphism invari-
ance, restricted to small deviations away from flat space. With this gauge invariance
in hand one can show that, just like QED, the negative norm states decouple from all
physical processes.
To summarize, theories of massless spin 2 fields only make sense if there is a gauge
symmetry to remove the negative norm states. In general relativity, this gauge symme-
try descends from diffeomorphism invariance. The argument of Feynman and Weinberg
now runs this logic in reverse. It goes as follows: suppose that we have a massless, spin
2 particle. Then, at the linearized level, it must be invariant under the gauge symmetry
(2.31) in order to eliminate the negative norm states. Moreover, this symmetry must
survive when interaction terms are introduced. But the only way to do this is to ensure
that the resulting theory obeys diffeomorpism invariance. That means the theory of
any interacting, massless spin 2 particle is Einstein gravity, perhaps supplemented by
higher derivative terms.
– 44 –
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String
String

More Related Content

Similar to String

Algebraic Topology - Hatcher.pdf
Algebraic Topology - Hatcher.pdfAlgebraic Topology - Hatcher.pdf
Algebraic Topology - Hatcher.pdfFaith Brown
 
Elastic wave propagation and generation in seismology by j. pujol (2003)
Elastic wave propagation and generation in seismology by j. pujol (2003)Elastic wave propagation and generation in seismology by j. pujol (2003)
Elastic wave propagation and generation in seismology by j. pujol (2003)Patricia Gutiérrez
 
A Short Introduction to Quantum Information and Quantum.pdf
A Short Introduction to Quantum Information and Quantum.pdfA Short Introduction to Quantum Information and Quantum.pdf
A Short Introduction to Quantum Information and Quantum.pdfSolMar38
 
Michael m. woolfson an introduction to x ray crystallography, second edition ...
Michael m. woolfson an introduction to x ray crystallography, second edition ...Michael m. woolfson an introduction to x ray crystallography, second edition ...
Michael m. woolfson an introduction to x ray crystallography, second edition ...PaReJaiiZz
 
PART VII.1 - Quantum Electrodynamics
PART VII.1 - Quantum ElectrodynamicsPART VII.1 - Quantum Electrodynamics
PART VII.1 - Quantum ElectrodynamicsMaurice R. TREMBLAY
 
A First Course In With Applications Complex Analysis
A First Course In With Applications Complex AnalysisA First Course In With Applications Complex Analysis
A First Course In With Applications Complex AnalysisElizabeth Williams
 
Cohen-Tannoudji, Diu and Laloë - Quantum Mechanics (vol. I, II and III, 2nd e...
Cohen-Tannoudji, Diu and Laloë - Quantum Mechanics (vol. I, II and III, 2nd e...Cohen-Tannoudji, Diu and Laloë - Quantum Mechanics (vol. I, II and III, 2nd e...
Cohen-Tannoudji, Diu and Laloë - Quantum Mechanics (vol. I, II and III, 2nd e...BEATRIZJAIMESGARCIA
 
Quantum Mechanics information imoortant.pptx
Quantum Mechanics information imoortant.pptxQuantum Mechanics information imoortant.pptx
Quantum Mechanics information imoortant.pptxXigfonSqitro
 
1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdfBEATRIZJAIMESGARCIA
 
1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdfBEATRIZJAIMESGARCIA
 
1000 Solved Problems In Modern Physics
1000 Solved Problems In Modern Physics1000 Solved Problems In Modern Physics
1000 Solved Problems In Modern PhysicsLaurie Smith
 
Elementary geometry from an advanced standpoint(Geometría Elemental Desde Un ...
Elementary geometry from an advanced standpoint(Geometría Elemental Desde Un ...Elementary geometry from an advanced standpoint(Geometría Elemental Desde Un ...
Elementary geometry from an advanced standpoint(Geometría Elemental Desde Un ...Minister Education
 
Short Review of the Unitary Quantum Theory
Short Review of the Unitary Quantum TheoryShort Review of the Unitary Quantum Theory
Short Review of the Unitary Quantum Theorytheijes
 
Practical Magnetotellurics Fiona Simpson.pdf
Practical Magnetotellurics Fiona Simpson.pdfPractical Magnetotellurics Fiona Simpson.pdf
Practical Magnetotellurics Fiona Simpson.pdfJorgeFloresCantu
 
Quantum cosmologyjj halliwell
Quantum cosmologyjj halliwellQuantum cosmologyjj halliwell
Quantum cosmologyjj halliwellLívia Rezende
 
Grimmett&Stirzaker--Probability and Random Processes Third Ed(2001).pdf
Grimmett&Stirzaker--Probability and Random Processes  Third Ed(2001).pdfGrimmett&Stirzaker--Probability and Random Processes  Third Ed(2001).pdf
Grimmett&Stirzaker--Probability and Random Processes Third Ed(2001).pdfAbdirahman Farah Ali
 

Similar to String (20)

Algebraic Topology - Hatcher.pdf
Algebraic Topology - Hatcher.pdfAlgebraic Topology - Hatcher.pdf
Algebraic Topology - Hatcher.pdf
 
Solid state
Solid stateSolid state
Solid state
 
Elastic wave propagation and generation in seismology by j. pujol (2003)
Elastic wave propagation and generation in seismology by j. pujol (2003)Elastic wave propagation and generation in seismology by j. pujol (2003)
Elastic wave propagation and generation in seismology by j. pujol (2003)
 
A Short Introduction to Quantum Information and Quantum.pdf
A Short Introduction to Quantum Information and Quantum.pdfA Short Introduction to Quantum Information and Quantum.pdf
A Short Introduction to Quantum Information and Quantum.pdf
 
Michael m. woolfson an introduction to x ray crystallography, second edition ...
Michael m. woolfson an introduction to x ray crystallography, second edition ...Michael m. woolfson an introduction to x ray crystallography, second edition ...
Michael m. woolfson an introduction to x ray crystallography, second edition ...
 
NBLovett
NBLovettNBLovett
NBLovett
 
PART VII.1 - Quantum Electrodynamics
PART VII.1 - Quantum ElectrodynamicsPART VII.1 - Quantum Electrodynamics
PART VII.1 - Quantum Electrodynamics
 
A First Course In With Applications Complex Analysis
A First Course In With Applications Complex AnalysisA First Course In With Applications Complex Analysis
A First Course In With Applications Complex Analysis
 
Cohen-Tannoudji, Diu and Laloë - Quantum Mechanics (vol. I, II and III, 2nd e...
Cohen-Tannoudji, Diu and Laloë - Quantum Mechanics (vol. I, II and III, 2nd e...Cohen-Tannoudji, Diu and Laloë - Quantum Mechanics (vol. I, II and III, 2nd e...
Cohen-Tannoudji, Diu and Laloë - Quantum Mechanics (vol. I, II and III, 2nd e...
 
Quantum Mechanics information imoortant.pptx
Quantum Mechanics information imoortant.pptxQuantum Mechanics information imoortant.pptx
Quantum Mechanics information imoortant.pptx
 
21 physics
21 physics21 physics
21 physics
 
1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf
 
1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf
 
1000 Solved Problems In Modern Physics
1000 Solved Problems In Modern Physics1000 Solved Problems In Modern Physics
1000 Solved Problems In Modern Physics
 
Elementary geometry from an advanced standpoint(Geometría Elemental Desde Un ...
Elementary geometry from an advanced standpoint(Geometría Elemental Desde Un ...Elementary geometry from an advanced standpoint(Geometría Elemental Desde Un ...
Elementary geometry from an advanced standpoint(Geometría Elemental Desde Un ...
 
diplom_master
diplom_masterdiplom_master
diplom_master
 
Short Review of the Unitary Quantum Theory
Short Review of the Unitary Quantum TheoryShort Review of the Unitary Quantum Theory
Short Review of the Unitary Quantum Theory
 
Practical Magnetotellurics Fiona Simpson.pdf
Practical Magnetotellurics Fiona Simpson.pdfPractical Magnetotellurics Fiona Simpson.pdf
Practical Magnetotellurics Fiona Simpson.pdf
 
Quantum cosmologyjj halliwell
Quantum cosmologyjj halliwellQuantum cosmologyjj halliwell
Quantum cosmologyjj halliwell
 
Grimmett&Stirzaker--Probability and Random Processes Third Ed(2001).pdf
Grimmett&Stirzaker--Probability and Random Processes  Third Ed(2001).pdfGrimmett&Stirzaker--Probability and Random Processes  Third Ed(2001).pdf
Grimmett&Stirzaker--Probability and Random Processes Third Ed(2001).pdf
 

Recently uploaded

module for grade 9 for distance learning
module for grade 9 for distance learningmodule for grade 9 for distance learning
module for grade 9 for distance learninglevieagacer
 
Module for Grade 9 for Asynchronous/Distance learning
Module for Grade 9 for Asynchronous/Distance learningModule for Grade 9 for Asynchronous/Distance learning
Module for Grade 9 for Asynchronous/Distance learninglevieagacer
 
Thyroid Physiology_Dr.E. Muralinath_ Associate Professor
Thyroid Physiology_Dr.E. Muralinath_ Associate ProfessorThyroid Physiology_Dr.E. Muralinath_ Associate Professor
Thyroid Physiology_Dr.E. Muralinath_ Associate Professormuralinath2
 
CYTOGENETIC MAP................ ppt.pptx
CYTOGENETIC MAP................ ppt.pptxCYTOGENETIC MAP................ ppt.pptx
CYTOGENETIC MAP................ ppt.pptxCherry
 
Selaginella: features, morphology ,anatomy and reproduction.
Selaginella: features, morphology ,anatomy and reproduction.Selaginella: features, morphology ,anatomy and reproduction.
Selaginella: features, morphology ,anatomy and reproduction.Cherry
 
Use of mutants in understanding seedling development.pptx
Use of mutants in understanding seedling development.pptxUse of mutants in understanding seedling development.pptx
Use of mutants in understanding seedling development.pptxRenuJangid3
 
Reboulia: features, anatomy, morphology etc.
Reboulia: features, anatomy, morphology etc.Reboulia: features, anatomy, morphology etc.
Reboulia: features, anatomy, morphology etc.Cherry
 
TransientOffsetin14CAftertheCarringtonEventRecordedbyPolarTreeRings
TransientOffsetin14CAftertheCarringtonEventRecordedbyPolarTreeRingsTransientOffsetin14CAftertheCarringtonEventRecordedbyPolarTreeRings
TransientOffsetin14CAftertheCarringtonEventRecordedbyPolarTreeRingsSérgio Sacani
 
GBSN - Microbiology (Unit 4) Concept of Asepsis
GBSN - Microbiology (Unit 4) Concept of AsepsisGBSN - Microbiology (Unit 4) Concept of Asepsis
GBSN - Microbiology (Unit 4) Concept of AsepsisAreesha Ahmad
 
LUNULARIA -features, morphology, anatomy ,reproduction etc.
LUNULARIA -features, morphology, anatomy ,reproduction etc.LUNULARIA -features, morphology, anatomy ,reproduction etc.
LUNULARIA -features, morphology, anatomy ,reproduction etc.Cherry
 
Genome Projects : Human, Rice,Wheat,E coli and Arabidopsis.
Genome Projects : Human, Rice,Wheat,E coli and Arabidopsis.Genome Projects : Human, Rice,Wheat,E coli and Arabidopsis.
Genome Projects : Human, Rice,Wheat,E coli and Arabidopsis.Cherry
 
COMPOSTING : types of compost, merits and demerits
COMPOSTING : types of compost, merits and demeritsCOMPOSTING : types of compost, merits and demerits
COMPOSTING : types of compost, merits and demeritsCherry
 
PODOCARPUS...........................pptx
PODOCARPUS...........................pptxPODOCARPUS...........................pptx
PODOCARPUS...........................pptxCherry
 
CURRENT SCENARIO OF POULTRY PRODUCTION IN INDIA
CURRENT SCENARIO OF POULTRY PRODUCTION IN INDIACURRENT SCENARIO OF POULTRY PRODUCTION IN INDIA
CURRENT SCENARIO OF POULTRY PRODUCTION IN INDIADr. TATHAGAT KHOBRAGADE
 
Pteris : features, anatomy, morphology and lifecycle
Pteris : features, anatomy, morphology and lifecyclePteris : features, anatomy, morphology and lifecycle
Pteris : features, anatomy, morphology and lifecycleCherry
 
FS P2 COMBO MSTA LAST PUSH past exam papers.
FS P2 COMBO MSTA LAST PUSH past exam papers.FS P2 COMBO MSTA LAST PUSH past exam papers.
FS P2 COMBO MSTA LAST PUSH past exam papers.takadzanijustinmaime
 
Role of AI in seed science Predictive modelling and Beyond.pptx
Role of AI in seed science  Predictive modelling and  Beyond.pptxRole of AI in seed science  Predictive modelling and  Beyond.pptx
Role of AI in seed science Predictive modelling and Beyond.pptxArvind Kumar
 
Site specific recombination and transposition.........pdf
Site specific recombination and transposition.........pdfSite specific recombination and transposition.........pdf
Site specific recombination and transposition.........pdfCherry
 
Human & Veterinary Respiratory Physilogy_DR.E.Muralinath_Associate Professor....
Human & Veterinary Respiratory Physilogy_DR.E.Muralinath_Associate Professor....Human & Veterinary Respiratory Physilogy_DR.E.Muralinath_Associate Professor....
Human & Veterinary Respiratory Physilogy_DR.E.Muralinath_Associate Professor....muralinath2
 
The Mariana Trench remarkable geological features on Earth.pptx
The Mariana Trench remarkable geological features on Earth.pptxThe Mariana Trench remarkable geological features on Earth.pptx
The Mariana Trench remarkable geological features on Earth.pptxseri bangash
 

Recently uploaded (20)

module for grade 9 for distance learning
module for grade 9 for distance learningmodule for grade 9 for distance learning
module for grade 9 for distance learning
 
Module for Grade 9 for Asynchronous/Distance learning
Module for Grade 9 for Asynchronous/Distance learningModule for Grade 9 for Asynchronous/Distance learning
Module for Grade 9 for Asynchronous/Distance learning
 
Thyroid Physiology_Dr.E. Muralinath_ Associate Professor
Thyroid Physiology_Dr.E. Muralinath_ Associate ProfessorThyroid Physiology_Dr.E. Muralinath_ Associate Professor
Thyroid Physiology_Dr.E. Muralinath_ Associate Professor
 
CYTOGENETIC MAP................ ppt.pptx
CYTOGENETIC MAP................ ppt.pptxCYTOGENETIC MAP................ ppt.pptx
CYTOGENETIC MAP................ ppt.pptx
 
Selaginella: features, morphology ,anatomy and reproduction.
Selaginella: features, morphology ,anatomy and reproduction.Selaginella: features, morphology ,anatomy and reproduction.
Selaginella: features, morphology ,anatomy and reproduction.
 
Use of mutants in understanding seedling development.pptx
Use of mutants in understanding seedling development.pptxUse of mutants in understanding seedling development.pptx
Use of mutants in understanding seedling development.pptx
 
Reboulia: features, anatomy, morphology etc.
Reboulia: features, anatomy, morphology etc.Reboulia: features, anatomy, morphology etc.
Reboulia: features, anatomy, morphology etc.
 
TransientOffsetin14CAftertheCarringtonEventRecordedbyPolarTreeRings
TransientOffsetin14CAftertheCarringtonEventRecordedbyPolarTreeRingsTransientOffsetin14CAftertheCarringtonEventRecordedbyPolarTreeRings
TransientOffsetin14CAftertheCarringtonEventRecordedbyPolarTreeRings
 
GBSN - Microbiology (Unit 4) Concept of Asepsis
GBSN - Microbiology (Unit 4) Concept of AsepsisGBSN - Microbiology (Unit 4) Concept of Asepsis
GBSN - Microbiology (Unit 4) Concept of Asepsis
 
LUNULARIA -features, morphology, anatomy ,reproduction etc.
LUNULARIA -features, morphology, anatomy ,reproduction etc.LUNULARIA -features, morphology, anatomy ,reproduction etc.
LUNULARIA -features, morphology, anatomy ,reproduction etc.
 
Genome Projects : Human, Rice,Wheat,E coli and Arabidopsis.
Genome Projects : Human, Rice,Wheat,E coli and Arabidopsis.Genome Projects : Human, Rice,Wheat,E coli and Arabidopsis.
Genome Projects : Human, Rice,Wheat,E coli and Arabidopsis.
 
COMPOSTING : types of compost, merits and demerits
COMPOSTING : types of compost, merits and demeritsCOMPOSTING : types of compost, merits and demerits
COMPOSTING : types of compost, merits and demerits
 
PODOCARPUS...........................pptx
PODOCARPUS...........................pptxPODOCARPUS...........................pptx
PODOCARPUS...........................pptx
 
CURRENT SCENARIO OF POULTRY PRODUCTION IN INDIA
CURRENT SCENARIO OF POULTRY PRODUCTION IN INDIACURRENT SCENARIO OF POULTRY PRODUCTION IN INDIA
CURRENT SCENARIO OF POULTRY PRODUCTION IN INDIA
 
Pteris : features, anatomy, morphology and lifecycle
Pteris : features, anatomy, morphology and lifecyclePteris : features, anatomy, morphology and lifecycle
Pteris : features, anatomy, morphology and lifecycle
 
FS P2 COMBO MSTA LAST PUSH past exam papers.
FS P2 COMBO MSTA LAST PUSH past exam papers.FS P2 COMBO MSTA LAST PUSH past exam papers.
FS P2 COMBO MSTA LAST PUSH past exam papers.
 
Role of AI in seed science Predictive modelling and Beyond.pptx
Role of AI in seed science  Predictive modelling and  Beyond.pptxRole of AI in seed science  Predictive modelling and  Beyond.pptx
Role of AI in seed science Predictive modelling and Beyond.pptx
 
Site specific recombination and transposition.........pdf
Site specific recombination and transposition.........pdfSite specific recombination and transposition.........pdf
Site specific recombination and transposition.........pdf
 
Human & Veterinary Respiratory Physilogy_DR.E.Muralinath_Associate Professor....
Human & Veterinary Respiratory Physilogy_DR.E.Muralinath_Associate Professor....Human & Veterinary Respiratory Physilogy_DR.E.Muralinath_Associate Professor....
Human & Veterinary Respiratory Physilogy_DR.E.Muralinath_Associate Professor....
 
The Mariana Trench remarkable geological features on Earth.pptx
The Mariana Trench remarkable geological features on Earth.pptxThe Mariana Trench remarkable geological features on Earth.pptx
The Mariana Trench remarkable geological features on Earth.pptx
 

String

  • 1. Preprint typeset in JHEP style - HYPER VERSION January 2009 String Theory University of Cambridge Part III Mathematical Tripos Dr David Tong Department of Applied Mathematics and Theoretical Physics, Centre for Mathematical Sciences, Wilberforce Road, Cambridge, CB3 OWA, UK http://www.damtp.cam.ac.uk/user/tong/string.html d.tong@damtp.cam.ac.uk – 1 –
  • 2. Recommended Books and Resources • J. Polchinski, String Theory This two volume work is the standard introduction to the subject. Our lectures will more or less follow the path laid down in volume one covering the bosonic string. The book contains explanations and descriptions of many details that have been deliberately (and, I suspect, at times inadvertently) swept under a very large rug in these lectures. Volume two covers the superstring. • M. Green, J. Schwarz and E. Witten, Superstring Theory Another two volume set. It is now over 20 years old and takes a slightly old-fashioned route through the subject, with no explicit mention of conformal field theory. How- ever, it does contain much good material and the explanations are uniformly excellent. Volume one is most relevant for these lectures. • B. Zwiebach, A First Course in String Theory This book grew out of a course given to undergraduates who had no previous exposure to general relativity or quantum field theory. It has wonderful pedagogical discussions of the basics of lightcone quantization. More surprisingly, it also has some very clear descriptions of several advanced topics, even though it misses out all the bits in between. • P. Di Francesco, P. Mathieu and D. S´en´echal, Conformal Field Theory This big yellow book is affectionately known as the yellow pages. It’s a great way to learn conformal field theory. At first glance, it comes across as slightly daunting because it’s big. (And yellow). But you soon realise that it’s big because it starts at the beginning and provides detailed explanations at every step. The material necessary for this course can be found in chapters 5 and 6. Further References: “String Theory and M-Theory” by Becker, Becker and Schwarz and “String Theory in a Nutshell” (it’s a big nutshell) by Kiritsis both deal with the bosonic string fairly quickly, but include more advanced topics that may be of interest. The book “D-Branes” by Johnson has lively and clear discussions about the many joys of D-branes. Links to several excellent online resources, including video lectures by Shiraz Minwalla, are listed on the course webpage.
  • 3. Contents 0. Introduction 1 0.1 Quantum Gravity 3 1. The Relativistic String 9 1.1 The Relativistic Point Particle 9 1.1.1 Quantization 11 1.1.2 Ein Einbein 13 1.2 The Nambu-Goto Action 14 1.2.1 Symmetries of the Nambu-Goto Action 17 1.2.2 Equations of Motion 18 1.3 The Polyakov Action 18 1.3.1 Symmetries of the Polyakov Action 20 1.3.2 Fixing a Gauge 22 1.4 Mode Expansions 25 1.4.1 The Constraints Revisited 26 2. The Quantum String 28 2.1 A Lightning Look at Covariant Quantization 28 2.1.1 Ghosts 30 2.1.2 Constraints 30 2.2 Lightcone Quantization 32 2.2.1 Lightcone Gauge 33 2.2.2 Quantization 36 2.3 The String Spectrum 40 2.3.1 The Tachyon 40 2.3.2 The First Excited States 41 2.3.3 Higher Excited States 45 2.4 Lorentz Invariance Revisited 46 2.5 A Nod to the Superstring 48 3. Open Strings and D-Branes 50 3.1 Quantization 53 3.1.1 The Ground State 54 3.1.2 First Excited States: A World of Light 55 – 1 –
  • 4. 3.1.3 Higher Excited States and Regge Trajectories 56 3.1.4 Another Nod to the Superstring 56 3.2 Brane Dynamics: The Dirac Action 57 3.3 Multiple Branes: A World of Glue 59 4. Introducing Conformal Field Theory 61 4.0.1 Euclidean Space 62 4.0.2 The Holomorphy of Conformal Transformations 63 4.1 Classical Aspects 63 4.1.1 The Stress-Energy Tensor 64 4.1.2 Noether Currents 66 4.1.3 An Example: The Free Scalar Field 67 4.2 Quantum Aspects 68 4.2.1 Operator Product Expansion 68 4.2.2 Ward Identities 70 4.2.3 Primary Operators 73 4.3 An Example: The Free Scalar Field 77 4.3.1 The Propagator 77 4.3.2 An Aside: No Goldstone Bosons in Two Dimensions 79 4.3.3 The Stress-Energy Tensor and Primary Operators 80 4.4 The Central Charge 82 4.4.1 c is for Casimir 84 4.4.2 The Weyl Anomaly 86 4.4.3 c is for Cardy 89 4.4.4 c has a Theorem 91 4.5 The Virasoro Algebra 92 4.5.1 Radial Quantization 92 4.5.2 The Virasoro Algebra 94 4.5.3 Representations of the Virasoro Algebra 96 4.5.4 Consequences of Unitarity 98 4.6 The State-Operator Map 99 4.6.1 Some Simple Consequences 101 4.6.2 Our Favourite Example: The Free Scalar Field 102 4.7 Brief Comments on Conformal Field Theories with Boundaries 105 5. The Polyakov Path Integral and Ghosts 108 5.1 The Path Integral 108 5.1.1 The Faddeev-Popov Method 109 – 2 –
  • 5. 5.1.2 The Faddeev-Popov Determinant 112 5.1.3 Ghosts 113 5.2 The Ghost CFT 114 5.3 The Critical “Dimension” of String Theory 117 5.3.1 The Usual Nod to the Superstring 118 5.3.2 An Aside: Non-Critical Strings 119 5.4 States and Vertex Operators 120 5.4.1 An Example: Closed Strings in Flat Space 122 5.4.2 An Example: Open Strings in Flat Space 123 5.4.3 More General CFTs 124 6. String Interactions 125 6.1 What to Compute? 125 6.1.1 Summing Over Topologies 127 6.2 Closed String Amplitudes at Tree Level 130 6.2.1 Remnant Gauge Symmetry: SL(2,C) 130 6.2.2 The Virasoro-Shapiro Amplitude 132 6.2.3 Lessons to Learn 135 6.3 Open String Scattering 139 6.3.1 The Veneziano Amplitude 141 6.3.2 The Tension of D-Branes 142 6.4 One-Loop Amplitudes 143 6.4.1 The Moduli Space of the Torus 143 6.4.2 The One-Loop Partition Function 146 6.4.3 Interpreting the String Partition Function 149 6.4.4 So is String Theory Finite? 152 6.4.5 Beyond Perturbation Theory? 153 6.5 Appendix: Games with Integrals and Gamma Functions 154 7. Low Energy Effective Actions 157 7.1 Einstein’s Equations 158 7.1.1 The Beta Function 159 7.1.2 Ricci Flow 163 7.2 Other Couplings 163 7.2.1 Charged Strings and the B field 163 7.2.2 The Dilaton 165 7.2.3 Beta Functions 167 7.3 The Low-Energy Effective Action 167 – 3 –
  • 6. 7.3.1 String Frame and Einstein Frame 168 7.3.2 Corrections to Einstein’s Equations 170 7.3.3 Nodding Once More to the Superstring 171 7.4 Some Simple Solutions 173 7.4.1 Compactifications 174 7.4.2 The String Itself 175 7.4.3 Magnetic Branes 177 7.4.4 Moving Away from the Critical Dimension 180 7.4.5 The Elephant in the Room: The Tachyon 183 7.5 D-Branes Revisited: Background Gauge Fields 183 7.5.1 The Beta Function 184 7.5.2 The Born-Infeld Action 187 7.6 The DBI Action 188 7.6.1 Coupling to Closed String Fields 189 7.7 The Yang-Mills Action 191 7.7.1 D-Branes in Type II Superstring Theories 195 8. Compactification and T-Duality 197 8.1 The View from Spacetime 197 8.1.1 Moving around the Circle 199 8.2 The View from the Worldsheet 200 8.2.1 Massless States 202 8.2.2 Charged Fields 202 8.2.3 Enhanced Gauge Symmetry 203 8.3 Why Big Circles are the Same as Small Circles 204 8.3.1 A Path Integral Derivation of T-Duality 206 8.3.2 T-Duality for Open Strings 207 8.3.3 T-Duality for Superstrings 208 8.3.4 Mirror Symmetry 208 8.4 Epilogue 209 – 4 –
  • 7. Acknowledgements These lectures are aimed at beginning graduate students. They assume a working knowledge of quantum field theory and general relativity. The lectures were given over one semester and are based broadly on Volume one of the book by Joe Polchinski. I inherited the course from Michael Green whose notes were extremely useful. I also benefited enormously from the insightful and entertaining video lectures by Shiraz Minwalla. I’m grateful to Anirban Basu, Niklas Beisert, Joe Bhaseen, Diego Correa, Nick Dorey, Michael Green, Anshuman Maharana, Malcolm Perry and Martin Schnabl for discus- sions and help with various aspects of these notes. I’m also grateful to the students, especially Carlos Guedes, for their excellent questions and superhuman typo-spotting abilities. Finally, my thanks to Alex Considine for infinite patience and understanding over the weeks these notes were written. I am supported by the Royal Society. – 5 –
  • 8. 0. Introduction String theory is an ambitious project. It purports to be an all-encompassing theory of the universe, unifying the forces of Nature, including gravity, in a single quantum mechanical framework. The premise of string theory is that, at the fundamental level, matter does not consist of point-particles but rather of tiny loops of string. From this slightly absurd beginning, the laws of physics emerge. General relativity, electromagnetism and Yang-Mills gauge theories all appear in a surprising fashion. However, they come with baggage. String theory gives rise to a host of other ingredients, most strikingly extra spatial dimensions of the universe beyond the three that we have observed. The purpose of this course is to understand these statements in detail. These lectures differ from most other courses that you will take in a physics degree. String theory is speculative science. There is no experimental evidence that string theory is the correct description of our world and scant hope that hard evidence will arise in the near future. Moreover, string theory is very much a work in progress and certain aspects of the theory are far from understood. Unresolved issues abound and it seems likely that the final formulation has yet to be written. For these reasons, I’ll begin this introduction by suggesting some answers to the question: Why study string theory? Reason 1. String theory is a theory of quantum gravity String theory unifies Einstein’s theory of general relativity with quantum mechanics. Moreover, it does so in a manner that retains the explicit connection with both quantum theory and the low-energy description of spacetime. But quantum gravity contains many puzzles, both technical and conceptual. What does spacetime look like at the shortest distance scales? How can we understand physics if the causal structure fluctuates quantum mechanically? Is the big bang truely the beginning of time? Do singularities that arise in black holes really signify the end of time? What is the microscopic origin of black hole entropy and what is it telling us? What is the resolution to the information paradox? Some of these issues will be reviewed later in this introduction. Whether or not string theory is the true description of reality, it offers a framework in which one can begin to explore these issues. For some questions, string theory has given very impressive and compelling answers. For others, string theory has been almost silent. – 1 –
  • 9. Reason 2. String theory may be the theory of quantum gravity With broad brush, string theory looks like an extremely good candidate to describe the real world. At low-energies it naturally gives rise to general relativity, gauge theories, scalar fields and chiral fermions. In other words, it contains all the ingredients that make up our universe. It also gives the only presently credible explanation for the value of the cosmological constant although, in fairness, I should add that the explanation is so distasteful to some that the community is rather amusingly split between whether this is a good thing or a bad thing. Moreover, string theory incorporates several ideas which do not yet have experimental evidence but which are considered to be likely candidates for physics beyond the standard model. Prime examples are supersymmetry and axions. However, while the broad brush picture looks good, the finer details have yet to be painted. String theory does not provide unique predictions for low-energy physics but instead offers a bewildering array of possibilities, mostly dependent on what is hidden in those extra dimensions. Partly, this problem is inherent to any theory of quantum gravity: as we’ll review shortly, it’s a long way down from the Planck scale to the domestic energy scales explored at the LHC. Using quantum gravity to extract predictions for particle physics is akin to using QCD to extract predictions for how coffee makers work. But the mere fact that it’s hard is little comfort if we’re looking for convincing evidence that string theory describes the world in which we live. While string theory cannot at present offer falsifiable predictions, it has nonetheless inspired new and imaginative proposals for solving outstanding problems in particle physics and cosmology. There are scenarios in which string theory might reveal itself in forthcoming experiments. Perhaps we’ll find extra dimensions at the LHC, perhaps we’ll see a network of fundamental strings stretched across the sky, or perhaps we’ll detect some feature of non-Gaussianity in the CMB that is characteristic of D-branes at work during inflation. My personal feeling however is that each of these is a long shot and we may not know whether string theory is right or wrong within our lifetimes. Of course, the history of physics is littered with naysayers, wrongly suggesting that various theories will never be testable. With luck, I’ll be one of them. Reason 3. String theory provides new perspectives on gauge theories String theory was born from attempts to understand the strong force. Almost forty years later, this remains one of the prime motivations for the subject. String theory provides tools with which to analyze down-to-earth aspects of quantum field theory that are far removed from high-falutin’ ideas about gravity and black holes. – 2 –
  • 10. Of immediate relevance to this course are the pedagogical reasons to invest time in string theory. At heart, it is the study of conformal field theory and gauge symmetry. The techniques that we’ll learn are not isolated to string theory, but apply to countless systems which have direct application to real world physics. On a deeper level, string theory provides new and very surprising methods to under- stand aspects of quantum gauge theories. Of these, the most startling is the AdS/CFT correspondence, first conjectured by Juan Maldacena, which gives a relationship be- tween strongly coupled quantum field theories and gravity in higher dimensions. These ideas have been applied in areas ranging from nuclear physics to condensed matter physics and have provided qualitative (and arguably quantitative) insights into strongly coupled phenomena. Reason 4. String theory provides new results in mathematics For the past 250 years, the close relationship between mathematics and physics has been almost a one-way street: physicists borrowed many things from mathematicians but, with a few noticeable exceptions, gave little back. In recent times, that has changed. Ideas and techniques from string theory and quantum field theory have been employed to give new “proofs” and, perhaps more importantly, suggest new directions and insights in mathematics. The most well known of these is mirror symmetry, a relationship between topologically different Calabi-Yau manifolds. The four reasons described above also crudely characterize the string theory commu- nity: there are “relativists” and “phenomenologists” and “field theorists” and “math- ematicians”. Of course, the lines between these different sub-disciplines are not fixed and one of the great attractions of string theory is its ability to bring together people working in different areas — from cosmology to condensed matter to pure mathematics — and provide a framework in which they can profitably communicate. In my opinion, it is this cross-fertilization between fields which is the greatest strength of string theory. 0.1 Quantum Gravity This is a starter course in string theory. Our focus will be on the perturbative approach to the bosonic string and, in particular, why this gives a consistent theory of quantum gravity. Before we leap into this, it is probably best to say a few words about quantum gravity itself. Like why it’s hard. And why it’s important. (And why it’s not). The Einstein Hilbert action is given by SEH = 1 16πGN d4 x √ −gR – 3 –
  • 11. Newton’s constant GN can be written as 8πGN = c M2 pl Throughout these lectures we work in units with = c = 1. The Planck mass Mpl defines an energy scale Mpl ≈ 2 × 1018 GeV . (This is sometimes referred to as the reduced Planck mass, to distinguish it from the scale without the factor of 8π, namely 1/GN ≈ 1 × 1019 GeV). There are a couple of simple lessons that we can already take from this. The first is that the relevant coupling in the quantum theory is 1/Mpl. To see that this is indeed the case from the perspective of the action, we consider small perturbations around flat Minkowski space, gµν = ηµν + 1 Mpl hµν The factor of 1/Mpl is there to ensure that when we expand out the Einstein-Hilbert action, the kinetic term for h is canonically normalized, meaning that it comes with no powers of Mpl. This then gives the kind of theory that you met in your first course on quantum field theory, albeit with an infinite series of interaction terms, SEH = d4 x (∂h)2 + 1 Mpl h (∂h)2 + 1 M2 pl h2 (∂h)2 + . . . Each of these terms is schematic: if you were to do this explicitly, you would find a mess of indices contracted in different ways. We see that the interactions are suppressed by powers of Mpl. This means that quantum perturbation theory is an expansion in the dimensionless ratio E2 /M2 pl, where E is the energy associated to the process of interest. We learn that gravity is weak, and therefore under control, at low-energies. But gravitational interactions become strong as the energy involved approaches the Planck scale. In the language of the renormalization group, couplings of this type are known as irrelevant. The second lesson to take away is that the Planck scale Mpl is very very large. The LHC will probe the electroweak scale, MEW ∼ 103 GeV. The ratio is MEW /Mpl ∼ 10−15 . For this reason, quantum gravity will not affect your daily life, even if your daily life involves the study of the most extreme observable conditions in the universe. – 4 –
  • 12. Gravity is Non-Renormalizable Quantum field theories with irrelevant couplings are typically ill-behaved at high- energies, rendering the theory ill-defined. Gravity is no exception. Theories of this type are called non-renormalizable, which means that the divergences that appear in the Feynman diagram expansion cannot be absorbed by a finite number of countert- erms. In pure Einstein gravity, the symmetries of the theory are enough to ensure that the one-loop S-matrix is finite. The first divergence occurs at two-loops and requires the introduction of a counterterm of the form, Γ ∼ 1 1 M4 pl d4 x √ −g Rµν ρσRρσ λκRλκ µν with = 4−D. All indications point towards the fact that this is the first in an infinite number of necessary counterterms. Coupling gravity to matter requires an interaction term of the form, Sint = d4 x 1 Mpl hµνTµν + O(h2 ) This makes the situation marginally worse, with the first diver- Figure 1: gence now appearing at one-loop. The Feynman diagram in the figure shows particle scattering through the exchange of two gravi- tons. When the momentum k running in the loop is large, the diagram is badly divergent: it scales as 1 M4 pl ∞ d4 k Non-renormalizable theories are commonplace in the history of physics, the most com- monly cited example being Fermi’s theory of the weak interaction. The first thing to say about them is that they are far from useless! Non-renormalizable theories are typically viewed as effective field theories, valid only up to some energy scale Λ. One deals with the divergences by simply admitting ignorance beyond this scale and treating Λ as a UV cut-off on any momentum integral. In this way, we get results which are valid to an accuracy of E/Λ (perhaps raised to some power). In the case of the weak interaction, Fermi’s theory accurately predicts physics up to an energy scale of 1/GF ∼ 100 GeV. In the case of quantum gravity, Einstein’s theory works to an accuracy of (E/Mpl)2 . – 5 –
  • 13. However, non-renormalizable theories are typically unable to describe physics at their cut-off scale Λ or beyond. This is because they are missing the true ultra-violet degrees of freedom which tame the high-energy behaviour. In the case of the weak force, these new degrees of freedom are the W and Z bosons. We would like to know what missing degrees of freedom are needed to complete gravity. Singularities Only a particle physicist would phrase all questions about the universe in terms of scattering amplitudes. In general relativity we typically think about the geometry as a whole, rather than bastardizing the Einstein-Hilbert action and discussing perturba- tions around flat space. In this language, the question of high-energy physics turns into one of short distance physics. Classical general relativity is not to be trusted in regions where the curvature of spacetime approaches the Planck scale and ultimately becomes singular. A quantum theory of gravity should resolve these singularities. The question of spacetime singularities is morally equivalent to that of high-energy scattering. Both probe the ultra-violet nature of gravity. A spacetime geometry is made of a coherent collection of gravitons, just as the electric and magnetic fields in a laser are made from a collection of photons. The short distance structure of spacetime is governed – after Fourier transform – by high momentum gravitons. Understanding spacetime singularities and high-energy scattering are different sides of the same coin. There are two situations in general relativity where singularity theorems tell us that the curvature of spacetime gets large: at the big bang and in the center of a black hole. These provide two of the biggest challenges to any putative theory of quantum gravity. Gravity is Subtle It is often said that general relativity contains the seeds of its own destruction. The theory is unable to predict physics at the Planck scale and freely admits to it. Problems such as non-renormalizability and singularities are, in a Rumsfeldian sense, known unknowns. However, the full story is more complicated and subtle. On the one hand, the issue of non-renormalizability may not quite be the crisis that it first appears. On the other hand, some aspects of quantum gravity suggest that general relativity isn’t as honest about its own failings as is usually advertised. The theory hosts a number of unknown unknowns, things that we didn’t even know that we didn’t know. We won’t have a whole lot to say about these issues in this course, but you should be aware of them. Here I mention only a few salient points. – 6 –
  • 14. Firstly, there is a key difference between Fermi’s theory of the weak interaction and gravity. Fermi’s theory was unable to provide predictions for any scattering process at energies above 1/GF . In contrast, if we scatter two objects at extremely high- energies in gravity — say, at energies E Mpl — then we know exactly what will happen: we form a big black hole. We don’t need quantum gravity to tell us this. Classical general relativity is sufficient. If we restrict attention to scattering, the crisis of non-renormalizability is not problematic at ultra-high energies. It’s troublesome only within a window of energies around the Planck scale. Similar caveats hold for singularities. If you are foolish enough to jump into a black hole, then you’re on your own: without a theory of quantum gravity, no one can tell you what fate lies in store at the singularity. Yet, if you are smart and stay outside of the black hole, you’ll be hard pushed to see any effects of quantum gravity. This is because Nature has conspired to hide Planck scale curvatures from our inquisitive eyes. In the case of black holes this is achieved through cosmic censorship which is a conjecture in classical general relativity that says singularities are hidden behind horizons. In the case of the big bang, it is achieved through inflation, washing away any traces from the very early universe. Nature appears to shield us from the effects of quantum gravity, whether in high-energy scattering or in singularities. I think it’s fair to say that no one knows if this conspiracy is pointing at something deep, or is merely inconvenient for scientists trying to probe the Planck scale. While horizons may protect us from the worst excesses of singularities, they come with problems of their own. These are the unknown unknowns: difficulties that arise when curvatures are small and general relativity says “trust me”. The entropy of black holes and the associated paradox of information loss strongly suggest that local quan- tum field theory breaks down at macroscopic distance scales. Attempts to formulate quantum gravity in de Sitter space, or in the presence of eternal inflation, hint at similar difficulties. Ideas of holography, black hole complimentarity and the AdS/CFT corre- spondence all point towards non-local effects and the emergence of spacetime. These are the deep puzzles of quantum gravity and their relationship to the ultra-violet properties of gravity is unclear. As a final thought, let me mention the one observation that has an outside chance of being related to quantum gravity: the cosmological constant. With an energy scale of Λ ∼ 10−3 eV it appears to have little to do with ultra-violet physics. If it does have its origins in a theory of quantum gravity, it must either be due to some subtle “unknown unknown”, or because it is explained away as an environmental quantity as in string theory. – 7 –
  • 15. Is the Time Ripe? Our current understanding of physics, embodied in the standard model, is valid up to energy scales of 103 GeV. This is 15 orders of magnitude away from the Planck scale. Why do we think the time is now ripe to tackle quantum gravity? Surely we are like the ancient Greeks arguing about atomism. Why on earth do we believe that we’ve developed the right tools to even address the question? The honest answer, I think, is hubris. Figure 2: However, there is mild circumstantial evidence that the framework of quantum field theory might hold all the way to the Planck scale without any- thing very dramatic happening in between. The main argument is unification. The three coupling constants of Nature run logarithmically, meeting miraculously at the GUT energy scale of 1015 GeV. Just slightly later, the fourth force of Nature, grav- ity, joins them. While not overwhelming, this does provide a hint that perhaps quantum field theory can be taken seriously at these ridiculous scales. Historically I suspect this was what convinced large parts of the community that it was ok to speak about processes at 1018 GeV. Finally, perhaps the most compelling argument for studying physics at the Planck scale is that string theory does provide a consistent unified quantum theory of gravity and the other forces. Given that we have this theory sitting in our laps, it would be foolish not to explore its consequences. The purpose of these lecture notes is to begin this journey. – 8 –
  • 16. 1. The Relativistic String All lecture courses on string theory start with a discussion of the point particle. Ours is no exception. We’ll take a flying tour through the physics of the relativistic point particle and extract a couple of important lessons that we’ll take with us as we move onto string theory. 1.1 The Relativistic Point Particle We want to write down the Lagrangian describing a relativistic particle of mass m. In anticipation of string theory, we’ll consider D-dimensional Minkowski space R1,D−1 . Throughout these notes, we work with signature ηµν = diag(−1, +1, +1, . . . , +1) Note that this is the opposite signature to my quantum field theory notes. If we fix a frame with coordinates Xµ = (t, x) the action is simple: S = −m dt 1 − ˙x · ˙x . (1.1) To see that this is correct we can compute the momentum p, conjugate to x, and the energy E which is equal to the Hamiltonian, p = m˙x 1 − ˙x · ˙x , E = m2 + p2 , both of which should be familiar from courses on special relativity. Although the Lagrangian (1.1) is correct, it’s not fully satisfactory. The reason is that time t and space x play very different roles in this Lagrangian. The position x is a dynamical degree of freedom. In contrast, time t is merely a parameter providing a label for the position. Yet Lorentz transformations are supposed to mix up t and x and such symmetries are not completely obvious in (1.1). Can we find a new Lagrangian in which time and space are on equal footing? One possibility is to treat both time and space as labels. This leads us to the concept of field theory. However, in this course we will be more interested in the other possibility: we will promote time to a dynamical degree of freedom. At first glance, this may appear odd: the number of degrees of freedom is one of the crudest ways we have to characterize a system. We shouldn’t be able to add more degrees of freedom – 9 –
  • 17. at will without fundamentally changing the system that we’re talking about. Another way of saying this is that the particle has the option to move in space, but it doesn’t have the option to move in time. It has to move in time. So we somehow need a way to promote time to a degree of freedom without it really being a true dynamical degree of freedom! How do we do this? The answer, as we will now show, is gauge symmetry. Consider the action, X X 0 Figure 3: S = −m dτ − ˙Xµ ˙Xνηµν , (1.2) where µ = 0, . . . , D − 1 and ˙Xµ = dXµ /dτ. We’ve introduced a new parameter τ which labels the position along the worldline of the particle as shown by the dashed lines in the figure. This action has a simple interpretation: it is just the proper time ds along the worldline. Naively it looks as if we now have D physical degrees of freedom rather than D − 1 because, as promised, the time direction X0 ≡ t is among our dynamical variables: X0 = X0 (τ). However, this is an illusion. To see why, we need to note that the action (1.2) has a very important property: reparameterization invariance. This means that we can pick a different parameter ˜τ on the worldline, related to τ by any monotonic function ˜τ = ˜τ(τ) . Let’s check that the action is invariant under transformations of this type. The inte- gration measure in the action changes as dτ = d˜τ |dτ/d˜τ|. Meanwhile, the velocities change as dXµ /dτ = (dXµ /d˜τ) (d˜τ/dτ). Putting this together, we see that the action can just as well be written in the ˜τ reparameterization, S = −m d˜τ − dXµ d˜τ dXν d˜τ ηµν . The upshot of this is that not all D degrees of freedom Xµ are physical. For example, suppose you find a solution to this system, so that you know how X0 changes with τ and how X1 changes with τ and so on. Not all of that information is meaningful because τ itself is not meaningful. In particular, we could use our reparameterization invariance to simply set τ = X0 (τ) ≡ t (1.3) – 10 –
  • 18. If we plug this choice into the action (1.2) then we recover our initial action (1.1). The reparameterization invariance is a gauge symmetry of the system. Like all gauge sym- metries, it’s not really a symmetry at all. Rather, it is a redundancy in our description. In the present case, it means that although we seem to have D degrees of freedom Xµ , one of them is fake. The fact that one of the degrees of freedom is a fake also shows up if we look at the momenta, pµ = ∂L ∂ ˙Xµ = m ˙Xν ηµν − ˙Xλ ˙Xρ ηλρ (1.4) These momenta aren’t all independent. They satisfy pµpµ + m2 = 0 (1.5) This is a constraint on the system. It is, of course, the mass-shell constraint for a relativistic particle of mass m. From the worldline perspective, it tells us that the particle isn’t allowed to sit still in Minkowski space: at the very least, it had better keep moving in a timelike direction with (p0 )2 ≥ m2 . One advantage of the action (1.2) is that the Poincar´e symmetry of the particle is now manifest, appearing as a global symmetry on the worldline Xµ → Λµ νXν + cµ (1.6) where Λ is a Lorentz transformation satisfying Λµ νηνρ Λσ ρ = ηµσ , while cµ corresponds to a constant translation. We have made all the symmetries manifest at the price of introducing a gauge symmetry into our system. A similar gauge symmetry will arise in the relativistic string and much of this course will be devoted to understanding its consequences. 1.1.1 Quantization It’s a trivial matter to quantize this action. We introduce a wavefunction Ψ(X). This satisfies the usual Schr¨odinger equation, i ∂Ψ ∂τ = HΨ . But, computing the Hamiltonian H = ˙Xµ pµ −L, we find that it vanishes: H = 0. This shouldn’t be surprising. It is simply telling us that the wavefunction doesn’t depend on – 11 –
  • 19. τ. Since the wavefunction is something physical while, as we have seen, τ is not, this is to be expected. Note that this doesn’t mean that time has dropped out of the problem. On the contrary, in this relativistic context, time X0 is an operator, just like the spatial coordinates x. This means that the wavefunction Ψ is immediately a function of space and time. It is not like a static state in quantum mechanics, but more akin to the fully integrated solution to the non-relativistic Schr¨odinger equation. The classical system has a constraint given by (1.5). In the quantum theory, we impose this constraint as an operator equation on the wavefunction, namely (pµ pµ + m2 )Ψ = 0. Using the usual representation of the momentum operator pµ = −i∂/∂Xµ , we recognize this constraint as the Klein-Gordon equation − ∂ ∂Xµ ∂ ∂Xν ηµν + m2 Ψ(X) = 0 (1.7) Although this equation is familiar from field theory, it’s important to realize that the interpretation is somewhat different. In relativistic field theory, the Klein-Gordon equa- tion is the equation of motion obeyed by a scalar field. In relativistic quantum mechan- ics, it is the equation obeyed by the wavefunction. In the early days of field theory, the fact that these two equations are the same led people to think one should view the wavefunction as a classical field and quantize it a second time. This isn’t cor- rect, but nonetheless the language has stuck and it is common to talk about the point particle perspective as “first quantization” and the field theory perspective as “second quantization”. So far we’ve considered only a free point particle. How can we Figure 4: introduce interactions into this framework? We would have to first decide which interactions are allowed: perhaps the particle can split into two; perhaps it can fuse with other particles? Obviously, there is a huge range of options for us to choose from. We would then assign amplitudes for these processes to happen. There would be certain restrictions coming from the requirement of unitarity which, among other things, would lead to the necessity of anti-particles. We could draw diagrams associated to the different interactions — an example is given in the figure — and in this manner we would slowly build up the Feynman diagram expansion that is familiar from field theory. In fact, this was pretty much the way Feynman himself approached the topic of QED. However, in practice we rarely construct particle interactions in this way because the field theory framework provides a much better way of looking at things. In contrast, this way of building up interactions is exactly what we will later do for strings. – 12 –
  • 20. 1.1.2 Ein Einbein There is another action that describes the relativistic point particle. We introduce yet another field on the worldline, e(τ), and write S = 1 2 dτ e−1 ˙X2 − em2 , (1.8) where we’ve used the notation ˙X2 = ˙Xµ ˙Xν ηµν. For the rest of these lectures, terms like X2 will always mean an implicit contraction with the spacetime Minkowski metric. This form of the action makes it look as if we have coupled the worldline theory to 1d gravity, with the field e(τ) acting as an einbein (in the sense of vierbeins that are introduced in general relativity). To see this, note that we could change notation and write this action in the more suggestive form S = − 1 2 dτ √ −gττ gττ ˙X2 + m2 . (1.9) where gττ = (gττ )−1 is the metric on the worldline and e = √ −gττ Although our action appears to have one more degree of freedom, e, it can be easily checked that it has the same equations of motion as (1.2). The reason for this is that e is completely fixed by its equation of motion, ˙X2 + e2 m2 = 0. Substituting this into the action (1.8) recovers (1.2) The action (1.8) has a couple of advantages over (1.2). Firstly, it works for massless particles with m = 0. Secondly, the absence of the annoying square root means that it’s easier to quantize in a path integral framework. The action (1.8) retains invariance under reparameterizations which are now written in a form that looks more like general relativity. For transformations parameterized by an infinitesimal η, we have τ → ˜τ = τ − η(τ) , δe = d dτ (η(τ)e) , δXµ = dXµ dτ η(τ) (1.10) The einbein e transforms as a density on the worldline, while each of the coordinates Xµ transforms as a worldline scalar. – 13 –
  • 21. 1.2 The Nambu-Goto Action A particle sweeps out a worldline in Minkowski space. A string σ τ Figure 5: sweeps out a worldsheet. We’ll parameterize this worldsheet by one timelike coordinate τ, and one spacelike coordinate σ. In this section we’ll focus on closed strings and take σ to be periodic, with range σ ∈ [0, 2π) . (1.11) We will sometimes package the two worldsheet coordinates to- gether as σα = (τ, σ), α = 0, 1. Then the string sweeps out a surface in spacetime which defines a map from the worldsheet to Minkowski space, Xµ (σ, τ) with µ = 0, . . . , D − 1. For closed strings, we require Xµ (σ, τ) = Xµ (σ + 2π, τ) . In this context, spacetime is sometimes referred to as the target space to distinguish it from the worldsheet. We need an action that describes the dynamics of this string. The key property that we will ask for is that nothing depends on the coordinates σα that we choose on the worldsheet. In other words, the string action should be reparameterization invariant. What kind of action does the trick? Well, for the point particle the action was proportional to the length of the worldline. The obvious generalization is that the action for the string should be proportional to the area, A, of the worldsheet. This is certainly a property that is characteristic of the worldsheet itself, rather than any choice of parameterization. How do we find the area A in terms of the coordinates Xµ (σ, τ)? The worldsheet is a curved surface embedded in spacetime. The induced metric, γαβ, on this surface is the pull-back of the flat metric on Minkowski space, γαβ = ∂Xµ ∂σα ∂Xν ∂σβ ηµν . (1.12) Then the action which is proportional to the area of the worldsheet is given by, S = −T d2 σ − det γ . (1.13) Here T is a constant of proportionality. We will see shortly that it is the tension of the string, meaning the mass per unit length. – 14 –
  • 22. We can write this action a little more explicitly. The pull-back of the metric is given by, γαβ = ˙X2 ˙X · X ˙X · X X 2 . where ˙Xµ = ∂Xµ /∂τ and Xµ = ∂Xµ /∂σ. The action then takes the form, S = −T d2 σ −( ˙X)2 (X )2 + ( ˙X · X )2 . (1.14) This is the Nambu-Goto action for a relativistic string. Action = Area: A Check If you’re unfamiliar with differential geometry, the argu- dl dl 1 2 τ σ Figure 6: ment about the pull-back of the metric may be a bit slick. Thankfully, there’s a more pedestrian way to see that the action (1.14) is equal to the area swept out by the world- sheet. It’s slightly simpler to make this argument for a sur- face embedded in Euclidean space rather than Minkowski space. We choose some parameterization of the sheet in terms of τ and σ, as drawn in the figure, and we write the coordinates of Euclidean space as X(σ, τ). We’ll compute the area of the infinitesimal shaded region. The vectors tangent to the boundary are, dl1 = ∂X ∂σ , dl2 = ∂X ∂τ . If the angle between these two vectors is θ, then the area is then given by ds2 = |dl1||dl2| sin θ = dl2 1 dl2 2(1 − cos2 θ) = dl2 1 dl2 2 − (dl1 · dl2)2 (1.15) which indeed takes the form of the integrand of (1.14). Tension and Dimension Let’s now see that T has the physical interpretation of tension. We write Minkowski coordinates as Xµ = (t, x). We work in a gauge with X0 ≡ t = Rτ, where R is a constant that is needed to balance up dimensions (see below) and will drop out at the end of the argument. Consider a snapshot of a string configuration at a time when – 15 –
  • 23. dx/dτ = 0 so that the instantaneous kinetic energy vanishes. Evaluating the action for a time dt gives S = −T dτdσR (dx/dσ)2 = −T dt (spatial length of string) . (1.16) But, when the kinetic energy vanishes, the action is proportional to the time integral of the potential energy, potential energy = T × (spatial length of string) . So T is indeed the energy per unit length as claimed. We learn that the string acts rather like an elastic band and its energy increases linearly with length. (This is different from the elastic bands you’re used to which obey Hooke’s law where energy increased quadratically with length). To minimize its potential energy, the string will want to shrink to zero size. We’ll see that when we include quantum effects this can’t happen because of the usual zero point energies. There is a slightly annoying way of writing the tension that has its origin in ancient history, but is commonly used today T = 1 2πα (1.17) where α is pronounced “alpha-prime”. In the language of our ancestors, α is referred to as the “universal Regge slope”. We’ll explain why later in this course. At this point, it’s worth pointing out some conventions that we have, until now, left implicit. The spacetime coordinates have dimension [X] = −1. In contrast, the worldsheet coordinates are taken to be dimensionless, [σ] = 0. (This can be seen in our identification σ ≡ σ + 2π). The tension is equal to the mass per unit length and has dimension [T] = 2. Obviously this means that [α ] = −2. We can therefore associate a length scale, ls, by α = l2 s (1.18) The string scale ls is the natural length that appears in string theory. In fact, in a certain sense (that we will make more precise later in the course) this length scale is the only parameter of the theory. – 16 –
  • 24. Actual Strings vs. Fundamental Strings There are several situations in Nature where string-like objects arise. Prime examples include magnetic flux tubes in superconductors and chromo-electric flux tubes in QCD. Cosmic strings, a popular speculation in cosmology, are similar objects, stretched across the sky. In each of these situations, there are typically two length scales associated to the string: the tension, T and the width of the string, L. For all these objects, the dynamics is governed by the Nambu-Goto action as long as the curvature of the string is much greater than L. (In the case of superconductors, one should work with a suitable non-relativistic version of the Nambu-Goto action). However, in each of these other cases, the Nambu-Goto action is not the end of the story. There will typically be additional terms in the action that depend on the width of the string. The form of these terms is not universal, but often includes a rigidity piece of form L K2 , where K is the extrinsic curvature of the worldsheet. Other terms could be added to describe fluctuations in the width of the string. The string scale, ls, or equivalently the tension, T, depends on the kind of string that we’re considering. For example, if we’re interested in QCD flux tubes then we would take T ∼ (1 GeV)2 (1.19) In this course we will consider fundamental strings which have zero width. What this means in practice is that we take the Nambu-Goto action as the complete description for all configurations of the string. These strings will have relevance to quantum gravity and the tension of the string is taken to be much larger, typically an order of magnitude or so below the Planck scale. T M2 pl = (1018 GeV)2 (1.20) However, I should point out that when we try to view string theory as a fundamental theory of quantum gravity, we don’t really know what value T should take. As we will see later in this course, it depends on many other aspects, most notably the string coupling and the volume of the extra dimensions. 1.2.1 Symmetries of the Nambu-Goto Action The Nambu-Goto action has two types of symmetry, each of a different nature. • Poincar´e invariance of the spacetime (1.6). This is a global symmetry from the perspective of the worldsheet, meaning that the parameters Λµ ν and cµ which label – 17 –
  • 25. the symmetry transformation are constants and do not depend on worldsheet coordinates σα . • Reparameterization invariance, σα → ˜σα (σ). As for the point particle, this is a gauge symmetry. It reflects the fact that we have a redundancy in our description because the worldsheet coordinates σα have no physical meaning. 1.2.2 Equations of Motion To derive the equations of motion for the Nambu-Goto string, we first introduce the momenta which we call Π because there will be countless other quantities that we want to call p later, Πτ µ = ∂L ∂ ˙Xµ = −T ( ˙X · X )Xµ − (X 2 ) ˙Xµ ( ˙X · X )2 − ˙X2 X 2 . Πσ µ = ∂L ∂X µ = −T ( ˙X · X ) ˙Xµ − ( ˙X2 )Xµ ( ˙X · X )2 − ˙X2 X 2 . The equations of motion are then given by, ∂Πτ µ ∂τ + ∂Πσ µ ∂σ = 0 These look like nasty, non-linear equations. In fact, there’s a slightly nicer way to write these equations, starting from the earlier action (1.13). Recall that the variation of a determinant is δ √ −γ = 1 2 √ −γ γαβ δγαβ. Using the definition of the pull-back metric γαβ, this gives rise to the equations of motion ∂α( − det γ γαβ ∂βXµ ) = 0 , (1.21) Although this notation makes the equations look a little nicer, we’re kidding ourselves. Written in terms of Xµ , they are still the same equations. Still nasty. 1.3 The Polyakov Action The square-root in the Nambu-Goto action means that it’s rather difficult to quantize using path integral techniques. However, there is another form of the string action which is classically equivalent to the Nambu-Goto action. It eliminates the square root at the expense of introducing another field, S = − 1 4πα d2 σ √ −ggαβ ∂αXµ ∂βXν ηµν (1.22) where g ≡ det g. This is the Polyakov action. (Polyakov didn’t discover the action, but he understood how to work with it in the path integral and for this reason it carries his name. The path integral treatment of this action will be the subject of Chapter 5). – 18 –
  • 26. The new field is gαβ. It is a dynamical metric on the worldsheet. From the perspective of the worldsheet, the Polyakov action is a bunch of scalar fields X coupled to 2d gravity. The equation of motion for Xµ is ∂α( √ −ggαβ ∂βXµ ) = 0 , (1.23) which coincides with the equation of motion (1.21) from the Nambu-Goto action, except that gαβ is now an independent variable which is fixed by its own equation of motion. To determine this, we vary the action (remembering again that δ √ −g = −1 2 √ −ggαβδgαβ = +1 2 √ −ggαβ δgαβ), δS = − T 2 d2 σ δgαβ √ −g ∂αXµ ∂βXν − 1 2 √ −g gαβgρσ ∂ρXµ ∂σXν ηµν = 0 .(1.24) The worldsheet metric is therefore given by, gαβ = 2f(σ) ∂αX · ∂βX , (1.25) where the function f(σ) is given by, f−1 = gρσ ∂ρX · ∂σX A comment on the potentially ambiguous notation: here, and below, any function f(σ) is always short-hand for f(σ, τ): it in no way implies that f depends only on the spatial worldsheet coordinate. We see that gαβ isn’t quite the same as the pull-back metric γαβ defined in equation (1.12); the two differ by the conformal factor f. However, this doesn’t matter because, rather remarkably, f drops out of the equation of motion (1.23). This is because the √ −g term scales as f, while the inverse metric gαβ scales as f−1 and the two pieces cancel. We therefore see that Nambu-Goto and the Polyakov actions result in the same equation of motion for X. In fact, we can see more directly that the Nambu-Goto and Polyakov actions coincide. We may replace gαβ in the Polyakov action (1.22) with its equation of motion gαβ = 2f γαβ. The factor of f also drops out of the action for the same reason that it dropped out of the equation of motion. In this manner, we recover the Nambu-Goto action (1.13). – 19 –
  • 27. 1.3.1 Symmetries of the Polyakov Action The fact that the presence of the factor f(σ, τ) in (1.25) didn’t actually affect the equations of motion for Xµ reflects the existence of an extra symmetry which the Polyakov action enjoys. Let’s look more closely at this. Firstly, the Polyakov action still has the two symmetries of the Nambu-Goto action, • Poincar´e invariance. This is a global symmetry on the worldsheet. Xµ → Λµ νXν + cµ . • Reparameterization invariance, also known as diffeomorphisms. This is a gauge symmetry on the worldsheet. We may redefine the worldsheet coordinates as σα → ˜σα (σ). The fields Xµ transform as worldsheet scalars, while gαβ transforms in the manner appropriate for a 2d metric. Xµ (σ) → ˜Xµ (˜σ) = Xµ (σ) gαβ(σ) → ˜gαβ(˜σ) = ∂σγ ∂˜σα ∂σδ ∂˜σβ gγδ(σ) It will sometimes be useful to work infinitesimally. If we make the coordinate change σα → ˜σα = σα − ηα (σ), for some small η. The transformations of the fields then become, δXµ (σ) = ηα ∂αXµ δgαβ(σ) = αηβ + βηα where the covariant derivative is defined by αηβ = ∂αηβ − Γσ αβησ with the Levi- Civita connection associated to the worldsheet metric given by the usual expres- sion, Γσ αβ = 1 2 gσρ (∂αgβρ + ∂βgρα − ∂ρgαβ) Together with these familiar symmetries, there is also a new symmetry which is novel to the Polyakov action. It is called Weyl invariance. • Weyl Invariance. Under this symmetry, Xµ (σ) → Xµ (σ), while the metric changes as gαβ(σ) → Ω2 (σ) gαβ(σ) . (1.26) Or, infinitesimally, we can write Ω2 (σ) = e2φ(σ) for small φ so that δgαβ(σ) = 2φ(σ) gαβ(σ) . – 20 –
  • 28. It is simple to see that the Polyakov action is invariant under this transformation: the factor of Ω2 drops out just as the factor of f did in equation (1.25), canceling between √ −g and the inverse metric gαβ . This is a gauge symmetry of the string, as seen by the fact that the parameter Ω depends on the worldsheet coordinates σ. This means that two metrics which are related by a Weyl transformation (1.26) are to be considered as the same physical state. Figure 7: An example of a Weyl transformation How should we think of Weyl invariance? It is not a coordinate change. Instead it is the invariance of the theory under a local change of scale which preserves the angles between all lines. For example the two worldsheet metrics shown in the figure are viewed by the Polyakov string as equivalent. This is rather surprising! And, as you might imagine, theories with this property are extremely rare. It should be clear from the discussion above that the property of Weyl invariance is special to two dimensions, for only there does the scaling factor coming from the determinant √ −g cancel that coming from the inverse metric. But even in two dimensions, if we wish to keep Weyl invariance then we are strictly limited in the kind of interactions that can be added to the action. For example, we would not be allowed a potential term for the worldsheet scalars of the form, d2 σ √ −g V (X) . These break Weyl invariance. Nor can we add a worldsheet cosmological constant term, µ d2 σ √ −g . This too breaks Weyl invariance. We will see later in this course that the requirement of Weyl invariance becomes even more stringent in the quantum theory. We will also see what kind of interactions terms can be added to the worldsheet. Indeed, much of this course can be thought of as the study of theories with Weyl invariance. – 21 –
  • 29. 1.3.2 Fixing a Gauge As we have seen, the equation of motion (1.23) looks pretty nasty. However, we can use the redundancy inherent in the gauge symmetry to choose coordinates in which they simplify. Let’s think about what we can do with the gauge symmetry. Firstly, we have two reparameterizations to play with. The worldsheet metric has three independent components. This means that we expect to be able to set any two of the metric components to a value of our choosing. We will choose to make the metric locally conformally flat, meaning gαβ = e2φ ηαβ , (1.27) where φ(σ, τ) is some function on the worldsheet. You can check that this is possible by writing down the change of the metric under a coordinate transformation and seeing that the differential equations which result from the condition (1.27) have solutions, at least locally. Choosing a metric of the form (1.27) is known as conformal gauge. We have only used reparameterization invariance to get to the metric (1.27). We still have Weyl transformations to play with. Clearly, we can use these to remove the last independent component of the metric and set φ = 0 such that, gαβ = ηαβ . (1.28) We end up with the flat metric on the worldsheet in Minkowski coordinates. A Diversion: How to make a metric flat The fact that we can use Weyl invariance to make any two-dimensional metric flat is an important result. Let’s take a quick diversion from our main discussion to see a different proof that isn’t tied to the choice of Minkowski coordinates on the worldsheet. We’ll work in 2d Euclidean space to avoid annoying minus signs. Consider two metrics related by a Weyl transformation, gαβ = e2φ gαβ. One can check that the Ricci scalars of the two metrics are related by, g R = √ g(R − 2 2 φ) . (1.29) We can therefore pick a φ such that the new metric has vanishing Ricci scalar, R = 0, simply by solving this differential equation for φ. However, in two dimensions (but not in higher dimensions) a vanishing Ricci scalar implies a flat metric. The reason is simply that there aren’t too many indices to play with. In particular, symmetry of the Riemann tensor in two dimensions means that it must take the form, Rαβγδ = R 2 (gαγgβδ − gαδgβγ) . – 22 –
  • 30. So R = 0 is enough to ensure that Rαβγδ = 0, which means that the manifold is flat. In equation (1.28), we’ve further used reparameterization invariance to pick coordinates in which the flat metric is the Minkowski metric. The equations of motion and the stress-energy tensor With the choice of the flat metric (1.28), the Polyakov action simplifies tremendously and becomes the theory of D free scalar fields. (In fact, this simplification happens in any conformal gauge). S = − 1 4πα d2 σ ∂αX · ∂α X , (1.30) and the equations of motion for Xµ reduce to the free wave equation, ∂α∂α Xµ = 0 . (1.31) Now that looks too good to be true! Are the horrible equations (1.23) really equivalent to a free wave equation? Well, not quite. There is something that we’ve forgotten: we picked a choice of gauge for the metric gαβ. But we must still make sure that the equation of motion for gαβ is satisfied. In fact, the variation of the action with respect to the metric gives rise to a rather special quantity: it is the stress-energy tensor, Tαβ. With a particular choice of normalization convention, we define the stress-energy tensor to be Tαβ = − 2 T 1 √ −g ∂S ∂gαβ . We varied the Polyakov action with respect to gαβ in (1.24). When we set gαβ = ηαβ we get Tαβ = ∂αX · ∂βX − 1 2 ηαβ ηρσ ∂ρX · ∂σX . (1.32) The equation of motion associated to the metric gαβ is simply Tαβ = 0. Or, more explicitly, T01 = ˙X · X = 0 T00 = T11 = 1 2 ( ˙X2 + X 2 ) = 0 . (1.33) We therefore learn that the equations of motion of the string are the free wave equations (1.31) subject to the two constraints (1.33) arising from the equation of motion Tαβ = 0. – 23 –
  • 31. Getting a feel for the constraints Let’s try to get some intuition for these constraints. There is a simple Figure 8: meaning of the first constraint in (1.33): we must choose our parame- terization such that lines of constant σ are perpendicular to the lines of constant τ, as shown in the figure. But we can do better. To gain more physical insight, we need to make use of the fact that we haven’t quite exhausted our gauge symmetry. We will discuss this more in Section 2.2, but for now one can check that there is enough remnant gauge symmetry to allow us to go to static gauge, X0 ≡ t = Rτ , so that (X0 ) = 0 and ˙X0 = R, where R is a constant that is needed on dimensional grounds. The interpretation of this constant will become clear shortly. Then, writing Xµ = (t, x), the equation of motion for spatial components is the free wave equation, ¨x − x = 0 while the constraints become ˙x · x = 0 ˙x 2 + x 2 = R2 (1.34) The first constraint tells us that the motion of the string must be perpendicular to the string itself. In other words, the physical modes of the string are transverse oscillations. There is no longitudinal mode. We’ll also see this again in Section 2.2. From the second constraint, we can understand the meaning of the constant R: it is related to the length of the string when ˙x = 0, dσ (dx/dσ)2 = 2πR . Of course, if we have a stretched string with ˙x = 0 at one moment of time, then it won’t stay like that for long. It will contract under its own tension. As this happens, the second constraint equation relates the length of the string to the instantaneous velocity of the string. – 24 –
  • 32. 1.4 Mode Expansions Let’s look at the equations of motion and constraints more closely. The equations of motion (1.31) are easily solved. We introduce lightcone coordinates on the worldsheet, σ± = τ ± σ , in terms of which the equations of motion simply read ∂+∂−Xµ = 0 The most general solution is, Xµ (σ, τ) = Xµ L(σ+ ) + Xµ R(σ− ) for arbitrary functions Xµ L and Xµ R. These describe left-moving and right-moving waves respectively. Of course the solution must still obey both the constraints (1.33) as well as the periodicity condition, Xµ (σ, τ) = Xµ (σ + 2π, τ) . (1.35) The most general, periodic solution can be expanded in Fourier modes, Xµ L(σ+ ) = 1 2 xµ + 1 2 α pµ σ+ + i α 2 n=0 1 n ˜αµ n e−inσ+ , Xµ R(σ− ) = 1 2 xµ + 1 2 α pµ σ− + i α 2 n=0 1 n αµ n e−inσ− . (1.36) This mode expansion will be very important when we come to the quantum theory. Let’s make a few simple comments here. • Various normalizations in this expression, such as the α and factor of 1/n have been chosen for later convenience. • XL and XR do not individually satisfy the periodicity condition (1.35) due to the terms linear in σ± . However, the sum of them is invariant under σ → σ + 2π as required. • The variables xµ and pµ are the position and momentum of the center of mass of the string. This can be checked, for example, by studying the Noether currents arising from the spacetime translation symmetry Xµ → Xµ + cµ . One finds that the conserved charge is indeed pµ . • Reality of Xµ requires that the coefficients of the Fourier modes, αµ n and ˜αµ n, obey αµ n = (αµ −n) , ˜αµ n = (˜αµ −n) . (1.37) – 25 –
  • 33. 1.4.1 The Constraints Revisited We still have to impose the two constraints (1.33). In the worldsheet lightcone coordi- nates σ± , these become, (∂+X)2 = (∂−X)2 = 0 . (1.38) These equations give constraints on the momenta pµ and the Fourier modes αµ n and ˜αµ n. To see what these are, let’s look at ∂−Xµ = ∂−Xµ R = α 2 pµ + α 2 n=0 αµ n e−inσ− = α 2 n αµ ne−inσ− where in the second line the sum is over all n ∈ Z and we have defined αµ 0 to be αµ 0 ≡ α 2 pµ . The constraint (1.38) can then be written as (∂−X)2 = α 2 m,p αm · αp e−i(m+p)σ− = α 2 m,n αm · αn−m e−inσ− ≡ α n Ln e−inσ− = 0 . where we have defined the sum of oscillator modes, Ln = 1 2 m αn−m · αm . (1.39) We can also do the same for the left-moving modes, where we again define an analogous sum of operator modes, ˜Ln = 1 2 m ˜αn−m · ˜αm . (1.40) with the zero mode defined to be, ˜αµ 0 ≡ α 2 pµ . – 26 –
  • 34. The fact that ˜αµ 0 = αµ 0 looks innocuous but is a key point to remember when we come to quantize the string. The Ln and ˜Ln are the Fourier modes of the constraints. Any classical solution of the string of the form (1.36) must further obey the infinite number of constraints, Ln = ˜Ln = 0 n ∈ Z . We’ll meet these objects Ln and ˜Ln again in a more general context when we come to discuss conformal field theory. The constraints arising from L0 and ˜L0 have a rather special interpretation. This is because they include the square of the spacetime momentum pµ . But, the square of the spacetime momentum is an important quantity in Minkowski space: it is the square of the rest mass of a particle, pµpµ = −M2 . So the L0 and ˜L0 constraints tell us the effective mass of a string in terms of the excited oscillator modes, namely M2 = 4 α n>0 αn · α−n = 4 α n>0 ˜αn · ˜α−n (1.41) Because both αµ 0 and ˜αµ 0 are equal to α /2 pµ , we have two expressions for the invariant mass: one in terms of right-moving oscillators αµ n and one in terms of left-moving oscillators ˜αµ n. And these two terms must be equal to each other. This is known as level matching. It will play an important role in the next section where we turn to the quantum theory. – 27 –
  • 35. 2. The Quantum String Our goal in this section is to quantize the string. We have seen that the string action involves a gauge symmetry and whenever we wish to quantize a gauge theory we’re presented with a number of different ways in which we can proceed. If we’re working in the canonical formalism, this usually boils down to one of two choices: • We could first quantize the system and then subsequently impose the constraints that arise from gauge fixing as operator equations on the physical states of the system. For example, in QED this is the Gupta-Bleuler method of quantization that we use in Lorentz gauge. In string theory it consists of treating all fields Xµ , including time X0 , as operators and imposing the constraint equations (1.33) on the states. This is usually called covariant quantization. • The alternative method is to first solve all of the constraints of the system to determine the space of physically distinct classical solutions. We then quantize these physical solutions. For example, in QED, this is the way we proceed in Coulomb gauge. Later in this chapter, we will see a simple way to solve the constraints of the free string. Of course, if we do everything correctly, the two methods should agree. Usually, each presents a slightly different challenge and offers a different viewpoint. In these lectures, we’ll take a brief look at the first method of covariant quantization. However, at the slightest sign of difficulties, we’ll bail! It will be useful enough to see where the problems lie. We’ll then push forward with the second method described above which is known as lightcone quantization in string theory. Although we’ll succeed in pushing quantization through to the end, our derivations will be a little cheap and unsatisfactory in places. In Section 5 we’ll return to all these issues, armed with more sophisticated techniques from conformal field theory. 2.1 A Lightning Look at Covariant Quantization We wish to quantize D free scalar fields Xµ whose dynamics is governed by the action (1.30). We subsequently wish to impose the constraints ˙X · X = ˙X2 + X 2 = 0 . (2.1) The first step is easy. We promote Xµ and their conjugate momenta Πµ = (1/2πα ) ˙Xµ to operator valued fields obeying the canonical equal-time commutation relations, [Xµ (σ, τ), Πν(σ , τ)] = iδ(σ − σ ) δµ ν , [Xµ (σ, τ), Xν (σ , τ)] = [Πµ(σ, τ), Πν(σ , τ)] = 0 . – 28 –
  • 36. We translate these into commutation relations for the Fourier modes xµ , pµ , αµ n and ˜αµ n. Using the mode expansion (1.36) we find [xµ , pν] = iδµ ν and [αµ n, αν m] = [˜αµ n, ˜αν m] = n ηµν δn+m, 0 , (2.2) with all others zero. The commutation relations for xµ and pµ are expected for oper- ators governing the position and momentum of the center of mass of the string. The commutation relations of αµ n and ˜αµ n are those of harmonic oscillator creation and anni- hilation operators in disguise. And the disguise isn’t that good. We just need to define (ignoring the µ index for now) an = αn √ n , a† n = α−n √ n n > 0 (2.3) Then (2.2) gives the familiar [an, a† m] = δmn. So each scalar field gives rise to two infinite towers of creation and annihilation operators, with αn acting as a rescaled annihilation operator for n > 0 and as a creation operator for n < 0. There are two towers because we have right-moving modes αn and left-moving modes ˜αn. With these commutation relations in hand we can now start building the Fock space of our theory. We introduce a vacuum state of the string |0 , defined to obey αµ n |0 = ˜αµ n |0 = 0 for n > 0 (2.4) The vacuum state of string theory has a different interpretation from the analogous object in field theory. This is not the vacuum state of spacetime. It is instead the vacuum state of a single string. This is reflected in the fact that the operators xµ and pµ give extra structure to the vacuum. The true ground state of the string is |0 , tensored with a spatial wavefunction Ψ(x). Alternatively, if we work in momentum space, the vacuum carries another quantum number, pµ , which is the eigenvalue of the momentum operator. We should therefore write the vacuum as |0; p , which still obeys (2.4), but now also ˆpµ |0; p = pµ |0; p (2.5) where (for the only time in these lecture notes) we’ve put a hat on the momentum operator ˆpµ on the left-hand side of this equation to distinguish it from the eigenvalue pµ on the right-hand side. We can now start to build up the Fock space by acting with creation operators αµ n and ˜αµ n with n < 0. A generic state comes from acting with any number of these creation operators on the vacuum, (αµ1 −1)nµ1 (αµ2 −2)nµ2 . . . (˜αν1 −1)nν1 (˜αν2 −2)nν2 . . . |0; p – 29 –
  • 37. Each state in the Fock space is a different excited state of the string. Each has the interpretation of a different species of particle in spacetime. We’ll see exactly what particles they are shortly. But for now, notice that because there’s an infinite number of ways to excite a string there are an infinite number of different species of particles in this theory. 2.1.1 Ghosts There’s a problem with the Fock space that we’ve constructed: it doesn’t have positive norm. The reason for this is that one of the scalar fields, X0 , comes with the wrong sign kinetic term in the action (1.30). From the perspective of the commutation relations, this issue raises its head in presence of the spacetime Minkowski metric in the expression [αµ n, αν † m ] = n ηµν δn,m . This gives rise to the offending negative norm states, which come with an odd number of timelike oscillators excited, for example p ; 0|α0 1α0 −1|0; p ∼ −δD (p − p ) This is the first problem that arises in the covariant approach to quantization. States with negative norm are referred to as ghosts. To make sense of the theory, we have to make sure that they can’t be produced in any physical processes. Of course, this problem is familiar from attempts to quantize QED in Lorentz gauge. In that case, gauge symmetry rides to the rescue since the ghosts are removed by imposing the gauge fixing constraint. We must hope that the same happens in string theory. 2.1.2 Constraints Although we won’t push through with this programme at the present time, let us briefly look at what kind of constraints we have in string theory. In terms of Fourier modes, the classical constraints can be written as Ln = ˜Ln = 0, where Ln = 1 2 m αn−m · αm and similar for ˜Ln. As in the Gupta-Bleuler quantization of QED, we don’t impose all of these as operator equations on the Hilbert space. Instead we only require that the operators Ln and ˜Ln have vanishing matrix elements when sandwiched between two physical states |phys and |phys , phys |Ln|phys = phys |˜Ln|phys = 0 – 30 –
  • 38. Because L† n = L−n, it is therefore sufficient to require Ln|phys = ˜Ln|phys = 0 for n > 0 (2.6) However, we still haven’t explained how to impose the constraints L0 and ˜L0. And these present a problem that doesn’t arise in the case of QED. The problem is that, unlike for Ln with n = 0, the operator L0 is not uniquely defined when we pass to the quantum theory. There is an operator ordering ambiguity arising from the commuta- tion relations (2.2). Commuting the αµ n operators past each other in L0 gives rise to extra constant terms. Question: How do we know what order to put the αµ n operators in the quantum operator L0? Or the ˜αµ n operators in ˜L0? Answer: We don’t! Yet. Naively it looks as if each different choice will define a different theory when we impose the constraints. To make this ambiguity manifest, for now let’s just pick a choice of ordering. We define the quantum operators to be normal ordered, with the annihilation operators αi n, n > 0, moved to the right, L0 = ∞ m=1 α−m · αm + 1 2 α2 0 , ˜L0 = ∞ m=1 ˜α−m · ˜αm + 1 2 ˜α2 0 Then the ambiguity rears its head in the different constraint equations that we could impose, namely (L0 − a)|phys = (˜L0 − a)|phys = 0 (2.7) for some constant a. As we saw classically, the operators L0 and ˜L0 play an important role in determining the spectrum of the string because they include a term quadratic in the momentum αµ 0 = ˜αµ 0 = α /2 pµ . Combining the expression (1.41) with our constraint equation for L0 and ˜L0, we find the spectrum of the string is given by, M2 = 4 α −a + ∞ m=1 α−m · αm = 4 α −a + ∞ m=1 ˜α−m · ˜αm We learn therefore that the undetermined constant a has a direct physical effect: it changes the mass spectrum of the string. In the quantum theory, the sums over αµ n modes are related to the number operators for the harmonic oscillator: they count the number of excited modes of the string. The level matching in the quantum theory tells us that the number of left-moving modes must equal the number of right-moving modes. – 31 –
  • 39. Ultimately, we will find that the need to decouple the ghosts forces us to make a unique choice for the constant a. (Spoiler alert: it turns out to be a = 1). In fact, the requirement that there are no ghosts is much stronger than this. It also restricts the number of scalar fields that we have in the theory. (Another spoiler: D = 26). If you’re interested in how this works in covariant formulation then you can read about it in the book by Green, Schwarz and Witten. Instead, we’ll show how to quantize the string and derive these values for a and D in lightcone gauge. However, after a trip through the world of conformal field theory, we’ll come back to these ideas in a context which is closer to the covariant approach. 2.2 Lightcone Quantization We will now take the second path described at the beginning of this section. We will try to find a parameterization of all classical solutions of the string. This is equivalent to finding the classical phase space of the theory. We do this by solving the constraints (2.1) in the classical theory, leaving behind only the physical degrees of freedom. Recall that we fixed the gauge to set the worldsheet metric to gαβ = ηαβ . However, this isn’t the end of our gauge freedom. There still remain gauge transforma- tions which preserve this choice of metric. In particular, any coordinate transformation σ → ˜σ(σ) which changes the metric by ηαβ → Ω2 (σ)ηαβ , (2.8) can be undone by a Weyl transformation. What are these coordinate transformations? It’s simplest to answer this using lightcone coordinates on the worldsheet, σ± = τ ± σ , (2.9) where the flat metric on the worldsheet takes the form, ds2 = −dσ+ dσ− In these coordinates, it’s clear that any transformation of the form σ+ → ˜σ+ (σ+ ) , σ− → ˜σ− (σ− ) , (2.10) simply multiplies the flat metric by an overall factor (2.8) and so can be undone by a compensating Weyl transformation. Some quick comments on this surviving gauge symmetry: – 32 –
  • 40. • Recall that in Section 1.3.2 we used the argument that 3 gauge invariances (2 reparameterizations + 1 Weyl) could be used to fix 3 components of the world- sheet metric gαβ. What happened to this argument? Why do we still have some gauge symmetry left? The reason is that ˜σ± are functions of just a single variable, not two. So we did fix nearly all our gauge symmetries. What is left is a set of measure zero amongst the full gauge symmetry that we started with. • The remaining reparameterization invariance (2.10) has an important physical implication. Recall that the solutions to the equations of motion are of the form Xµ L(σ+ ) + Xµ R(σ− ) which looks like 2D functions worth of solutions. Of course, we still have the constraints which, in terms of σ± , read (∂+X)2 = (∂−X)2 = 0 , (2.11) which seems to bring the number down to 2(D − 1) functions. But the reparam- eterization invariance (2.10) tells us that even some of these are fake since we can always change what we mean by σ± . The physical solutions of the string are therefore actually described by 2(D − 2) functions. But this counting has a nice interpretation: the degrees of freedom describe the transverse fluctuations of the string. • The above comment reaches the same conclusion as the discussion in Section 1.3.2. There, in an attempt to get some feel for the constraints, we claimed that we could go to static gauge X0 = Rτ for some dimensionful parameter R. It is easy to check that this is simple to do using reparameterizations of the form (2.10). However, to solve the string constraints in full, it turns out that static gauge is not that useful. Rather we will use something called “lightcone gauge”. 2.2.1 Lightcone Gauge We would like to gauge fix the remaining reparameterization invariance (2.10). The best way to do this is called lightcone gauge. In counterpoint to the worldsheet lightcone coordinates (2.9), we introduce the spacetime lightcone coordinates, X± = 1 2 (X0 ± XD−1 ) . (2.12) Note that this choice picks out a particular time direction and a particular spatial direction. It means that any calculations that we do involving X± will not be manifestly Lorentz invariant. You might think that we needn’t really worry about this. We could try to make the following argument: “The equations may not look Lorentz invariant – 33 –
  • 41. but, since we started from a Lorentz invariant theory, at the end of the day any physical process is guaranteed to obey this symmetry”. Right?! Well, unfortunately not. One of the more interesting and subtle aspects of quantum field theory is the possibility of anomalies: these are symmetries of the classical theory that do not survive the journey of quantization. When we come to the quantum theory, if our equations don’t look Lorentz invariant then there’s a real possibility that it’s because the underlying physics actually isn’t Lorentz invariant. Later we will need to spend some time figuring out under what circumstances our quantum theory keeps the classical Lorentz symmetry. In lightcone coordinates, the spacetime Minkowski metric reads ds2 = −2dX+ dX− + D−2 i=1 dXi dXi This means that indices are raised and lowered with A+ = −A− and A− = −A+ and Ai = Ai . The product of spacetime vectors reads A · B = −A+ B− − A− B+ + Ai Bi . Let’s look at the solution to the equation of motion for X+ . It reads, X+ = X+ L (σ+ ) + X+ R (σ− ) . We now gauge fix. We use our freedom of reparameterization invariance to choose coordinates such that X+ L = 1 2 x+ + 1 2 α p+ σ+ , X+ R = 1 2 x+ + 1 2 α p+ σ− . You might think that we could go further and eliminate p+ X+ Figure 9: and x+ but this isn’t possible because we don’t quite have the full freedom of reparameterization invariance since all functions should remain periodic in σ. The upshot of this choice of gauge is that X+ = x+ + α p+ τ . (2.13) This is lightcone gauge. Notice that, as long as p+ = 0, we can always shift x+ by a shift in τ. There’s something a little disconcerting about the choice (2.13). We’ve identified a timelike worldsheet coordinate with a null spacetime coor- dinate. Nonetheless, as you can see from the figure, it seems to be a good parameter- ization of the worldsheet. One could imagine that the parameterization might break if the string is actually massless and travels in the X− direction, with p+ = 0. But otherwise, all should be fine. – 34 –
  • 42. Solving for X− The choice (2.13) does the job of fixing the reparameterization invariance (2.10). As we will now see, it also renders the constraint equations trivial. The first thing that we have to worry about is the possibility of extra constraints arising from this new choice of gauge fixing. This can be checked by looking at the equation of motion for X+ , ∂+∂−X− = 0 But we can solve this by the usual ansatz, X− = X− L (σ+ ) + X− R (σ− ) . We’re still left with all the other constraints (2.11). Here we see the real benefit of working in lightcone gauge (which is actually what makes quantization possible at all): X− is completely determined by these constraints. For example, the first of these reads 2∂+X− ∂+X+ = D−2 i=1 ∂+Xi ∂+Xi (2.14) which, using (2.13), simply becomes ∂+X− L = 1 α p+ D−2 i=1 ∂+Xi ∂+Xi . (2.15) Similarly, ∂−X− R = 1 α p+ D−2 i=1 ∂−Xi ∂−Xi . (2.16) So, up to an integration constant, the function X− (σ+ , σ− ) is completely determined in terms of the other fields. If we write the usual mode expansion for X− L/R X− L (σ+ ) = 1 2 x− + 1 2 α p− σ+ + i α 2 n=0 1 n ˜α− n e−inσ+ , X− R (σ− ) = 1 2 x− + 1 2 α p− σ− + i α 2 n=0 1 n α− n e−inσ− . then x− is the undetermined integration constant, while p− , α− n and ˜α− n are all fixed by the constraints (2.15) and (2.16). For example, the oscillator modes α− n are given by, α− n = 1 2α 1 p+ +∞ m=−∞ D−2 i=1 αi n−mαi m , (2.17) – 35 –
  • 43. A special case of this is the α− 0 = α /2 p− equation, which reads α p− 2 = 1 2p+ D−2 i=1 1 2 α pi pi + n=0 αi nαi −n . (2.18) We also get another equation for p− from the ˜α− 0 equation arising from (2.15) α p− 2 = 1 2p+ D−2 i=1 1 2 α pi pi + n=0 ˜αi n ˜αi −n . (2.19) From these two equations, we can reconstruct the old, classical, level matching condi- tions (1.41). But now with a difference: M2 = 2p+ p− − D−2 i=1 pi pi = 4 α D−2 i=1 n>0 αi −nαi n = 4 α D−2 i=1 n>0 ˜αi −n ˜αi n . (2.20) The difference is that now the sum is over oscillators αi and ˜αi only, with i = 1, . . . , D− 2. We’ll refer to these as transverse oscillators. Note that the string isn’t necessarily living in the X0 -XD−1 plane, so these aren’t literally the transverse excitations of the string. Nonetheless, if we specify the αi then all other oscillator modes are determined. In this sense, they are the physical excitation of the string. Let’s summarize the state of play so far. The most general classical solution is described in terms of 2(D − 2) transverse oscillator modes αi n and ˜αi n, together with a number of zero modes describing the center of mass and momentum of the string: xi , pi , p+ and x− . But x+ can be absorbed by a shift of τ in (2.13) and p− is constrained to obey (2.18) and (2.19). In fact, p− can be thought of as (proportional to) the lightcone Hamiltonian. Indeed, we know that p− generates translations in x+ , but this is equivalent to shifts in τ. 2.2.2 Quantization Having identified the physical degrees of freedom, let’s now quantize. We want to impose commutation relations. Some of these are easy: [xi , pj ] = iδij , [x− , p+ ] = −i [αi n, αj m] = [˜αi n, ˜αj m] = nδij δn+m,0 . (2.21) – 36 –
  • 44. all of which follow from the commutation relations (2.2) that we saw in covariant quantization1 . What to do with x+ and p− ? We could implement p− as the Hamiltonian acting on states. In fact, it will prove slightly more elegant (but equivalent) if we promote both x+ and p− to operators with the expected commutation relation, [x+ , p− ] = −i . (2.22) This is morally equivalent to writing [t, H] = −i in non-relativistic quantum mechanics, which is true on a formal level. In the present context, it means that we can once again choose states to be eigenstates of pµ , with µ = 0, . . . , D, but the constraints (2.18) and (2.19) must still be imposed as operator equations on the physical states. We’ll come to this shortly. The Hilbert space of states is very similar to that described in covariant quantization: we define a vacuum state, |0; p such that ˆpµ |0; p = pµ |0; p , αi n|0; p = ˜αi n|0; p = 0 for n > 0 (2.23) and we build a Fock space by acting with the creation operators αi −n and ˜αi −n with n > 0. The difference with the covariant quantization is that we only act with transverse oscillators which carry a spatial index i = 1, . . . , D − 2. For this reason, the Hilbert space is, by construction, positive definite. We don’t have to worry about ghosts. 1 Mea Culpa: We’re not really supposed to do this. The whole point of the approach that we’re taking is to quantize just the physical degrees of freedom. The resulting commutation relations are not, in general, inherited from the larger theory that we started with simply by closing our eyes and forgetting about all the other fields that we’ve gauge fixed. We can see the problem by looking at (2.17), where α− n is determined in terms of αi n. This means that the commutation relations for αi n might be infected by those of α− n which could potentially give rise to extra terms. The correct procedure to deal with this is to figure out the Poisson bracket structure of the physical degrees of freedom in the classical theory. Or, in fancier language, the symplectic form on the phase space which schematically looks like ω ∼ dσ − d ˙X+ ∧ dX− − d ˙X− ∧ dX+ + 2d ˙Xi ∧ dXi , The reason that the commutation relations (2.21) do not get infected is because the α− terms in the symplectic form come multiplying X+ . Yet X+ is given in (2.13). It has no oscillator modes. That means that the symplectic form doesn’t pick up the Fourier modes of X− and so doesn’t receive any corrections from α− n . The upshot of this is that the naive commutation relations (2.21) are actually right. – 37 –
  • 45. The Constraints Because p− is not an independent variable in our theory, we must impose the constraints (2.18) and (2.19) by hand as operator equations which define the physical states. In the classical theory, we saw that these constraints are equivalent to mass-shell conditions (2.20). But there’s a problem when we go to the quantum theory. It’s the same problem that we saw in covariant quantization: there’s an ordering ambiguity in the sum over oscillator modes on the right-hand side of (2.20). If we choose all operators to be normal ordered then this ambiguity reveals itself in an overall constant, a, which we have not yet determined. The final result for the mass of states in lightcone gauge is: M2 = 4 α D−2 i=1 n>0 αi −nαi n − a = 4 α D−2 i=1 n>0 ˜αi −n ˜αi n − a Since we’ll use this formula quite a lot in what follows, it’s useful to introduce quantities related to the number operators of the harmonic oscillator, N = D−2 i=1 n>0 αi −nαi n , ˜N = D−2 i=1 n>0 ˜αi −n ˜αi n . (2.24) These are not quite number operators because of the factor of 1/ √ n in (2.3). The value of N and ˜N is often called the level. Which, if nothing else, means that the name “level matching” makes sense. We now have M2 = 4 α (N − a) = 4 α ( ˜N − a) . (2.25) How are we going to fix a? Later in the course we’ll see the correct way to do it. For now, I’m just going to give you a quick and dirty derivation. The Casimir Energy “I told him that the sum of an infinite no. of terms of the series: 1 + 2 + 3 + 4 + . . . = − 1 12 under my theory. If I tell you this you will at once point out to me the lunatic asylum as my goal. ” Ramanujan, in a letter to G.H.Hardy. What follows is a heuristic derivation of the normal ordering constant a. Suppose that we didn’t notice that there was any ordering ambiguity and instead took the naive classical result directly over to the quantum theory, that is 1 2 n=0 αi −nαi n = 1 2 n<0 αi −nαi n + 1 2 n>0 αi −nαi n . – 38 –
  • 46. where we’ve left the sum over i = 1, . . . , D − 2 implicit. We’ll now try to put this in normal ordered form, with the annihilation operators αi n with n > 0 on the right-hand side. It’s the first term that needs changing. We get 1 2 n<0 αi nαi −n − n(D − 2) + 1 2 n>0 αi −nαi n = n>0 αi −nαi n + D − 2 2 n>0 n . The final term clearly diverges. But it at least seems to have a physical interpretation: it is the sum of zero point energies of an infinite number of harmonic oscillators. In fact, we came across exactly the same type of term in the course on quantum field theory where we learnt that, despite the divergence, one can still extract interesting physics from this. This is the physics of the Casimir force. Let’s recall the steps that we took to derive the Casimir force. Firstly, we introduced an ultra-violet cut-off 1, probably muttering some words about no physical plates being able to withstand very high energy quanta. Unfortunately, those words are no longer available to us in string theory, but let’s proceed regardless. We replace the divergent sum over integers by the expression, ∞ n=1 n −→ ∞ n=1 ne− n = − ∂ ∂ ∞ n=1 e− n = − ∂ ∂ (1 − e− )−1 = 1 2 − 1 12 + O( ) Obviously the 1/ 2 piece diverges as → 0. This term should be renormalized away. In fact, this is necessary to preserve the Weyl invariance of the Polyakov action since it contributes to a cosmological constant on the worldsheet. After this renormalization, we’re left with the wonderful answer, first intuited by Ramanujan ∞ n=1 n = − 1 12 . While heuristic, this argument does predict the correct physical Casimir energy mea- sured in one-dimensional systems. For example, this effect is seen in simulations of quantum spin chains. What does this mean for our string? It means that we should take the unknown constant a in the mass formula (2.25) to be, M2 = 4 α N − D − 2 24 = 4 α ˜N − D − 2 24 . (2.26) This is the formula that we will use to determine the spectrum of the string. – 39 –
  • 47. Zeta Function Regularization I appreciate that the preceding argument is not totally convincing. We could spend some time making it more robust at this stage, but it’s best if we wait until later in the course when we will have the tools of conformal field theory at our disposal. We will eventually revisit this issue and provide a respectable derivation of the Casimir energy in Section 4.4.1. For now I merely offer an even less convincing argument, known as zeta-function regularization. The zeta-function is defined, for Re(s) > 1, by the sum ζ(s) = ∞ n=1 n−s . But ζ(s) has a unique analytic continuation to all values of s. In particular, ζ(−1) = − 1 12 . Good? Good. This argument is famously unconvincing the first time you meet it! But it’s actually a very useful trick for getting the right answer. 2.3 The String Spectrum Finally, we’re in a position to analyze the spectrum of a single, free string. 2.3.1 The Tachyon Let’s start with the ground state |0; p defined in (2.23). With no oscillators excited, the mass formula (2.26) gives M2 = − 1 α D − 2 6 . (2.27) But that’s a little odd. It’s a negative mass-squared. Such particles are called tachyons. In fact, tachyons aren’t quite as pathological as you might think. If you’ve heard of these objects before, it’s probably in the context of special relativity where they’re strange beasts which always travel faster than the speed of light. But that’s not the right interpretation. Rather we should think more in the language of quantum field theory. Suppose that we have a field in spacetime — let’s call it T(X) — whose quanta will give rise to this particle. The mass-squared of the particle is simply the quadratic term in the action, or M2 = ∂2 V (T) ∂T2 T=0 – 40 –
  • 48. So the negative mass-squared in (2.27) is telling us that we’re expanding around a maximum of the potential for the tachyon field as shown in the figure. Note that from this perspective, the Higgs field in the standard model at H = 0 is also a tachyon. The fact that string theory turns out to sit at an unstable T V(T) ? Figure 11: point in the tachyon field is unfortunate. The natural ques- tion is whether the potential has a good minimum elsewhere, as shown in the figure to the right. No one knows the answer to this! Naive attempts to understand this don’t work. We know that around T = 0, the leading order contribution to the potential is negative and quadratic. But there are fur- ther terms that we can compute using techniques that we’ll describe in Section 6. An expansion of the tachyon potential around T = 0 looks like V (T) = 1 2 M2 T2 + c3T3 + c4T4 + . . . It turns out that the T3 term in the potential does give rise to a minimum. But the T4 term destabilizes it again. Moreover, the T field starts to mix with other scalar fields in the theory that we will come across soon. The ultimate fate of the tachyon in the bosonic string is not yet understood. The tachyon is a problem for the bosonic string. It may well be that this theory makes no sense — or, at the very least, has no time-independent stable solutions. Or perhaps we just haven’t worked out how to correctly deal with the tachyon. Either way, the problem does not arise when we introduce fermions on the worldsheet and study the superstring. This will involve several further technicalities which we won’t get into in this course. Instead, our time will be put to better use if we continue to study the bosonic string since all the lessons that we learn will carry over directly to the superstring. However, one should be aware that the problem of the unstable vacuum will continue to haunt us throughout this course. Although we won’t describe it in detail, at several times along our journey we’ll make an aside about how calculations work out for the superstring. 2.3.2 The First Excited States We now look at the first excited states. If we act with a creation operator αj −1, then the level matching condition (2.25) tells us that we also need to act with a ˜αi −1 operator. This gives us (D − 2)2 particle states, ˜αi −1αj −1 |0; p , (2.28) – 41 –
  • 49. each of which has mass M2 = 4 α 1 − D − 2 24 . But now we seem to have a problem. Our states have space indices i, j = 1, . . . , D − 2. The operators αi and ˜αi each transform in the vector representation of SO(D − 2) ⊂ SO(1, D−1) which is manifest in lightcone gauge. But ultimately we want these states to fit into some representation of the full Lorentz SO(1, D − 1) group. That looks as if it’s going to be hard to arrange. This is the first manifestation of the comment that we made after equation (2.12): it’s tricky to see Lorentz invariance in lightcone gauge. To proceed, let’s recall Wigner’s classification of representations of the Poincar´e group. We start by looking at massive particles in R1,D−1 . After going to the rest frame of the particle by setting pµ = (p, 0, . . . , 0), we can watch how any internal indices transform under the little group SO(D − 1) of spatial rotations. The upshot of this is that any massive particle must form a representation of SO(D − 1). But the particles described by (2.28) have (D − 2)2 states. There’s no way to package these states into a representation of SO(D − 1) and this means that there’s no way that the first excited states of the string can form a massive representation of the D-dimensional Poincar´e group. It looks like we’re in trouble. Thankfully, there’s a way out. If the states are massless, then we can’t go to the rest frame. The best that we can do is choose a spacetime momentum for the particle of the form pµ = (p, 0, . . . , 0, p). In this case, the particles fill out a representation of the little group SO(D−2). This means that massless particles get away with having fewer internal states than massive particles. For example, in four dimensions the photon has two polarization states, but a massive spin-1 particle must have three. The first excited states (2.28) happily sit in a representation of SO(D −2). We learn that if we want the quantum theory to preserve the SO(1, D − 1) Lorentz symmetry that we started with, then these states will have to be massless. And this is only the case if the dimension of spacetime is D = 26 . This is our first derivation of the critical dimension of the bosonic string. Moreover, we’ve found that our theory contains a bunch of massless particles. And massless particles are interesting because they give rise to long range forces. Let’s look – 42 –
  • 50. more closely at what massless particles the string has given us. The states (2.28) trans- form in the 24 ⊗ 24 representation of SO(24). These decompose into three irreducible representations: traceless symmetric ⊕ anti-symmetric ⊕ singlet (=trace) To each of these modes, we associate a massless field in spacetime such that the string oscillation can be identified with a quantum of these fields. The fields are: Gµν(X) , Bµν(X) , Φ(X) (2.29) Of these, the first is the most interesting and we shall have more to say momentarily. The second is an anti-symmetric tensor field which is usually called the anti-symmetric tensor field. It also goes by the names of the “Kalb-Ramond field” or, in the language of differential geometry, the “2-form”. The scalar field is called the dilaton. These three massless fields are common to all string theories. We’ll learn more about the role these fields play later in the course. The particle in the symmetric traceless representation of SO(24) is particularly in- teresting. This is a massless spin 2 particle. However, there are general arguments, due originally to Feynman and Weinberg, that any theory of interacting massless spin two particles must be equivalent to general relativity2 . We should therefore identify the field Gµν(X) with the metric of spacetime. Let’s pause briefly to review the thrust of these arguments. Why Massless Spin 2 = General Relativity Let’s call the spacetime metric Gµν(X). We can expand around flat space by writing Gµν = ηµν + hµν(X) . Then the Einstein-Hilbert action has an expansion in powers of h. If we truncate to quadratic order, we simply have a free theory which we may merrily quantize in the usual canonical fashion: we promote hµν to an operator and introduce the associated creation and annihilation operators aµν and a† µν. This way of looking at gravity is anathema to those raised in the geometrical world of general relativity. But from a particle physics language it is very standard: it is simply the quantization of a massless spin 2 field, hµν. 2 A very readable description of this can be found in the first few chapters of the Feynman Lectures on Gravitation. – 43 –
  • 51. However, even on this simple level, there is a problem due to the indefinite signature of the spacetime Minkowski metric. The canonical quantization relations of the creation and annihilation operators are schematically of the form, [aµν, a† ρσ] ∼ ηµρηνσ + ηµσηνρ But this will lead to a Hilbert space with negative norm states coming from acting with time-like creation operators. For example, the one-graviton state of the form, a† 0i|0 (2.30) suffers from a negative norm. This should be becoming familiar by now: it is the usual problem that we run into if we try to covariantly quantize a gauge theory. And, indeed, general relativity is a gauge theory. The gauge transformations are diffeomorphisms. We would hope that this saves the theory of quantum gravity from these negative norm states. Let’s look a little more closely at what the gauge symmetry looks like for small fluctu- ations hµν. We’ve butchered the Einstein-Hilbert action and left only terms quadratic in h. Including all the index contractions, we find SEH = M2 pl 2 d4 x ∂µhρ ρ∂νhµν − ∂ρ hµν ∂µhρν + 1 2 ∂ρhµν∂ρ hµν − 1 2 ∂µhν ν∂µ hρ ρ + . . . One can check that this truncated action is invariant under the gauge symmetry, hµν −→ hµν + ∂µξν + ∂νξµ (2.31) for any function ξµ(X). The gauge symmetry is the remnant of diffeomorphism invari- ance, restricted to small deviations away from flat space. With this gauge invariance in hand one can show that, just like QED, the negative norm states decouple from all physical processes. To summarize, theories of massless spin 2 fields only make sense if there is a gauge symmetry to remove the negative norm states. In general relativity, this gauge symme- try descends from diffeomorphism invariance. The argument of Feynman and Weinberg now runs this logic in reverse. It goes as follows: suppose that we have a massless, spin 2 particle. Then, at the linearized level, it must be invariant under the gauge symmetry (2.31) in order to eliminate the negative norm states. Moreover, this symmetry must survive when interaction terms are introduced. But the only way to do this is to ensure that the resulting theory obeys diffeomorpism invariance. That means the theory of any interacting, massless spin 2 particle is Einstein gravity, perhaps supplemented by higher derivative terms. – 44 –