15.01.2013 Views

macroscopic and microscopic structure of black ... - Stefan Kowalczyk

macroscopic and microscopic structure of black ... - Stefan Kowalczyk

macroscopic and microscopic structure of black ... - Stefan Kowalczyk

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

AUGUST 2006<br />

MACROSCOPIC AND MICROSCOPIC<br />

STRUCTURE OF BLACK HOLES<br />

Master’s Thesis<br />

SZCZEPAN WOJCIECH KOWALCZYK 1<br />

SUPERVISOR: PROF. DR. ERIK P. VERLINDE<br />

1 email: stefan@kowalczyk.nl<br />

Keywords: Black Holes, Hawking Radiation,<br />

Supersymmetry, D-branes, Elliptic Genus


ABSTRACT<br />

This thesis is an introduction to classical, quantum, <strong>and</strong> stringy <strong>black</strong> holes with special<br />

focus on <strong>black</strong> hole entropy.<br />

We start with a brief review <strong>of</strong> the geometry <strong>of</strong> classical <strong>black</strong> holes in a variety <strong>of</strong><br />

different coordinate systems. The curious analogy between thermodynamics <strong>and</strong> the<br />

laws <strong>of</strong> <strong>black</strong> hole mechanics is demystified. We try to give an answer to the question<br />

whether general relativity can be interpreted as a statistical theory by studying a<br />

derivation <strong>of</strong> the Einstein equations from thermodynamical principles. Using quantum<br />

field theory in Rindler space we give an explicit derivation <strong>of</strong> the Hawking effect. In<br />

the main part <strong>of</strong> this thesis we try to underst<strong>and</strong> the <strong>microscopic</strong> origin <strong>of</strong> the entropy<br />

<strong>of</strong> certain extremal, BPS-saturated <strong>black</strong> holes by making use <strong>of</strong> D-brane technology.<br />

As a prelude to this famous entropy matching due to Strominger <strong>and</strong> Vafa, we examine<br />

the so-called correspondence principle between strings <strong>and</strong> <strong>black</strong> holes. We also explain<br />

why extremal <strong>black</strong> holes are sometimes called supersymmetric solitons. In order<br />

to illustrate some st<strong>and</strong>ard counting techniques, we count 1/2-BPS heterotic states on<br />

T 4 × T 2 . Finally, we focus on modular forms, elliptic genera <strong>and</strong> other mathematical<br />

tools, which we use to review the counting <strong>of</strong> dyons in N = 4 string theory.


ACKNOWLEDGEMENTS<br />

I would like to thank my advisor, Pr<strong>of</strong>. dr. Erik Verlinde, for motivating me to dive<br />

into the physics <strong>of</strong> <strong>black</strong> holes <strong>and</strong> explaining many things to me ever since we met at<br />

the student initiated seminar.<br />

I am also very grateful to Pr<strong>of</strong>. dr. Jan de Boer, Pr<strong>of</strong>. dr. Klaas L<strong>and</strong>sman, <strong>and</strong><br />

Pr<strong>of</strong>. dr. Erik Verlinde for writing recommendation letters for me, which helped me<br />

getting into the State University <strong>of</strong> New York at Stony Brook.<br />

Last but not least, I would like to thank my girlfriend, Rebecca Hegeman, <strong>and</strong> my<br />

family for always being there for me.


CONTENTS<br />

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

2. Black Holes in Einstein Gravity . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2.1 Einstein gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2.2 The Schwarzschild <strong>black</strong> hole . . . . . . . . . . . . . . . . . . . . . 7<br />

2.3 Charged <strong>black</strong> holes . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

2.3.1 The extreme Reissner-Nordström solution . . . . . . . . . . . 12<br />

3. Motivation for Black Hole Thermodynamics . . . . . . . . . . . . . . . . 14<br />

3.1 Two derivations <strong>of</strong> the first <strong>black</strong> hole law . . . . . . . . . . . . . . . 15<br />

3.2 Black hole entropy <strong>and</strong> the generalized second law . . . . . . . . . . 18<br />

3.3 The holographic principle . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

3.4 Derivation <strong>of</strong> Einstein equations from thermodynamics . . . . . . . . 20<br />

3.5 Wald entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

4. Quantum Black Holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

4.1 Rindler space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

4.2 Field theory in Rindler space . . . . . . . . . . . . . . . . . . . . . . 28<br />

4.2.1 Derivation <strong>of</strong> the Hawking effect . . . . . . . . . . . . . . . . 30<br />

4.3 Path integral approach . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

5. Extremal Black Holes in String Theory . . . . . . . . . . . . . . . . . . . 38<br />

5.1 Perturbative microstates in string theory . . . . . . . . . . . . . . . . 38<br />

5.2 Correspondence principle between strings <strong>and</strong> <strong>black</strong> holes . . . . . . 40<br />

5.3 Extremal <strong>black</strong> holes <strong>and</strong> extended supersymmetry . . . . . . . . . . 41<br />

5.4 Type II supergravity . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

5.5 Form fields <strong>and</strong> D-branes . . . . . . . . . . . . . . . . . . . . . . . . 49<br />

5.6 Dualities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

5.7 Counting <strong>of</strong> 1/2 BPS heterotic states on T 4 × T 2 . . . . . . . . . . . 52<br />

5.8 D=5 Reissner-Nordström <strong>black</strong> holes . . . . . . . . . . . . . . . . . . 55<br />

5.8.1 Derivation <strong>of</strong> <strong>macroscopic</strong> entropy . . . . . . . . . . . . . . . 56<br />

5.8.2 Derivation <strong>of</strong> <strong>microscopic</strong> entropy . . . . . . . . . . . . . . . 61


5.9 Link with the Strominger-Vafa result . . . . . . . . . . . . . . . . . . 63<br />

6. Mathematical Tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

6.1 Modular forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

6.2 (Weak) Jacobi forms . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

6.3 Elliptic genus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

6.3.1 Definition <strong>and</strong> properties . . . . . . . . . . . . . . . . . . . . 69<br />

6.3.2 A geometrical description . . . . . . . . . . . . . . . . . . . 71<br />

7. Counting N = 4 Dyons via Automorphic Forms . . . . . . . . . . . . . . 74<br />

7.1 Electric-magnetic duality, S-duality <strong>and</strong> dyons <strong>of</strong> charge eθ/2π . . . . 74<br />

7.2 Making use <strong>of</strong> the attractor mechanism . . . . . . . . . . . . . . . . . 76<br />

7.3 The Dijkgraaf-Verlinde-Verlinde degeneracy formula . . . . . . . . . 77<br />

7.3.1 Consistency checks . . . . . . . . . . . . . . . . . . . . . . . 79<br />

7.4 State counting <strong>and</strong> the elliptic genus . . . . . . . . . . . . . . . . . . 80<br />

8. Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

Appendix A: Geodesic Deviation <strong>and</strong> the Raychaudhuri Equation . . . . . . . . 90<br />

Appendix B: Samenvatting in het Nederl<strong>and</strong>s . . . . . . . . . . . . . . . . . . 90<br />

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

1


1. INTRODUCTION<br />

Theoretical physics is founded on two pillars: quantum theory <strong>and</strong> general relativity.<br />

The former is extremely successful in describing physics at <strong>microscopic</strong> scales, while<br />

the latter is equally successful in describing physics at cosmological scales. However,<br />

the two are very different. For example general relativity is deterministic, whereas<br />

quantum theory is not. Attempts to unify them into a quantum theory <strong>of</strong> gravity stum-<br />

ble upon the problem <strong>of</strong> the non-renormalizability <strong>of</strong> the theory. Yet one could ask:<br />

is it really necessary to have a unified theory <strong>of</strong> quantum gravity? The answer is yes.<br />

In the case <strong>of</strong> purely classical interactions, one could, without such a unified theory,<br />

obtain exact position <strong>and</strong> velocity <strong>of</strong> particles, thus violating Heisenberg’s uncertainty<br />

principle. Both quantum theory <strong>and</strong> general theory are supposed to apply everywhere<br />

<strong>and</strong> there cannot be two separate everywhere’s. Nature works <strong>and</strong> does not care about<br />

our renormalization problems. So our mission is clear. We have to find a theory <strong>of</strong><br />

quantum gravity. With this in mind, <strong>black</strong> holes are very interesting objects to study.<br />

For a complete underst<strong>and</strong>ing <strong>of</strong> <strong>black</strong> holes we need to apply both the idea <strong>of</strong> dynam-<br />

ical spacetime geometry <strong>and</strong> <strong>of</strong> quantum physics simultaneously.<br />

Classically, <strong>black</strong> holes are completely <strong>black</strong>. Objects inside their event horizon<br />

are eternally trapped. Their gravitational pull does not allow even light rays to escape.<br />

In the seventies it was found that there was a very close analogy between the laws that<br />

govern <strong>black</strong> hole physics <strong>and</strong> the four laws <strong>of</strong> thermodynamics [1]. The laws <strong>of</strong> <strong>black</strong><br />

hole mechanics become the laws <strong>of</strong> thermodynamics if one replaces the surface gravity<br />

κ <strong>of</strong> the <strong>black</strong> hole by the temperature T <strong>of</strong> a body in thermal equilibrium, the mass M<br />

<strong>of</strong> the <strong>black</strong> hole by the energy <strong>of</strong> the system E, etc. At a classical level this seems to<br />

be just a formal analogy with no physical meaning, since classically <strong>black</strong> holes have<br />

strictly zero temperature. However, Hawking showed, under very general assumptions,


1. Introduction 3<br />

that quantum mechanics implies that <strong>black</strong> holes emit particles [2]. In his so-called<br />

semi-classical approximation 1 , the <strong>black</strong> hole radiation is exactly thermal <strong>and</strong> contains<br />

no information about the state <strong>of</strong> the <strong>black</strong> hole. This leads to the famous “information<br />

loss paradox”, since infalling particles can carry information, while outgoing ones<br />

cannot [3]. As Hawking argued, this would lead to non-unitary evolution, so that the<br />

basic principles <strong>of</strong> quantum mechanics would have to be modified. Intuitively, it is<br />

clear that information should not be lost in reality (<strong>and</strong> thus that the semi-classical<br />

approximation breaks down at some point), but should somehow be returned in subtle<br />

corrections <strong>of</strong> the outgoing radiation.<br />

The above mentioned analogy between the laws <strong>of</strong> <strong>black</strong> holes <strong>and</strong> the laws <strong>of</strong><br />

thermodynamics, combined with the Hawking effect, suggests to assign to a <strong>black</strong><br />

hole <strong>of</strong> area A a ‘<strong>macroscopic</strong>’ or ‘thermodynamic’ entropy 2<br />

Smacro = A<br />

. (1.1)<br />

4<br />

This ‘<strong>macroscopic</strong>’ entropy can be measured far away from the <strong>black</strong> hole <strong>and</strong> is de-<br />

termined by its mass M, angular momentum J <strong>and</strong> charge Q. A quantum theory <strong>of</strong><br />

gravity should also be able to specify the microstates <strong>of</strong> the <strong>black</strong> hole which give rise<br />

to this macrostate. If there are N states corresponding to such a <strong>black</strong> hole, then the<br />

associated ‘<strong>microscopic</strong>’ or ‘statistical’ entropy is<br />

Smicro = log N. (1.2)<br />

These two entropies are expected to agree. The <strong>microscopic</strong> origin <strong>of</strong> <strong>black</strong> hole en-<br />

tropy was not understood until the discovery <strong>of</strong> so-called “D-branes”, dynamical ob-<br />

jects in string theory.<br />

In [4, 5] it was proposed to take the fundamental strings themselves as c<strong>and</strong>idates<br />

for the <strong>black</strong> hole microstates. This string-<strong>black</strong> hole correspondence principle pre-<br />

dicts the <strong>black</strong> hole entropy in terms <strong>of</strong> string states correctly up to a constant <strong>of</strong> order<br />

1 In the semi-classical approximation the gravitational field is treated classically (it is a solution <strong>of</strong> the<br />

Einstein equation) while the matter field is treated quantum mechanically in this classical background<br />

(one replaces the flat Minkowski metric in the wave equation by the curved metric that solves the<br />

Einstein equation).<br />

2 For the moment we set c = � = GN = kB = 1.


1. Introduction 4<br />

unity <strong>and</strong> it also explains the final state <strong>of</strong> a Schwarzschild <strong>black</strong> hole as a highly<br />

excited single string state.<br />

For charged extremal <strong>black</strong> holes (i.e., <strong>black</strong> holes with Q = M) one can apply<br />

even more quantitative tests. These turn out to be “BPS solitons” when embedded into<br />

theories with extended supersymmetry [6]. The essential properties <strong>of</strong> such objects are<br />

determined by supersymmetry, <strong>and</strong> because <strong>of</strong> this it is possible to do calculations in<br />

the weak coupling perturbation regime <strong>and</strong> then to increase the string coupling, thereby<br />

moving to a regime where the backreaction onto spacetime has led to the formation <strong>of</strong><br />

a <strong>black</strong> hole, see figure 1.1.<br />

Fig. 1.1: It turns out that one can study (the <strong>microscopic</strong> <strong>structure</strong> <strong>of</strong>) <strong>black</strong> holes at strong<br />

string coupling by doing ‘easy’ calculations in the weak string coupling regime.<br />

This weak ↔ strong coupling became a very powerful tool after it was discovered<br />

that charged <strong>black</strong> holes have a dual, perturbative description in terms <strong>of</strong> D-branes [7].<br />

Strominger <strong>and</strong> Vafa showed, by making use <strong>of</strong> these D-branes, how one can derive<br />

the Bekenstein-Hawking area-entropy relation for a class <strong>of</strong> five-dimensional extremal<br />

<strong>black</strong> holes in string theory by counting the degeneracy <strong>of</strong> BPS soliton bound states<br />

[8]. At the end <strong>of</strong> this brief introduction, <strong>and</strong> at the beginning <strong>of</strong> the real work, let us<br />

state that <strong>black</strong> holes are very interesting objects to study <strong>and</strong> may well turn out to be<br />

the ‘Hydrogen atom’ <strong>of</strong> quantum gravity.


2. BLACK HOLES IN EINSTEIN GRAVITY<br />

In this chapter we examine the geometry <strong>of</strong> classical <strong>black</strong> holes in a variety <strong>of</strong> differ-<br />

ent coordinate systems. Each coordinate system turns out to have its own advantages<br />

<strong>and</strong> disadvantages.<br />

2.1 Einstein gravity<br />

The crucial idea <strong>of</strong> Einstein gravity is that spacetime is dynamical: the geometry <strong>of</strong><br />

spacetime is determined by the distribution <strong>of</strong> matter, whereas the motion <strong>of</strong> matter is<br />

determined by the spacetime geometry.<br />

Mathematically speaking, spacetime is a four dimensional (pseudo-) Riemannian<br />

manifold with metric gµν. We choose our signature to be (− + + +). Once we re-<br />

strict the action to be at most quadratic in derivatives <strong>and</strong> ignore the possibility <strong>of</strong> a<br />

cosmological constant, the unique gravitational action is the Einstein-Hilbert action<br />

SEH = 1<br />

2κ2 �<br />

d 4 x √ −gR. (2.1)<br />

Here κ is related to Newton’s constant by κ 2 = 8πGN, g is the trace <strong>of</strong> the metric, <strong>and</strong><br />

R is the Ricci scalar. The coupling <strong>of</strong> matter to gravity is determined by the principle<br />

<strong>of</strong> minimal coupling. This principle basically amounts to taking a law <strong>of</strong> physics<br />

valid in inertial coordinates in flat spacetime, writing it in a coordinate-invariant form<br />

(usually this boils down to replacing partial derivatives by covariant ones), <strong>and</strong>, finally,<br />

asserting that the resulting law remains true in curved spacetime. We define an energy-<br />

momentum tensor <strong>of</strong> matter by<br />

where SM is the action for matter.<br />

Tµν = −2 δSM<br />

√ , (2.2)<br />

−g δg µν


2. Black Holes in Einstein Gravity 7<br />

The Einstein equations, which are the Euler-Lagrange equations that one gets by<br />

varying the combined action SEH + SM with respect to the metric,<br />

Rµν − 1<br />

2 gµνR = κ 2 Tµν, (2.3)<br />

where Rµν is the Ricci tensor, form the a central result <strong>of</strong> Einstein gravity<br />

The left-h<strong>and</strong> side is a reflection <strong>of</strong> the geometry <strong>of</strong> spacetime, while the right-h<strong>and</strong><br />

side contains all information about the location <strong>and</strong> motion <strong>of</strong> the matter.<br />

2.2 The Schwarzschild <strong>black</strong> hole<br />

First let’s take a look at the simplest <strong>black</strong> hole that general relativity predicts: a spher-<br />

ically symmetric static uncharged <strong>black</strong> hole. It is described by the Schwarzschild<br />

metric<br />

ds 2 �<br />

= − 1 − 2GM<br />

�<br />

dt<br />

r<br />

2 �<br />

+ 1 − 2GM<br />

�−1 dr<br />

r<br />

2 + r 2 dΩ 2 , (2.4)<br />

where dΩ 2 ≡ dθ 2 + sin 2 θdφ 2 is the metric on the unit two-sphere.<br />

According to Birkh<strong>of</strong>f’s theorem the Schwarzschild solution is the unique spheri-<br />

cally symmetric vacuum solution.<br />

In vacuum Tµν = 0. By taking the trace <strong>of</strong> equation (2.3) this implies R = 0,<br />

which in turn implies Rµν = 0. So the vacuum solutions to Einstein’s equations are<br />

precisely the Ricci-flat spacetimes.<br />

The coefficients <strong>of</strong> the Schwarzschild metric in the form (2.4) become infinite at<br />

r = 0 <strong>and</strong> r = 2GM. It is well known that r = 0 represents an honest singularity<br />

(as may be verified by calculation <strong>of</strong> the invariant RµνλψR µνλψ ), whereas r = 2GM is<br />

just a coordinate singularity which disappears when going to appropriate coordinates.<br />

An observer in free fall at r = 2GM would not notice anything unusual. Globally,<br />

however, the horizon is a very special place, since to a distant observer it represents<br />

the boundary <strong>of</strong> the world, or, more precisely, the boundary <strong>of</strong> the part <strong>of</strong> the world<br />

which can influence his detectors.<br />

One way in which we can study a spacetime is by looking at its causal <strong>structure</strong>.<br />

This can be done by studying the behavior <strong>of</strong> light cones. Setting ds 2 = 0 in equation


2. Black Holes in Einstein Gravity 8<br />

(2.4) gives us for radial null curves (for constant angles θ <strong>and</strong> φ)<br />

�<br />

0 = − 1 − 2GM<br />

�<br />

dt<br />

r<br />

2 �<br />

+ 1 − 2GM<br />

�−1 dr<br />

r<br />

2 , (2.5)<br />

thus<br />

dt<br />

dr<br />

�<br />

= ± 1 − 2GM<br />

�−1 . (2.6)<br />

r<br />

Clearly, in Schwarzschild coordinates, light cones appear to close up as we approach<br />

the Schwarzschild radius r = 2GM. Does this mean that a light ray that approaches<br />

r = 2GM never gets there? No! This is just an illusion, caused by the fact that our<br />

coordinates are not well behaved at the horizon. Therefor we would like to find a<br />

coordinate system that is better behaved at the horizon.<br />

We can explicitly solve the above equation (2.6) to obtain<br />

�<br />

� ��<br />

r<br />

t = ± r + 2GM ln − 1 + const. (2.7)<br />

2GM<br />

The expression in square brackets is known as the tortoise coordinate r ∗ . Since beyond<br />

r < 2GM our coordinates are not good anyway, we don’t have to be worried about the<br />

logarithm. We have that<br />

dr ∗ �<br />

�<br />

r<br />

= d r + 2GM ln<br />

2GM<br />

�� �<br />

− 1 = 1 − 2GM<br />

�−1 dr. (2.8)<br />

r<br />

Plugging this into the Schwarzschild metric gives<br />

ds 2 �<br />

= 1 − 2GM<br />

��<br />

− dt<br />

r<br />

2 + dr ∗2<br />

�<br />

+ r 2 dΩ 2 . (2.9)<br />

Here r should be thought <strong>of</strong> as a function <strong>of</strong> r ∗ . And indeed, in these coordinates we<br />

see that none <strong>of</strong> the metric coefficients is infinite at the horizon. However, we also see<br />

that in these coordinates the horizon has been pushed to r ∗ = −∞, which is not so<br />

nice. In a moment we will introduce new coordinates to repair this awkward behavior.<br />

Notice that the lines v ≡ t + r ∗ = constant (at constant angles θ <strong>and</strong> φ) are the<br />

geodesics <strong>of</strong> infalling light rays. This can immediately be read <strong>of</strong> from equation (2.7).<br />

The outgoing ones satisfy u ≡ t − r ∗ = constant. In the (v, r) coordinates we can<br />

follow future-directed timelike paths across the event horizon r = 2GM (thereby<br />

arriving in region III in figure 2.1).


2. Black Holes in Einstein Gravity 9<br />

Note that for infalling null curves we have dv/dr = 0, whereas outgoing null<br />

curves satisfy<br />

dv<br />

dr = d(t − r∗ + 2r∗ )<br />

dr<br />

= d(2r∗ )<br />

dr<br />

�<br />

= 2 1 − 2GM<br />

�−1 . (2.10)<br />

r<br />

So we see that this time the light cones don’t close up but tilt over. In the region<br />

0 < r < 2GM the local future light cone points entirely toward the singularity at<br />

r = 0.<br />

At first sight it may look as if we were done. After all, our spacetime includes<br />

r ≤ 2GM. However, we have no guarantee that there are no other directions in which<br />

to extend our manifold. In fact there are other directions. For example, we could have<br />

taken (u, r) as our new coordinates, making it possible to cross the horizon across past-<br />

directed curves (thereby arriving in region IV in figure 2.1). In this way one arrives<br />

at a different place than by crossing the horizon along future-directed curves in (v, r)<br />

coordinates.<br />

Let us now introduce new coordinates that are good all over the place:<br />

¯v = e v/4GM , ū = −e −u/4GM<br />

(2.11)<br />

In these new coordinates (¯v, ū, θ, φ), known as the Kruskal coordinates 1 , the metric<br />

becomes<br />

ds 2 = − 32G3M 3<br />

e<br />

r<br />

−r/2GM dūd¯v + r 2 dΩ 2 . (2.12)<br />

Since the metric components are nonsingular at r = 2M we can analytically continue<br />

the solution to the whole region −∞ < ū, ¯v < ∞. This is the extended Schwarzschild<br />

solution, shown in figure 2.1.<br />

For every r, t we have two solutions for ū <strong>and</strong> ¯v. Therefore the region r < 2GM<br />

occurs twice in (ū,¯v) space; these two regions are indicated by III <strong>and</strong> IV. Region III<br />

is the <strong>black</strong> hole. Once an observer has entered region III, he or she will not be able<br />

to escape from it, <strong>and</strong> will fall into the singularity at r = 0 within finite proper time.<br />

1 The related coordinates (T, R, θ, φ), where T ≡ (¯v + ū)/2 <strong>and</strong> R ≡ (¯v − ū)/2, are sometimes also<br />

called Kruskal coordinates. The nice thing about taking (T, R, θ, φ) is that in that case one coordinate<br />

is timelike while the remaining ones are spacelike.


2. Black Holes in Einstein Gravity 10<br />

u<br />

I I<br />

r=0<br />

r=0<br />

T<br />

III<br />

IV<br />

I<br />

v<br />

r=2GM<br />

R<br />

r=2GM<br />

Fig. 2.1: The Schwarzschild solution in Kruskal coordinates.<br />

Region IV is simply the time-reverse <strong>of</strong> region III. It can be thought <strong>of</strong> as a white hole.<br />

The central region, |ū| ≪ 1, |¯v| ≪ 1, will turn out to be very important in our study<br />

<strong>of</strong> the quantum mechanics <strong>of</strong> the <strong>black</strong> hole. If we ignore the curvature in this central<br />

region, <strong>and</strong> replace it by flat space, we have the so-called Rindler space.<br />

2.3 Charged <strong>black</strong> holes<br />

In this section we will briefly describe some main facts about electrically charged <strong>black</strong><br />

holes. For compactness we set G = 1 from now on. The energy-momentum tensor for<br />

electromagnetism is given by<br />

Tµν = FµρF ρ ν − 1<br />

4 gµνFρσF ρσ , (2.13)<br />

where Fµν is the electromagnetic field strength tensor. Thus the gravitational field<br />

equations are<br />

Rµν − 1<br />

2 gµνR = κ 2<br />

�<br />

FµρF ρ ν − 1<br />

ρσ<br />

gµνFρσF<br />

4<br />

�<br />

. (2.14)<br />

A generalized version <strong>of</strong> Birkh<strong>of</strong>f’s theorem now states that the unique spherically<br />

symmetric solution <strong>of</strong> equation (2.14) is the Reissner-Nordström solution<br />

ds 2 = −C(r)dt 2 + dr2<br />

C(r) + r2 dΩ 2 , (2.15)


where<br />

2. Black Holes in Einstein Gravity 11<br />

C(r) = 1 − 2M<br />

r<br />

Q2<br />

+ . (2.16)<br />

r2 In this expression Q is the charge <strong>of</strong> the <strong>black</strong> hole. By definition<br />

Qe = 1<br />

�<br />

4π<br />

⋆<br />

F, (2.17)<br />

where S 2 ∞ is an asymptotic spacelike sphere at infinity <strong>and</strong> ⋆ F is the dual <strong>of</strong> the field<br />

strength two-form F = 1<br />

2 Fµνdx µ ∧ dx ν . Here µ, ν denote the indices t, r, φ, θ.<br />

Although isolated magnetic charges have never been observed in nature, we may<br />

write in general Q = � Q 2 e + Q 2 m where Qe is the electric <strong>and</strong> Qm the magnetic charge<br />

(which is defined like the electric charge (2.17) but without the star). In the electrically<br />

charged case the radial component <strong>of</strong> the electric field is Er = Frt = Qe/r 2 <strong>and</strong> in<br />

the magnetically charged case Br = Fθφ = Qm/r 2 . (The electric field is curl free,<br />

whereas the magnetic field is divergence free.)<br />

Clearly for a charged <strong>black</strong> hole the horizon <strong>structure</strong> is more complicated than for<br />

the Schwarzschild <strong>black</strong> hole. By solving<br />

1 − 2M<br />

r<br />

S 2 ∞<br />

Q2<br />

+ = 0 (2.18)<br />

r2 we get two horizons: an inner one (r−) <strong>and</strong> an outer one (r+) located at<br />

� �<br />

r± = M 1 ± 1 − Q2<br />

M 2<br />

�<br />

. (2.19)<br />

There are three cases to distinguish: Q 2 > M 2 , Q 2 < M 2 , <strong>and</strong> Q 2 = M 2 .<br />

In the case Q 2 > M 2 the metric is completely regular all the way down to r = 0.<br />

At r = 0 we have a singularity: a naked one, which according to the cosmic censorship<br />

conjecture is not physical. It is however good to note that objects satisfying Q 2 > M 2<br />

do exist in nature (an electron is an example <strong>of</strong> such an object), but just cannot be<br />

described by classical general relativity.<br />

The case M 2 > Q 2 is called the non-extreme Reissner-Nordström <strong>black</strong> hole <strong>and</strong><br />

is certainly the most realistic case, but we will not discuss it here.<br />

Our main point <strong>of</strong> focus will be on the somewhat less realistic, though mathemat-<br />

ically easier to study, so-called extreme Reissner-Nordström <strong>black</strong> holes, for which


2. Black Holes in Einstein Gravity 12<br />

the charge is equal to the mass, Q 2 = M 2 . In this case the two horizons coincide at<br />

r+ = r− = M.<br />

2.3.1 The extreme Reissner-Nordström solution<br />

Plugging Q2 = M 2 into equation (2.16) gives<br />

�<br />

C(r) = 1 − Q<br />

�2 . (2.20)<br />

r<br />

The metric thus takes the form<br />

ds 2 � �2 r − Q<br />

= − dt<br />

r<br />

2 � �2 r<br />

+ dr<br />

r − Q<br />

2 + r 2 dΩ 2 . (2.21)<br />

It is convenient to define a shifted radial coordinate<br />

ρ ≡ r − Q. (2.22)<br />

We can write the metric as<br />

ds 2 � �2 ρ<br />

= − dt<br />

ρ + Q<br />

2 � �2 ρ + Q<br />

+ d(ρ + Q)<br />

ρ<br />

2 + (ρ + Q) 2 dΩ 2<br />

(2.23)<br />

�<br />

= − 1 + Q<br />

�−2 dt<br />

ρ<br />

2 �<br />

+ 1 + Q<br />

�2 [dρ<br />

ρ<br />

2 + ρ 2 dΩ 2 ]. (2.24)<br />

And since dρ 2 + ρ 2 dΩ 2 is just the flat metric in three spatial dimensions, we can write<br />

where<br />

ds 2 = −H −2 (�x)dt 2 + H 2 (�x)d�x 2 , (2.25)<br />

H = 1 + Q<br />

. (2.26)<br />

|�x|<br />

For N <strong>black</strong> holes with charges Qi (or equivalently multiple masses Mi) sitting at<br />

positions �xi we have [9]<br />

H = 1 +<br />

N�<br />

i=1<br />

Qi<br />

. (2.27)<br />

|�x − �xi|<br />

Such a stationary configuration <strong>of</strong> N extremal <strong>black</strong> holes is possible because <strong>of</strong> the<br />

exact cancelation <strong>of</strong> the electric repulsion <strong>and</strong> gravitational attraction. This is called<br />

the no-force property.


2. Black Holes in Einstein Gravity 13<br />

We can find an additional special property <strong>of</strong> the extremal <strong>black</strong> hole geometry as<br />

follows. Near the horizon ρ = 0 the metric (see (2.23) <strong>and</strong> (2.24)) behaves like<br />

Defining another new coordinate<br />

so that dz/dr = Q 2 /r 2 = z/r, we find<br />

ds 2 → − ρ2<br />

Q 2 dt2 + Q2<br />

ρ 2 dρ2 + Q 2 dΩ 2 2<br />

z ≡ Q2<br />

ρ<br />

(2.28)<br />

(2.29)<br />

ds 2 → Q2<br />

z 2 (−dt2 + dz 2 ) + Q 2 dΩ 2 2. (2.30)<br />

This is AdS2×S 2 , the direct product <strong>of</strong> an Anti-de-Sitter spacetime ((SO(2, 1)/SO(1, 1))<br />

with a sphere. This is a maximally symmetric space. It is interesting to note that the<br />

Reissner-Nordström spacetime is also maximally symmetric <strong>and</strong> asymptotically flat.<br />

The extreme Reissner-Nordström <strong>black</strong> hole will turn out to possess a certain super-<br />

symmetry which makes it an interesting <strong>and</strong> relatively easy object to study.


3. MOTIVATION FOR BLACK HOLE THERMODYNAMICS<br />

An amazing result from <strong>black</strong> hole physics is that one can derive a certain set <strong>of</strong> laws<br />

for <strong>black</strong> holes that has the same <strong>structure</strong> as the laws <strong>of</strong> thermodynamics. They are<br />

called the laws <strong>of</strong> <strong>black</strong> hole mechanics [1]. One can think <strong>of</strong> these “<strong>black</strong> hole laws”<br />

as theorems within the theory <strong>of</strong> general relativity. In this section we will investigate<br />

the question whether this curious analogy between thermodynamics <strong>and</strong> <strong>black</strong> hole<br />

laws is more than a formal coincidence.<br />

Let us first introduce some terminology. We assume that the reader is familiar with<br />

Killing vectors (see for example [10] or [11]). A vector ξ is called a Killing vector if<br />

it satisfies the Killing equation<br />

∇µξν + ∇νξµ = 0, (3.1)<br />

where ∇µ is the covariant derivative. Remember that if a spacetime has a Killing<br />

vector, we can find a coordinate system in which the metric is independent <strong>of</strong> one <strong>of</strong><br />

the coordinates, <strong>and</strong> the quantity pµξ µ will be constant along geodesics with tangent<br />

vector p µ . If a Killing vector field ξ µ is null along some null hypersurface Σ, we say<br />

that Σ is a Killing horizon 1 <strong>of</strong> ξ µ .<br />

Now let K be a Killing horizon with normal Killing field ξ µ . The Killing horizon<br />

is defined by the equation ξ ν ξν = 0. Therefore the gradient ∇ µ (ξ ν ξν) is normal to the<br />

horizon. By definition ξ µ is also normal to the horizon. So the two must be proportional<br />

to each other. In other words, there exist a proportionality function κS on K. κS is<br />

defined by the equation<br />

∇ µ (ξ ν ξν) = −2κSξ µ<br />

(3.2)<br />

1 It can be proven that in Einstein gravity event horizons <strong>of</strong> stationary <strong>black</strong> holes are always Killing<br />

horizons (i.e., one can always find an appropriate Killing vector field). See [12].


3. Motivation for Black Hole Thermodynamics 15<br />

<strong>and</strong> is called the surface gravity 2 . It turns out that this so-called surface gravity is<br />

actually a constant function on the horizon,<br />

κS = const. (3.3)<br />

This is the case whenever Einstein’s equations hold, with the energy-momentum tensor<br />

satisfying the dominant energy condition (which basically means that the energy <strong>and</strong><br />

pressure always remain positive), see [1].<br />

This constancy <strong>of</strong> the surface gravity for a stationary <strong>black</strong> hole is called the zeroth<br />

law <strong>of</strong> <strong>black</strong> hole mechanics. It is reminiscent <strong>of</strong> the zeroth law <strong>of</strong> thermodynam-<br />

ics, which states that the temperature is uniform everywhere in a system in thermal<br />

equilibrium. This suggests that surface gravity is analogous to temperature.<br />

The first law <strong>of</strong> <strong>black</strong> hole mechanics is a law <strong>of</strong> energy conservation. It gives a<br />

relation between the change in mass (δM), horizon area (δA), charge (δQ), <strong>and</strong> angular<br />

momentum (δJ) <strong>of</strong> a <strong>black</strong> hole when it is perturbed, as follows<br />

δM = κS<br />

δA + µδQ + ΩδJ. (3.4)<br />

8π<br />

Obviously, it has the same mathematical form as the first law <strong>of</strong> thermodynamics<br />

dE = T dS − P dV + µdN. (3.5)<br />

Note that the µ in equation (3.5) st<strong>and</strong>s for the chemical potential, but in equation (3.4)<br />

is does not. (We will soon find out where it does st<strong>and</strong> for.) Above we concluded that<br />

surface gravity is analogous to temperature. We can now guess that area plays the role<br />

<strong>of</strong> entropy.<br />

3.1 Two derivations <strong>of</strong> the first <strong>black</strong> hole law<br />

We will now present a short derivation <strong>of</strong> equation (3.4). Obviously, we will not find<br />

the ΩδJ term, since the <strong>black</strong> hole we study is not spinning. Starting from equation<br />

2 One can think <strong>of</strong> the surface gravity κS as the limiting force which would have to be exerted at<br />

infinity to keep an object stationary at the horizon. Another - basically the same - way <strong>of</strong> thinking about<br />

it, is that surface gravity is the acceleration <strong>of</strong> a static particle near the horizon as measured at spatial<br />

infinity.


(2.16) we have<br />

So<br />

3. Motivation for Black Hole Thermodynamics 16<br />

C ′ (r+)dr+ = dC(r+) = − 2dM<br />

r+<br />

1<br />

2 C′ (r+)r+dr+ = −dM + Q<br />

+ 2Q<br />

r2 dQ. (3.6)<br />

+<br />

r+<br />

dQ. (3.7)<br />

Now we can plug in the area <strong>of</strong> the (outer) horizon into this equation using A = 4πr 2 +,<br />

so dA = 8πr+dr+ <strong>and</strong> we get<br />

dM = −1<br />

16π C′ (r+)dA + µdQ, (3.8)<br />

where we put µ ≡ Q/r+. We see that this indeed looks like equation (3.4), but without<br />

the ΩδJ term. This is as expected, since our <strong>black</strong> hole is not spinning.<br />

In order to find this spinning term one has to consider a rotating <strong>black</strong> hole. We<br />

will now give another derivation <strong>of</strong> equation (3.4), but this time for a rotating <strong>black</strong><br />

hole. The metric for a charged rotating <strong>black</strong> hole is the Kerr-Newman metric, [13]<br />

[14]. We will use it in the form<br />

ds 2 � �<br />

2 2 ∆ − a sin θ<br />

= −<br />

dt<br />

Σ<br />

2 − 2a sin2 θ(r2 + a2 − ∆)<br />

dtdφ +<br />

Σ<br />

� �<br />

2 2 2 2 2 (r + a ) − ∆a sin θ<br />

+<br />

sin<br />

Σ<br />

2 θdφ 2 + Σ<br />

∆ dr2 + Σdθ 2 , (3.9)<br />

where � Σ = r 2 + a 2 cos 2 θ,<br />

∆ = r 2 + a 2 + Q 2 − 2Mr.<br />

Here a is the angular momentum per unit mass, a = J/M. Notice that for a = 0 the<br />

Kerr-Newman metric reduces to the Reissner-Nordström metric.<br />

Now we will study the variation <strong>of</strong> the area <strong>of</strong> the event horizon <strong>of</strong> a Kerr <strong>black</strong><br />

hole (A) with respect to M <strong>and</strong> J. Let us first write down an explicit equation for A:<br />

A =<br />

=<br />

�<br />

�<br />

r=r+<br />

r=r+<br />

dθdφ √ gθθgφφ<br />

�<br />

dθdφ Σ (r2 + + a2 ) 2 sin2 Σ<br />

= 4π(2Mr+ − Q 2 ), (3.10)


3. Motivation for Black Hole Thermodynamics 17<br />

where the rotating version <strong>of</strong> equation (2.19) was used, namely<br />

r+ = M + � M 2 − a 2 − Q 2 . (3.11)<br />

We are now ready to calculate the variation <strong>of</strong> A with respect to M <strong>and</strong> J ≡ Ma.<br />

Ignoring the δQ contribution for the moment we have from equation (3.10) that<br />

Let’s work this out:<br />

So, since<br />

we get<br />

1<br />

8π δA = r+δM + Mδr+. (3.12)<br />

δr+ = δ(M + � M 2 − a2 − Q2 )<br />

= δM + 1<br />

2 (M 2 − a 2 − Q 2 1<br />

−<br />

) 2 2(MδM − aδa)<br />

= δM + (r+ − M) −1 (MδM − aδa). (3.13)<br />

δa = δ(J/M) = 1 a<br />

δJ − δM (3.14)<br />

M M<br />

1<br />

8π = r+δM + MδM + (r+ − M) −1 {(M 2 + a 2 )δM − aδJ}. (3.15)<br />

Rewriting this gives<br />

δM = r+ − M<br />

8π(r2 + + a2 a<br />

δA +<br />

) r2 δJ. (3.16)<br />

+ + a2 We see that κS should equal (r+ − M)/(r 2 + + a 2 ) <strong>and</strong> ΩH = a/(r 2 + + a 2 ) in order to<br />

get the desired<br />

δM = κS<br />

8π δA + ΩHδJ. (3.17)<br />

Note that if we would also have varied with respect to Q we would have found an<br />

additional µδQ term in the above equation, as one can immediately see from equation<br />

(3.10).


3. Motivation for Black Hole Thermodynamics 18<br />

3.2 Black hole entropy <strong>and</strong> the generalized second law<br />

So what do we learn from this first law <strong>of</strong> <strong>black</strong> hole mechanics? By comparing<br />

equations (3.4) <strong>and</strong> (3.5) it seems that <strong>black</strong> holes have entropy (<strong>and</strong> that their entropy<br />

is proportional to their area). This was already anticipated by Bekenstein, who noted<br />

that we can violate the second law <strong>of</strong> thermodynamics by throwing matter into a <strong>black</strong><br />

hole. It is clear that the second law must be reformulated in the presence <strong>of</strong> <strong>black</strong><br />

holes. The way this is done is by associating a certain entropy with each <strong>black</strong> hole<br />

<strong>and</strong> reformulating the second law to a generalized second law (GSL) which states that<br />

the sum <strong>of</strong> the entropy outside the <strong>black</strong> hole <strong>and</strong> the entropy <strong>of</strong> the <strong>black</strong> hole itself<br />

will never decrease [15].<br />

It is not obvious at all that this should work, <strong>and</strong> actually at a purely classical level<br />

it simply appears to be false. Perhaps the easiest way to see why, is by considering a<br />

box containing entropy in the form <strong>of</strong> radiation which can be lowered to the horizon<br />

<strong>of</strong> a <strong>black</strong> hole <strong>and</strong> be dropped in. Now, in principle, for an ideal infinitesimal box all<br />

<strong>of</strong> its energy can be extracted at infinity, so when the box falls into the <strong>black</strong> hole, no<br />

mass is added to the <strong>black</strong> hole. Therefore its area <strong>and</strong> thus its entropy is unchanged.<br />

The entropy <strong>of</strong> the exterior, however, has decreased. It thus appears that the GSL is<br />

simply not true at a purely classical level.<br />

Another problem was that if <strong>black</strong> holes really have entropy, they must have a<br />

temperature because <strong>of</strong> the thermodynamic identity 1/T = ∂S/∂E. Since E ∼ M<br />

<strong>and</strong> S ∼ A ∼ M 2 we have that ∂S/∂M ∼ M ∼ T −1 <strong>and</strong> our guess would thus be that<br />

<strong>black</strong> holes have a temperature T ∼ M −1 . Indeed, Hawking’s calculation [2] shows<br />

that a <strong>black</strong> hole <strong>of</strong> mass M indeed has a temperature<br />

<strong>and</strong> we get that<br />

T =<br />

�<br />

, (3.18)<br />

8πGM<br />

S = A<br />

. (3.19)<br />

4G�<br />

We will look at a derivation <strong>of</strong> this using the path integral approach in section 4.3.<br />

This solves yet another mystery, because it looked like the units did not match, since


3. Motivation for Black Hole Thermodynamics 19<br />

entropy is dimensionless, whereas horizon area is length squared. But now we see that<br />

indeed S = A/4 where the area is measured in units <strong>of</strong> the Planck length.<br />

There are also second <strong>and</strong> third laws <strong>of</strong> <strong>black</strong> hole mechanics, which we will not<br />

discuss here, see for example [1].<br />

3.3 The holographic principle<br />

At this point it is worthwhile to pause for a moment <strong>and</strong> wonder about the strange fact<br />

that <strong>black</strong> hole entropy scales with its area <strong>and</strong> not with its volume, as we are used<br />

for thermodynamic systems. (Consider for example a simple thermodynamic system<br />

<strong>of</strong> volume V divided in little unit cubes which can take either value 0 or 1. The total<br />

number <strong>of</strong> states for such a system is 2 V . To get the entropy we have to take the<br />

logarithm <strong>of</strong> this quantity, so we indeed see that S ∝ V .)<br />

Imagine the following. Suppose we have a volume V somewhere in the universe;<br />

for simplicity let’s take it to be a sphere with radius R. It seems logical that the amount<br />

<strong>of</strong> information (or entropy) in this volume depends on the volume V ∝ R 3 . Our claim,<br />

<strong>and</strong> also the claim <strong>of</strong> the holographic principle which is due to ’t Ho<strong>of</strong>t [16], is that<br />

this is not the case! The amount <strong>of</strong> information we can store in a sphere with radius R<br />

is proportional to the area A ∝ R 2 . So, so to speak, if we can put a maximum <strong>of</strong> 100<br />

Gigabyte into a spherical laptop with radius R, we can put only 400 (<strong>and</strong> not 800!)<br />

Gigabyte into a laptop <strong>of</strong> radius 2R.<br />

We can underst<strong>and</strong> this as follows. Our volume V can be turned into a <strong>black</strong> hole<br />

in the same region by throwing matter in. Its entropy will than be equal to the area<br />

SBH ∼ A surrounding V , in suitable units. Thus, we learn that it is impossible for<br />

the entropy <strong>of</strong> our volume to scale with the volume, since we always can make a <strong>black</strong><br />

hole with entropy proportional to area out <strong>of</strong> it.<br />

So in short, the holographic principle states that information (entropy) scales with<br />

area <strong>and</strong> not with volume, as one would expect. More precisely, it states that a field<br />

theory on a two-dimensional closed surface suffices to describe all processes that take<br />

place in the three-dimensional volume within this surface. Further evidence for holog-


3. Motivation for Black Hole Thermodynamics 20<br />

raphy is given by the AdS/CFT correspondence [17], which basically is a dictionary<br />

encoding the equivalence between string theory on five dimensional Anti-de-Sitter<br />

space <strong>and</strong> conformal fields on its four dimensional boundary.<br />

3.4 Derivation <strong>of</strong> Einstein equations from thermodynamics<br />

Let us turn the logic around <strong>and</strong> show that not only does general relativity give ther-<br />

modynamic kind <strong>of</strong> identities, but that one actually can derive the Einstein equations<br />

from thermodynamics principles.<br />

In this section we will study Jacobson’s derivation [18] <strong>of</strong> the Einstein equations<br />

from thermodynamical principles in some detail. In turns out that locally the Einstein<br />

equations can be derived from the proportionality <strong>of</strong> entropy <strong>and</strong> horizon area together<br />

with the fundamental relation δQ = T dS. If this is indeed the right way to do things,<br />

gravity should not be quantized but rather seen as a many particle effect (just like, for<br />

example, sound waves). For some general relativity background (geodesic deviation,<br />

Raychaudhuri equation, etc.), see Appendix A.<br />

Jacobson considered locally static horizons <strong>and</strong> created locally accelerated ob-<br />

servers in order to study information entering the spacetime through the local past<br />

horizon <strong>of</strong> the local Rindler space. Assuming local equilibrium he then derived the<br />

Einstein equations. He was thus led to the conclusion that the Einstein equations give<br />

an effective field description <strong>of</strong> some more fundamental fields, <strong>and</strong> actually do not<br />

describe a fundamental force: the Einstein equation is an equation <strong>of</strong> state.<br />

So what exactly is meant here? It is clear what we mean by heat in thermodynam-<br />

ics: it is energy that flows between degrees <strong>of</strong> freedom that are not <strong>macroscopic</strong>ally<br />

observable. We will define heat to be energy that flows across some causal horizon<br />

(not necessarily a <strong>black</strong> hole event horizon). Next, we have to say what we mean with<br />

the temperature <strong>of</strong> the system into which the heat is flowing. We define it to be the<br />

Unruh temperature [19] associated with a uniformly accelerated observer hovering just<br />

inside the horizon. Since different accelerated observers will measure different things,<br />

the same observer should be used to measure the heat flow (which is defined by the


3. Motivation for Black Hole Thermodynamics 21<br />

energy flux through the horizon). Another quantity appearing in the relation Q = T dS<br />

that needs our attention is the entropy, S. We use a comoving volume element which<br />

moves along with the energy towards the horizon to define the entropy. As the volume<br />

element falls towards the horizon it is flattened (as seen by an outside observer) <strong>and</strong> the<br />

entropy <strong>of</strong> the volume element scales as the surface area <strong>of</strong> the local horizon. Having<br />

identified all the relevant quantities, we will now derive the (local) Einstein equations.<br />

Let p be some point in the spacetime. According to the equivalence principle we<br />

can find local coordinates in which the metric is Minkowski with vanishing first deriva-<br />

tives at p. A local horizon can be created by considering Lorentz boosted observers<br />

near p, <strong>and</strong> is called a “local Rindler horizon”. In Jacobson’s terminology, we have<br />

a small spacelike 2-surface element P through p, whose past directed null normal<br />

congruence to one side (the “inside”) has vanishing expansion <strong>and</strong> shear at p (see Ap-<br />

pendix A). It is always possible to choose this P such that both the expansion <strong>and</strong><br />

shear vanish close to p. He calls the past horizon <strong>of</strong> such a p the “local Rindler horizon<br />

<strong>of</strong> P”.<br />

Let χ µ be a (local) Killing field generating boosts orthogonal to P <strong>and</strong> vanishing<br />

at P. Note that our Rindler spacetime is local, <strong>and</strong> thus posses only local symmetries,<br />

which means that χ µ has the Killing property only in a spacelike neighborhood <strong>of</strong> p. (It<br />

exists also in the future <strong>and</strong> past <strong>of</strong> p, but does not satisfy Killing’s equation there.) The<br />

heat flow is defined to be Tµνχ µ , where Tµν is the matter energy-momentum tensor.<br />

Now let us consider any such local Rindler horizon through a spacetime point p<br />

(see figure 3.1). We assume that all the heat flux across the horizon is energy carried<br />

by matter. Since we take the thermodynamic limit we can assume all fluctuations in<br />

Tµν to be negligible. The heat flux to the past <strong>of</strong> P can now be obtained by integrating<br />

the boost-energy current <strong>of</strong> the matter, Tµνχ µ , over a pencil <strong>of</strong> generators <strong>of</strong> the “in-<br />

side” past horizon H <strong>of</strong> P. The χ µ can be reparameterised by using vectors k µ , which<br />

are tangent to the horizon generators for an affine parameter λ, which vanishes at a<br />

neighborhood <strong>of</strong> point p <strong>and</strong> is negative to the past <strong>of</strong> P, as follows: χ µ = −κλk µ ,<br />

where κ is the acceleration <strong>of</strong> the Killing orbit on which the norm <strong>of</strong> χ is one. The<br />

volume dV over which we integrate is the volume <strong>of</strong> the spacelike hypersurface gen-


3. Motivation for Black Hole Thermodynamics 22<br />

¢¡<br />

a<br />

Q<br />

Fig. 3.1: Spacetime diagram showing the heat flux δQ across the local Rindler horizon H <strong>of</strong> a<br />

2-surface element P. Each point in the diagram represents a two dimensional spacelike<br />

surface. The hyperbola is a uniformly accelerated worldline, <strong>and</strong> χ a is the approximate<br />

boost Killing vector on H. Adapted from [18].<br />

erated by the timelike geodesic vector field k µ . This is the natural normal vector to the<br />

volume. Putting all this in one equation, we get<br />

�<br />

δQ = −κ<br />

λTµνk<br />

H<br />

µ k ν dλdA, (3.20)<br />

where we substituted dV = dλdA (where A is the area element on a cross section <strong>of</strong><br />

the horizon) since near the horizon the volume becomes flatter <strong>and</strong> flatter, as seen by<br />

the observer on χ µ . This gives us, on the one h<strong>and</strong>, from dS = δQ/T , an expression<br />

for dS. On the other h<strong>and</strong>, as we have seen, entropy seems to be proportional with<br />

area <strong>and</strong> we can assume that the entropy variation associated with a piece <strong>of</strong> horizon<br />

satisfies dS = ηδA, where δA is the area variation <strong>of</strong> a cross section <strong>of</strong> a pencil <strong>of</strong><br />

generators <strong>of</strong> H <strong>and</strong> η is a constant (equal to 1/4). So by working out δA we can find<br />

some condition. It will turn out that this condition is precisely the Einstein equation.<br />

One can show that the change <strong>of</strong> the horizon area is<br />

dA<br />

dλ<br />

= θA, (3.21)<br />

with θ the expansion <strong>of</strong> the congruence <strong>of</strong> null geodesics k µ , which generates the<br />

horizon. Integrating (3.21) the horizon area element at affine parameter λ is<br />

A(λ) = A0e R λ<br />

0 dλ′ θ . (3.22)


3. Motivation for Black Hole Thermodynamics 23<br />

where A0 is the surface element <strong>of</strong> the horizon at λ = 0. Exp<strong>and</strong>ing the exponent we<br />

find<br />

A(λ) ≈ A0(1 + λθ(0) + λ2<br />

2<br />

dθ<br />

(0)). (3.23)<br />

dλ<br />

Noting that, due to the fact that the accelerated observer observes a local heat bath<br />

which is homogeneous, there is no contribution from the first order term <strong>and</strong> that we<br />

can use the Raychaudhuri equation (8.11) to work out dθ/dλ. We get<br />

A(λ) ≈ A0<br />

�<br />

1 + λ2<br />

2 Rµνk µ k ν<br />

�<br />

. (3.24)<br />

Here we used the fact that the local Rindler horizon was chosen to be instantaneously<br />

stationary at P, so that θ <strong>and</strong> σ (the shear) vanish at P. We can thus neglect the con-<br />

tributions from θ 2 <strong>and</strong> σ 2 = σµνσ µν when integrating to find θ near P. Summarizing<br />

all this,<br />

�<br />

δA = − λRabk<br />

H<br />

µ k ν dλdA. (3.25)<br />

As announced, comparing (3.20) <strong>and</strong> (3.25) we conclude that<br />

or, for λ �= 0,<br />

− 2πλTµνk µ k ν dλdA = ηλRµνk µ k ν dλdA, (3.26)<br />

− 2π<br />

η Tµνk µ k ν = Rµνk µ k ν . (3.27)<br />

Because the k µ are null vector fields we can add a term fgµνk µ k ν proportional to the<br />

metric to the right h<strong>and</strong> side (3.27). As we will see in a moment, this corresponds to<br />

the freedom one has in adding an integration constant in the Einstein equation. There<br />

is a close relation between f <strong>and</strong> Λ, the cosmological constant. We have<br />

− 2π<br />

η Tµνk µ k ν = Rµνk µ k ν + fgµνk µ k ν . (3.28)<br />

Since δQ = T dS holds for any family <strong>of</strong> geodesics, the above equation holds for all<br />

null vector fields on the horizon <strong>and</strong> we have<br />

2π<br />

η Tµν + Rµν + fgµν = 0. (3.29)


3. Motivation for Black Hole Thermodynamics 24<br />

We would like to find out what f is. A way to find this is out, is to take the diver-<br />

gence <strong>of</strong> equation (3.29). The divergence <strong>of</strong> the energy momentum tensor vanishes<br />

(by definition) <strong>and</strong> using the Bianchi identities<br />

we find<br />

∇µ(Rµν − R<br />

2 gµν) = 0 (3.30)<br />

f = − R<br />

2<br />

with C some constant. Plugging this f into equation 3.29 gives<br />

Rµν − 1<br />

2 gµνR + Cgµν = − 2π<br />

η Tµν<br />

+ C (3.31)<br />

(3.32)<br />

which is precisely the Einstein equation if we identify C with (minus) the cosmological<br />

constant, C = Λ, <strong>and</strong> with η = 1/4.<br />

A perhaps somewhat preliminary conclusion <strong>of</strong> all this would be that the general<br />

theory <strong>of</strong> relativity can be interpreted as a statistical theory. The equivalence principle<br />

corresponds to the statement that the neighborhood <strong>of</strong> each point in spacetime is in<br />

local thermodynamic equilibrium.<br />

3.5 Wald entropy<br />

We saw that for the Einstein-Hilbert action (2.1) we get the Bekenstein-Hawking equa-<br />

tion S = AH/4. In general (in string theory) we can get a more complicated action I<br />

with corrections involving the Riemann tensor <strong>and</strong> other fields. This will modify the<br />

laws <strong>of</strong> <strong>black</strong> hole thermodynamics. However, Wald [20] has a ‘simple’ solution for<br />

this: he gives the following formal expression for the entropy<br />

�<br />

ˆS<br />

δI<br />

:= 2π<br />

S2 ɛ<br />

δRµναβ<br />

µν ɛ ρσ√ hd 2 Ω. (3.33)<br />

Here Rµναβ is regarded as formally independent <strong>of</strong> the metric.<br />

As an example let us calculate the Wald entropy for the Einstein-Hilbert action<br />

L = 1<br />

16π Rµναβg µα g νβ . (3.34)


We get<br />

Thus<br />

3. Motivation for Black Hole Thermodynamics 25<br />

∂L<br />

∂Rµναβ<br />

= 1 1<br />

16π 2 (gµαg νβ − g να g µβ ). (3.35)<br />

ˆS = 1<br />

�<br />

8 S2 1<br />

2 (gµαg νβ − g να g µβ √<br />

2<br />

)ɛµνɛαβ hd σ)). (3.36)<br />

Only the tt <strong>and</strong> rr terms in the metric survive. So we get<br />

ˆS = 1<br />

�<br />

8<br />

For the Schwarzschild <strong>black</strong> hole<br />

so we get<br />

g tt =<br />

g rr =<br />

which is indeed what we hoped to get.<br />

S2 2g tt g rr√ hd 2 σ. (3.37)<br />

�<br />

1 − 2GM<br />

�<br />

, (3.38)<br />

r<br />

�<br />

1 − 2GM<br />

�−1 , (3.39)<br />

r<br />

ˆS = 1<br />

�<br />

4 S2 √<br />

2<br />

hd σ = A/4, (3.40)<br />

Wald’s redefinition <strong>of</strong> entropy is needed (besides the proper inclusion <strong>of</strong> higher<br />

curvature terms on the supergravity side) to find agreement between geometric entropy<br />

computed in four-dimensional N = 2 supergravity <strong>and</strong> the statistical entropy found by<br />

state counting [21].


4. QUANTUM BLACK HOLES<br />

First, we shall describe the st<strong>and</strong>ard picture <strong>of</strong> Rindler space arising from the near-<br />

horizon approximation <strong>of</strong> the Schwarzschild metric. Then we derive the Hawking<br />

effect following the treatment by ’t Ho<strong>of</strong>t in [22]. Some attention will be paid to<br />

the gravitational back reaction. We also comment on the information loss paradox, or<br />

better said, on information recovery. Finally, we discuss an alternative approach which<br />

makes use <strong>of</strong> Feynman’s path integral method.<br />

4.1 Rindler space<br />

Let us start with motivating the introduction <strong>of</strong> the Rindler geometry. It is not difficult<br />

to see that the (t − r)-part <strong>of</strong> the Schwarzschild metric (2.4) is very well approximated<br />

by flat space near the horizon. Introducing the positive coordinate x defined by<br />

r = 2M + x2<br />

8M<br />

(4.1)<br />

near the horizon r ≈ 2M (or r ≈ 0) we get for the Schwarzschild metric the approxi-<br />

mation<br />

ds 2 ≈ −(κx) 2 dt 2 + dx 2<br />

+ (2M)<br />

� �� �<br />

2−dim Rindler space<br />

2 dΩ 2<br />

,<br />

� �� �<br />

(4.2)<br />

2−sphere<br />

where κ = 1/4M is the surface gravity <strong>of</strong> the Schwarzschild metric <strong>and</strong> where we<br />

used the near horizon expansion<br />

1 − 2M<br />

r<br />

= 1 −<br />

2M<br />

2M + x2<br />

8M<br />

= 1 + x2<br />

16M 2 − 1<br />

1 + x2<br />

16M 2<br />

= x2 x4<br />

+ O( ). (4.3)<br />

16M 2 M 4<br />

We can easily check that x = 0 is just a coordinate singularity <strong>and</strong> that Rindler space-<br />

time is just (a part <strong>of</strong>) Minkowski space in unusual coordinates by introducing Kruskal-<br />

type coordinates<br />

U = −xe −κt , V = xe κt . (4.4)


One indeed checks that<br />

dUdV =<br />

4. Quantum Black Holes 27<br />

�<br />

∂U ∂U<br />

dx +<br />

∂x ∂t dt<br />

��<br />

∂V ∂V<br />

dx +<br />

∂x ∂t dt<br />

�<br />

= (−e −κt dx + κxe −κt dt)(e κt dx + κxe κt dt)<br />

= (κx) 2 dt 2 − dx 2 = −ds 2 , (4.5)<br />

which is Minkowski space in null coordinates 1 . So the Rindler coordinates with x > 0<br />

cover precisely the U < 0, V > 0 region <strong>of</strong> Minkowski space.<br />

For the study <strong>of</strong> uniformly accelerated observers (simply called ‘Rindler observers’;<br />

we encountered them already in section 3.4) it is convenient to introduce new coordi-<br />

nates (ξ, η) on the right wedge <strong>of</strong> two-dimensional Minkowski space (x > |t|), given<br />

by<br />

t = a −1 e aξ sinh aη, (4.6)<br />

x = a −1 e aξ cosh aη, (4.7)<br />

where a is a positive constant. As one can easily check, the near horizon two-dimensional<br />

Rindler metric becomes<br />

Lines <strong>of</strong> constant ξ are hyperbolae<br />

ds 2 = e 2aξ (dη 2 − dξ 2 ). (4.8)<br />

x 2 − t 2 = a −2 e 2aξ = constant, (4.9)<br />

<strong>and</strong> represent the world lines <strong>of</strong> uniformly accelerated observers, while lines <strong>of</strong> con-<br />

stant η are straight (from (4.6) <strong>and</strong> (4.7) we see that x ∝ t). This is indicated in figure<br />

4.1.<br />

We can also define coordinates in the region II by<br />

t = −a −1 e aξ sinh aη, (4.10)<br />

x = −a −1 e aξ cosh aη. (4.11)<br />

We will use the coordinates (ξ, η) for both region I <strong>and</strong> II although their ranges are not<br />

the same in each region. We will simply indicate to which region we are referring. The<br />

1 Setting U = T − X <strong>and</strong> V = T + X we get the st<strong>and</strong>ard form ds 2 = −dT 2 + dX 2 .


4. Quantum Black Holes 28<br />

t<br />

constant<br />

z<br />

constant<br />

Fig. 4.1: Rindler spacetime is a wedge <strong>of</strong> Minkowski spacetime. Lines <strong>of</strong> constant ξ represent<br />

the world lines <strong>of</strong> uniformly accelerated observers.<br />

advantage is that we will be able to use the same metric, namely the one in equation<br />

(4.8), for both region I <strong>and</strong> II.<br />

The important point to keep in mind is that Rindler space is a model for a gravita-<br />

tional field, although there are no real gravitational forces present.<br />

4.2 Field theory in Rindler space<br />

For the above discussion we could safely ignore two spatial directions, but now we<br />

want to study four <strong>and</strong> not two dimensional Minkowski space, so let us repair this:<br />

we consider Minkowski space described by the coordinates t, z <strong>and</strong> ˜x = (x, y) <strong>and</strong><br />

introduce new coordinates ξ, η <strong>and</strong> ˜x, given by<br />

z = a −1 e aξ cosh aη,<br />

t = a −1 e aξ sinh aη, (4.12)<br />

˜x = ˜x.<br />

Note that the coordinate z is playing the role <strong>of</strong> the coordinate x in section 4.1.<br />

In this section we will consider a quantum field theory, possibly with interacting<br />

particles, in a flat Minkowski background. The Hamiltonian, HM, <strong>of</strong> such a theory can<br />

usually be written as<br />

HM =<br />

�<br />

HM(�x)d 3 �x, (4.13)


4. Quantum Black Holes 29<br />

where HM is the Hamiltonian density. A time boost in Rindler space is generated by<br />

the operator<br />

HR =<br />

�<br />

(H(�x)z − Pz(�x)t)d 3 �x. (4.14)<br />

This can be understood by analogy by looking at how the boost matrix<br />

�<br />

cosh α<br />

�<br />

sinh α<br />

sinh α cosh α<br />

works on a vector (z, t). In the infinitesimal case we get<br />

� � � � � � � � �<br />

z 1 α z z + αt z<br />

→<br />

=<br />

=<br />

t α 1 t αz + t t<br />

We have that<br />

�<br />

�<br />

t<br />

+ α<br />

z<br />

(4.15)<br />

�<br />

. (4.16)<br />

δϕ(z, t) = tδzϕ − zδtϕ. (4.17)<br />

Usually one considers the operator HR at t = 0,<br />

�<br />

HR = HM(�x)zd 3 �x. (4.18)<br />

In contrast with (4.13), this is not bounded from below. We would like to write this as<br />

with<br />

H1 =<br />

HR = H1 − H2, (4.19)<br />

�<br />

�<br />

z>0<br />

H(�x)zd 3 �x, (4.20)<br />

H2 = H(�x)|z|d<br />

z


4. Quantum Black Holes 30<br />

In the limit ξ → −∞ the last two terms do not contribute <strong>and</strong> we get plane waves.<br />

There is, however, a problem in that there is an asymptotic region at ξ → −∞ into<br />

which wave packets may disappear at η → ∞ or may come from at η → −∞. We<br />

need a boundary condition at ξ = −∞ in order to formulate scattering against the<br />

Rindler horizon. We do not know this boundary condition <strong>and</strong> would like to find it.<br />

The explanation for this weird situation is that there seems to be thermal (Hawking)<br />

radiation emerging from the region ξ = −∞. We will now consider the derivation<br />

<strong>of</strong> this phenomenon. Good references are Gerard ’t Ho<strong>of</strong>t’s article in [22] <strong>and</strong> the<br />

excellent book by Birrell <strong>and</strong> Davies [23].<br />

4.2.1 Derivation <strong>of</strong> the Hawking effect<br />

Let us consider only non-interacting scalar particles. Other cases are quite similar. For<br />

a scalar field Φ in Minkowski space we can make an expansion (see [24] for a nice<br />

introduction to quantum field theory)<br />

�<br />

Φ(�r, t) =<br />

d 3� k<br />

�<br />

2k0( � k)(2π) 3<br />

Things get nicer if we work in lightcone coordinates<br />

We define new annihilation operators a1 by<br />

�<br />

a( �k)e i�k�r−ik0t †<br />

+ a ( � −i<br />

k)e � �<br />

k�r+ik0t<br />

. (4.24)<br />

u = (t − z)/2, v = (t + z)/2, (4.25)<br />

k+ = k0 + k3, k− = k0 − k3. (4.26)<br />

a( � k) � k0 =: a1( ˜ k, k+) � k+. (4.27)<br />

It turns out to be convenient to Fourier transform a1 with respect to the logarithm <strong>of</strong><br />

k+ (see [22]). We will call the new annihilation operator obtained in this way a2, so<br />

a1( ˜ k, k+) � k+ = (2π) −1/2<br />

� ∞<br />

where µ 2 = ˜ k 2 + m 2 = k+k−.<br />

−∞<br />

dωa2( ˜ k, ω)e −iω ln(k+/µ) , (4.28)


From the Minkowski Hamiltonian<br />

we get the Rindler Hamiltonian<br />

4. Quantum Black Holes 31<br />

HM(r, t) = 1<br />

2 ((� ∂ϕ) 2 + (∂tϕ) 2 + m 2 ϕ 2 ) (4.29)<br />

HR =<br />

�<br />

zHM(�r, 0)d 3 �r, (4.30)<br />

<strong>and</strong> one can check by plugging a bit modified version <strong>of</strong> (4.24) <strong>and</strong> (4.28) into (4.30)<br />

that<br />

HR =<br />

�<br />

d 2˜<br />

� ∞<br />

k dωa<br />

−∞<br />

†<br />

2( ˜ k, ω)a2( ˜ k, ω). (4.31)<br />

So indeed, our claim that a2 is an annihilation operator was well grounded; it corre-<br />

sponds to a Rindler energy ω.<br />

Nevertheless a2 is not the annihilation operator we want to work with. As is well<br />

motivated in [22] we are interested in an operator aI( ˜ k, ω) that is defined as follows:<br />

<strong>and</strong> similarly<br />

aI( ˜ k, ω) √ 1 − e 2πω = a2( ˜ k, ω) + e −πω a †<br />

2(− ˜ k, ω), (4.32)<br />

aII( ˜ k, ω) √ 1 − e 2πω = a2( ˜ k, −ω) + e −πω a †<br />

2(− ˜ k, ω). (4.33)<br />

These transformations are called Bogolyubov transformations. One can check that the<br />

value <strong>of</strong> Φ in region I depends only on aI <strong>and</strong> its Hermitean conjugate, <strong>and</strong> similarly<br />

for aII. The above definitions are motivated by the fact that we can exp<strong>and</strong> Φ(�r, t) as<br />

� ∞<br />

d<br />

Φ(�r, t) = dω<br />

2˜ k<br />

�<br />

4π(2π) 3 ei˜ k˜r<br />

where<br />

0<br />

�<br />

K(−ω, µu, µv) � a2( ˜ k, ω) + e −πω a †<br />

2(− ˜ k, −ω) � +<br />

K ∗ (−ω, µu, µv) � a †<br />

2(− ˜ k, ω) + e −πω a2( ˜ k, −ω) ��<br />

, (4.34)<br />

� ∞<br />

K(ω, α, β) =<br />

0<br />

dx<br />

x xiω e −ixα−iβ/x<br />

(4.35)<br />

is an integration kernel; see [22] for some properties <strong>of</strong> K. From the commutation<br />

relations<br />

[aI( ˜ k, ω), a †<br />

I (˜ k ′ , ω ′ )] = δ 2 ( ˜ k − ˜ k ′ )δ(ω − ω ′ ) (4.36)


(<strong>and</strong> similarly for [aII, a †<br />

II ]) <strong>and</strong><br />

4. Quantum Black Holes 32<br />

[aI, aI] = [aII, aII] = [aI, aII] = [aII, aI] = 0 (4.37)<br />

we see that all observables in region II commute with all aI, a †<br />

I . Our conclusion is<br />

that not a2 <strong>and</strong> a †<br />

2, but aI <strong>and</strong> a †<br />

I are the proper annihilation <strong>and</strong> creation operators for<br />

region I (<strong>and</strong> aII <strong>and</strong> a †<br />

II for region II).<br />

Let us now study the vacuum Ω as defined by an observer in Minkowski space:<br />

a |Ω〉 = a1 |Ω〉 = a2 |Ω〉 = 0 , for all k, ω. (4.38)<br />

In our Rindler basis this turns, using the Bogolyubov transformations (4.32) <strong>and</strong> (4.33),<br />

into<br />

aI( ˜ k, ω) |Ω〉 − e −πω a †<br />

II (−˜ k, ω) |Ω〉 = 0, (4.39)<br />

aII( ˜ k, ω) |Ω〉 − e −πω a †<br />

I (−˜ k, ω) |Ω〉 = 0, = 0. (4.40)<br />

So we learn that the vacuum only consists <strong>of</strong> states with equal number operators nI =<br />

nII, since the above implies that when acting on the vacuum Ω we have<br />

nI := a †<br />

I (˜ k, ω)aI( ˜ kω) = e −πω a †<br />

Ia† II = e−πωa †<br />

IIa† I = a†<br />

IIaII =: nII. (4.41)<br />

We thus find by plugging<br />

into (4.39) that<br />

�<br />

n<br />

fn<br />

Ω = �<br />

fn |n, n〉 (4.42)<br />

√ n |n − 1, n〉 = e −πω �<br />

<strong>and</strong> we thus see that the coefficients fn satisfy<br />

n<br />

n<br />

√<br />

fn n + 1 |n, n + 1〉, (4.43)<br />

fn+1 = e −πω fn. (4.44)<br />

Consequently, we can write the Minkowski vacuum in the Rindler basis as<br />

|Ω〉 = � √<br />

1 − e−2πω ˜k,ω<br />

∞�<br />

e −πnω |n, n〉. (4.45)<br />

n=0


4. Quantum Black Holes 33<br />

The strange conclusion we can make now, is that the Rindler observer detects par-<br />

ticles radiating in all directions at some temperature TR! To see this we write down the<br />

probability that a Rindler observer in region I, while looking at the states Ω, observes<br />

nI particles with energy ω <strong>and</strong> transverse momentum ˜ k in region I. From (4.42) <strong>and</strong><br />

(4.45) we see that this is<br />

PnI<br />

= �<br />

nII<br />

We can write this as<br />

〈Ω|nI, nII〉〈nI, nII|Ω〉 = |fnI |2 = (1 − e −2πω ) −1 e −2πnIω . (4.46)<br />

e −β(E−F ) , (4.47)<br />

where E = nω is the energy, β = 1/T the inverse temperature, <strong>and</strong> F the free energy.<br />

The Rindler temperature can be read <strong>of</strong> to be TR = 1/2π.<br />

Now, for the distant observer the central region <strong>of</strong> the Kruskal space <strong>of</strong> a <strong>black</strong> hole<br />

is just Rindler space. All the particles he sees are ingoing particles. There seem to be<br />

no particles coming out (in the local Minkowski frame). We thus need the state |Ω〉 to<br />

describe these. So the reason why an observer in the Kruskal region I will see thermal<br />

radiation with a temperature T is that he thinks he sees the state |Ω〉. Finally, notice<br />

that when Rindler space is replaced by Kruskal space we must insert an extra factor<br />

<strong>of</strong> 4M in the unit <strong>of</strong> time. Therefore the temperature <strong>of</strong> the expected radiation from a<br />

<strong>black</strong> hole is<br />

which is the Hawking temperature [2].<br />

TH = 1/8πM, (4.48)<br />

4.3 Path integral approach<br />

There is also an elegant derivation <strong>of</strong> the Hawking temperature which makes use <strong>of</strong><br />

path integrals. The evaluation <strong>of</strong> the generating function W [J] for the Green functions<br />

<strong>of</strong> the quantum field theory (at J = 0) gives us the partition function, which can<br />

be compared to the partition function ln Z = −βF <strong>of</strong> a thermodynamic system at<br />

temperature T = β −1 . Let us now take a brief look at how this goes.


4. Quantum Black Holes 34<br />

As is well know, the generating function W [J] is a functional integral over the<br />

exponent <strong>of</strong> the action:<br />

where<br />

�<br />

W [J] =<br />

�<br />

S[φ, J] =<br />

Dφe iS[φ,J] , (4.49)<br />

d 4 x[L(φ, ∂µφ) + Jφ] (4.50)<br />

is the action; L is the Lagrangian density, <strong>and</strong> J is a source current. One could try to<br />

take the Einstein-Hilbert action <strong>and</strong> just try to do the integral. But this gives an infinite<br />

result. What one should do, instead, is to consider a finite region <strong>and</strong> add a boundary<br />

term to the action. The action looks then like<br />

S[g, 0] = 1<br />

�<br />

�<br />

√ D 1<br />

−gRd x +<br />

16π<br />

8π<br />

M<br />

∂M<br />

√ −hBd D−1 ξ (4.51)<br />

where hµν is the induced metric on the boundary ∂M <strong>of</strong> a D-dimensional spacetime,<br />

ξ µ are the coordinates on this boundary, <strong>and</strong> B is essentially the trace <strong>of</strong> the extrinsic<br />

curvature (the second fundamental form). We will not treat this in detail, but by doing<br />

this, <strong>and</strong> going to imaginary time, one can evaluate this action <strong>and</strong> find it to be<br />

for the Schwarzschild action, <strong>and</strong><br />

S[g(S), 0] = iπ<br />

M (4.52)<br />

κ<br />

S[g(KN), 0] = iπ<br />

κ (M − ΦHQ) (4.53)<br />

for the Kerr-Newman metrics (3.9). So how do we get our partition function? The<br />

dominant contribution to the path integral (4.49) comes from the fields (in our case<br />

metrics) that minimize the action (thus solutions <strong>of</strong> the equations <strong>of</strong> motion). So in the<br />

case <strong>of</strong> rotating charged <strong>black</strong> holes we find<br />

ln W [0] = ln Z � iS[g(KN), 0]. (4.54)<br />

On the other h<strong>and</strong> we know that the partition function <strong>of</strong> a gr<strong>and</strong> canonical ensemble at<br />

inverse temperature β with chemical potentials µi associated with conserved charges<br />

Ni is given (by definition) by<br />

ln Z = −β(E − T S − �<br />

µiNi). (4.55)<br />

i


4. Quantum Black Holes 35<br />

For the case we are considering β equals 2π/κ (for a motivation, see the end <strong>of</strong> this<br />

section) <strong>and</strong> we find combining equations (4.53), (4.54), (4.55) that<br />

1<br />

1<br />

M = T S +<br />

2 2 ΦHQ + ΩHJ, (4.56)<br />

where ΦH <strong>and</strong> ΩH are our ‘conserved charges’ at the horizon: respectively the co-<br />

rotating electrostatic potential, <strong>and</strong> the angular velocity. Comparing this with the fol-<br />

lowing equation for M, due to Smarr [25],<br />

1 κ<br />

M =<br />

2 8π AH + 1<br />

2 ΦHQ + ΩHJ, (4.57)<br />

<strong>and</strong> restoring all constants that we have not been writing we indeed find<br />

T =<br />

<strong>and</strong> also the Bekenstein-Hawking formula<br />

�c 3<br />

8πGNMkB<br />

(4.58)<br />

S = c3AH . (4.59)<br />

4GD�<br />

Let us give a concrete example — which is at the same time a shortcut to find the<br />

Hawking temperature — <strong>of</strong> how this imaginary time trick is being used in practice.<br />

We will take the Schwarzschild metric for simplicity.<br />

In the (T, R, θ, φ) Kruskal coordinates from chapter 2 the Schwarzschild metric<br />

takes the form<br />

with<br />

ds 2 = −<br />

32M 3<br />

r<br />

T 2 − X 2 =<br />

e −r/2GM (dT 2 − dX 2 ) + r 2 dΩ 2<br />

(4.60)<br />

�<br />

1 − r<br />

�<br />

e<br />

2M<br />

r/2M , (4.61)<br />

X + T<br />

X − T = et/2M , (4.62)<br />

(cf. equations (2.11) <strong>and</strong> (2.12)). From equation (4.61) we see that the singularity<br />

at r = 0 corresponds to the surface −T 2 + X 2 = −1. A smart way <strong>of</strong> avoiding this<br />

singularity is by defining a new coordinate ˜ T = iT , so that r = 0 is being avoided since<br />

it corresponds to ˜ T 2 + X 2 = −1, which has no real solutions. The metric becomes<br />

ds 2 =<br />

32M 3<br />

r<br />

e −r/2GM (d ˜ T 2 + dX 2 ) + r 2 dΩ 2<br />

(4.63)


with<br />

4. Quantum Black Holes 36<br />

˜T 2 + X 2 =<br />

� r<br />

2M<br />

�<br />

− 1 e r/2M , (4.64)<br />

X − i ˜ T<br />

X + i ˜ T = et/2M . (4.65)<br />

Note that on the contour where ˜ T <strong>and</strong> X are real, r is real <strong>and</strong> larger than 2M. As<br />

promised, we define an imaginary time coordinate τ by<br />

τ = it = i2M ln X − i ˜ T<br />

X + i ˜ T<br />

(4.66)<br />

The logarithm <strong>of</strong> a complex number can be easily found: it is not defined on the<br />

negative real axis, <strong>and</strong> every other point the complex number (a + ib) can be written as<br />

re iφ (with r > 0, −π < φ < π) so we see that ln(a + ib) = ln|a 2 + b 2 | + iArg(a + ib).<br />

Thus from 4.66 we have<br />

τ = i2M[ln(X − i ˜ T ) − ln(X + i ˜ T )] (4.67)<br />

= i2Mi(Arg(X − i ˜ T ) − Arg(X + i ˜ T )) = 4MArg(X + i ˜ T ) (4.68)<br />

This is an angular coordinate with period β = 4M ·2π = 8πM. The functional integral<br />

defining the generating function should therefore be over matter fields <strong>and</strong> metrics with<br />

this periodicity in τ. But this is just the partition function Z for a canonical ensemble<br />

<strong>of</strong> the fields at temperature T = β −1 = (8πM) −1 . For a Schwarzschild <strong>black</strong> hole<br />

the surface gravity is known to be κ = (4M) −1 , so that the temperature is indeed<br />

T = κ/2π in accordance with the above.


5. EXTREMAL BLACK HOLES IN STRING THEORY<br />

There is only one c<strong>and</strong>idate theory, named string theory, that can provide us with an<br />

answer to the question what the microstates <strong>of</strong> a <strong>black</strong> hole are. In this chapter we will<br />

examine the <strong>macroscopic</strong> <strong>and</strong> <strong>microscopic</strong> <strong>structure</strong> <strong>of</strong> <strong>black</strong> holes in string theory.<br />

Some good references for this chapter are [26, 27, 28, 29, 30, 31, 32].<br />

5.1 Perturbative microstates in string theory<br />

As we have seen in the previous chapters, <strong>black</strong> holes should be thought <strong>of</strong> as thermo-<br />

dynamical systems. Since we know that thermodynamics is only an approximation <strong>of</strong><br />

a more fundamental description in terms <strong>of</strong> microstates <strong>of</strong> the system, we would like<br />

to find out what these microstates <strong>of</strong> a <strong>black</strong> hole are. Roughly speaking, a system<br />

with an entropy S has e S microstates. As we have seen before, for the Schwarzschild<br />

<strong>black</strong> hole Sbh ∼ A ∼ M 2 M 2<br />

. The associated number <strong>of</strong> states thus grows like e . At<br />

this point it is interesting to think about whether this agrees or not with the number <strong>of</strong><br />

perturbative string states at mass level M, as this is what one would perhaps expect<br />

(or hope for?). One could object that associating perturbative string states with a <strong>black</strong><br />

hole is absurd. In the end, perturbative string states are obtained by quantizing a string<br />

in a flat background spacetime, so how can they be equivalent with a <strong>black</strong> hole? This<br />

is certainly a legitimate question. To jump ahead a bit, let us say something about this.<br />

As we will see, as the string coupling strength gS increases, the mass M <strong>of</strong> the pertur-<br />

bative state is unaltered both in the classical (independence <strong>of</strong> gS) as in the quantum<br />

world (no corrections due to supersymmetry). Now the gravitational field <strong>of</strong> the state<br />

is determined by GNM (in four dimensions) <strong>and</strong> since GN is proportional to g2 S , we<br />

find that as gS increases, the gravitational field increases. There is thus a back reaction<br />

on the perturbative state <strong>and</strong>, eventually, it may be described by a curved spacetime


5. Extremal Black Holes in String Theory 39<br />

with large curvature. Thus, at least in principle, it might be possible to associate cer-<br />

tain <strong>black</strong>-hole spacetimes with perturbative states. We will see that this is the case for<br />

extreme (BPS) <strong>black</strong> holes.<br />

Consider the open bosonic string, with mode expansion<br />

X µ = x µ + l 2 sp µ � 1<br />

τ + ils<br />

n<br />

n�=0<br />

αµ n cos(nσ). (5.1)<br />

Remember that in this expression the oscillator coefficients α µ n are creation operators<br />

for n ≤ 0 <strong>and</strong> annihilation operators for n ≥ 0. We will take units in which the string<br />

length scale ls = 1/ √ πT (where T is the string tension) is 1. Then one can easily<br />

show that the mass eigenvalues are given by the eigenvalues <strong>of</strong><br />

where<br />

1<br />

2 M 2 = N − 1, (5.2)<br />

N =<br />

∞�<br />

α i −nα i n, (5.3)<br />

n=1<br />

is the number operator (where there is an implied sum over i, the 24 transverse dimen-<br />

sions <strong>of</strong> the bosonic string). Our goal is to estimate the number dn <strong>of</strong> states at level n.<br />

It is convenient to define a generating function<br />

G(w) ≡ tr w N =<br />

∞�<br />

dnw n , (5.4)<br />

with |w| < 1. We will now examine the behavior <strong>of</strong> G(w) <strong>and</strong> use Cauchy’s equation<br />

to get an approximation <strong>of</strong> dn<br />

dn = 1<br />

�<br />

2πi C<br />

n=0<br />

G(w)<br />

dw. (5.5)<br />

wn+1 where C is a closed loop around w = 0. Plugging in the expression for the number<br />

operator N, we have (using the geometric series to get rid <strong>of</strong> the sum over i) that<br />

tr w N ∞�<br />

= tr w αi −nαi ∞�<br />

�<br />

1<br />

n =<br />

1 − wm �24 . (5.6)<br />

Let us try to simplify<br />

m=1<br />

f(w) ≡<br />

m=1<br />

∞�<br />

(1 − w m ) (5.7)<br />

m=1


5. Extremal Black Holes in String Theory 40<br />

by using (once again) the geometric series <strong>and</strong> a Taylor expansion for the logarithm:<br />

ln f(w) =<br />

∞�<br />

ln(1 − w m ) = −<br />

m=1<br />

∞�<br />

m,p=1<br />

wmp p =<br />

∞�<br />

p=1<br />

wp p(1 − wp . (5.8)<br />

)<br />

Close to w = 1 we can Taylor exp<strong>and</strong> w p as w p ∼ = 1 + p(w − 1) + . . .. We get<br />

<strong>and</strong> thus<br />

when w ∼ 1.<br />

From (5.5) we get that<br />

ln f(w) ∼ = −1<br />

1 − w<br />

∞�<br />

p=1<br />

p −2 = − π2<br />

6(1 − w)<br />

� � 2 4π<br />

G(w) ∼ exp<br />

1 − w<br />

dn ∝ e 4π√ n<br />

(5.9)<br />

(5.10)<br />

as n → ∞. (5.11)<br />

Therefore, from (5.2), the number <strong>of</strong> states <strong>of</strong> a free string increases as<br />

ρ(M) ∝ e √ 2πM . (5.12)<br />

That is, with an M in the exponent. Note that this is very different from the <strong>black</strong> hole<br />

requirement that the number <strong>of</strong> states increases as<br />

which has an M 2 in the exponent.<br />

ρbh(M) ∝ e 4πGN M 2<br />

, (5.13)<br />

5.2 Correspondence principle between strings <strong>and</strong> <strong>black</strong> holes<br />

In [4] Susskind speculates that all <strong>black</strong> hole states are in one to one correspondence<br />

with single string states. Susskind’s proposition is based on the following. By in-<br />

creasing the string coupling, the size <strong>of</strong> a highly excited string becomes less than its<br />

Schwarzschild radius. We should thus think <strong>of</strong> it as a <strong>black</strong> hole. Decreasing the<br />

coupling, on the other h<strong>and</strong>, generically makes the size <strong>of</strong> a <strong>black</strong> hole smaller, <strong>and</strong><br />

eventually even smaller than the string scale. In such a case, the metric becomes ill-<br />

defined <strong>and</strong> we can no longer interpret the <strong>black</strong> hole as being a <strong>black</strong> hole. Susskind


5. Extremal Black Holes in String Theory 41<br />

suggests that we should describe this configuration in terms <strong>of</strong> some string state. At<br />

weak coupling the <strong>black</strong> hole can be thought <strong>of</strong> as a small number <strong>of</strong> (or to leading<br />

order just one) highly excited strings. There seems to be a problem, however. As we<br />

have shown in section 5.1, the entropy <strong>of</strong> a free string is proportional to the mass <strong>of</strong><br />

the string state, whereas the Bekenstein-Hawking entropy is proportional to the square<br />

<strong>of</strong> the mass <strong>of</strong> the <strong>black</strong> hole. As a resolution to this discrepancy, Susskind’s suggests<br />

that in the above M versus M 2 comparison the effect <strong>of</strong> a large gravitational redshift<br />

is not taken into account.<br />

In [5] Horowitz <strong>and</strong> Polchinski give a more precise, although still somewhat qual-<br />

itative, formulation <strong>of</strong> the relation between strings <strong>and</strong> <strong>black</strong> holes. Based on this fact<br />

that for <strong>black</strong> holes in string theory generically the Schwarzschild radius in string units<br />

decreases when the string coupling is reduced, they formulate a correspondence prin-<br />

ciple, which states that (i) when the size <strong>of</strong> the horizon drops below the size <strong>of</strong> a string,<br />

the typical <strong>black</strong> hole states becomes a typical state <strong>of</strong> strings (<strong>and</strong> so-called D-branes)<br />

with the same charges, <strong>and</strong> (ii) the mass does not change abruptly during the transi-<br />

tion. Their approach does not give right numerical coefficients, but it does give the<br />

correct dependence on mass <strong>and</strong> charge for a large variety <strong>of</strong> cases. For the exact right<br />

numerical coefficients we have to introduce some more concepts, which we will do in<br />

the following sections.<br />

5.3 Extremal <strong>black</strong> holes <strong>and</strong> extended supersymmetry<br />

Let us next review the supersymmetry algebra <strong>and</strong> its representations. We can view<br />

extremal Reissner-Nordström <strong>black</strong> holes as ‘supersymmetric solitons’ because they<br />

behave very much like BPS objects where M ≥ Q is the BPS bound. The reason why<br />

we can associate BPS states with extremal <strong>black</strong> holes is that they have the crucial<br />

property that they cannot receive quantum corrections, which is similar to the fact that<br />

extremal <strong>black</strong> holes have vanishing temperature <strong>and</strong> thus don’t radiate. In this section<br />

we will clarify all this.<br />

Supersymmetry is a symmetry between fermions <strong>and</strong> bosons <strong>and</strong> supersymmetric


5. Extremal Black Holes in String Theory 42<br />

theories are theories with conserved spinorial currents. For each such current we get<br />

four conserved charges. We can organize these charges into a Weyl spinor, or, in gen-<br />

eral, into N Weyl spinors Q A α if we have N spinorial currents (<strong>and</strong> thus 4N conserved<br />

charges). Here A = 1, ..., N counts the supersymmetries <strong>and</strong> α = 1, 2 is a Weyl spinor<br />

index. A theorem <strong>of</strong> Haag, Lopuzanski <strong>and</strong> Sohnius states that the most general form<br />

<strong>of</strong> the supersymmetry algebra 1 (in four spacetime dimensions), is given by<br />

{Q A α, ¯ Q B ˙ β } = 2δ A Bσ µ<br />

α ˙ β Pµ, (5.14)<br />

{Q A α, Q B β } = 2ɛαβZ AB . (5.15)<br />

The Q bars are the hermitean conjugates <strong>and</strong> have opposite chirality (this is indicated<br />

by the dot), so ¯ Q B ˙ β = (Q B β )† ; the 2 × 2 matrices σ µ with µ = 1, 2, 3 are the Pauli<br />

matrices, <strong>and</strong> σ 0 is the identity matrix; ɛαβ is antisymmetric with ɛ12 = +1; the matrix<br />

Z AB is also antisymmetric <strong>and</strong> the operators in it commute with all the operators in<br />

the super Poincaré algebra which explains why they are called central charges; by<br />

definition we set Z 12 ≡ Z, which we may choose without loss <strong>of</strong> generality to be non-<br />

negative. Let us for concreteness take N = 2 <strong>and</strong> start with massive representations,<br />

M 2 > 0. We may work in the rest frame Pµ = (M,�0). By forming two appropriate<br />

linear combinations aα = Q 1 α + ɛαβQ 2 β <strong>and</strong> bα = Q 1 α − ɛαβQ 2 β<br />

can bring the algebra to the form<br />

{aα, a †<br />

β } = 4(M + Z)δαβ,<br />

{bα, b †<br />

β } = 4(M − Z)δαβ,<br />

<strong>of</strong> the generators, we<br />

{aα, b †<br />

β } = {aα, bβ} = 0. (5.16)<br />

Note that (up to a non-relevant normalization factor) these are precisely the anticom-<br />

mutation relations satisfied by two independent sets <strong>of</strong> fermion annihilation <strong>and</strong> cre-<br />

ation operators. Starting with some state |ψ〉 that is annihilated by the annihilation<br />

operators aα <strong>and</strong> bα we can construct a total <strong>of</strong> 16 states using the creation operators<br />

a † α <strong>and</strong> b † α. Since quantum mechanics is a unitary theory we require the absence <strong>of</strong><br />

1 This is is an extension <strong>of</strong> the usual Poincaré algebra (the Lie algebra <strong>of</strong> the 10-dimensional Poincaré<br />

Lie group <strong>of</strong> isometries <strong>of</strong> Minkowski space).


5. Extremal Black Holes in String Theory 43<br />

negative norm states, i.e., we have 〈φ|{aα, a † α}|φ〉 ≥ 0 for any state |φ〉 (<strong>and</strong> similarly<br />

for bα). This implies that the mass is bounded by the central charge. This is called the<br />

BPS-bound:<br />

M ≥ |Z|. (5.17)<br />

In the case M = Z we have to be careful, since in this case the representation contains<br />

null states. These have to be divided out in order to get unitary representations, which<br />

corresponds to |ψ〉 being annihilated by one set <strong>of</strong> creation operators (b † α if Z > 0).<br />

Such ‘BPS’-states are thus constructed using only the a †<br />

β<br />

creation operators. As a<br />

consequence they are invariant under half <strong>of</strong> the supersymmetry algebra. They are<br />

so-called ‘short’ massive representations <strong>of</strong> the supersymmetry algebra.<br />

An interesting feature <strong>of</strong> BPS states is that they cannot receive quantum correc-<br />

tions, assuming that the full theory is supersymmetric. This is the case because the<br />

relation M = Z is a consequence <strong>of</strong> the supersymmetry algebra. The analogy with ex-<br />

tremal Reissner-Nordström <strong>black</strong> holes, which have exactly vanishing Hawking tem-<br />

perature, is clear.<br />

The extreme Reissner-Nordström <strong>black</strong> hole is sometimes called a supersymmetric<br />

soliton. In the remainder <strong>of</strong> this section we will explain this terminology.<br />

First let us say what we mean by a soliton. A soliton is a stationary, regular, <strong>and</strong><br />

stable finite energy solution to the equations <strong>of</strong> motion. Some <strong>of</strong> these features are<br />

immediately clear for the extreme Reissner-Nordström <strong>black</strong> hole: it is static <strong>and</strong> thus<br />

in particular it is stationary; it has a finite mass M <strong>and</strong> thus finite energy; it is regular in<br />

the sense that it has no naked singularity. It is also stable, at least when considered as<br />

a solution <strong>of</strong> N = 2 supergravity into which it can be embedded, as we will see below.<br />

The next question is: what do we mean by a ‘supersymmetric’ soliton? A solution is<br />

called supersymmetric if it is invariant under a rigid transformation, which can be seen<br />

as an analogue <strong>of</strong> an isometry. By ‘being invariant’ we mean here<br />

δɛ(x)Φ |Φ0= 0. (5.18)<br />

That is to say, a field configuration Φ0 is supersymmetric if there exists a choice ɛ(x) <strong>of</strong><br />

the supersymmetry transformation parameters such that the configuration is invariant.


5. Extremal Black Holes in String Theory 44<br />

The notation should speak for itself: one should perform a supersymmetry variation<br />

<strong>of</strong> all the fields Φ <strong>and</strong> then evaluate it at Φ0; this should be zero. It is instructive<br />

to note that the transformation parameters ɛ(x) are fermionic analogues <strong>of</strong> Killing<br />

vectors. (They are called Killing spinors <strong>and</strong> equation (5.18) is the Killing spinor<br />

equation.) Killing vectors are the generators <strong>of</strong> spacetime symmetries. The trivial<br />

vacuum (Minkowski space) has ten Killing vectors. A generic spacetime has none,<br />

whereas more symmetric spacetimes can have some. For example the RN <strong>black</strong> hole<br />

has 4 Killing vectors (1 for time, 3 for rotation invariance). Similarly, we have ‘Killing<br />

spinors’ ɛ(x) which are supersymmetry transformation parameters.<br />

As mentioned above, the extreme RN <strong>black</strong> hole can be embedded into N = 2<br />

supergravity. This can be done by adding two gravitini ψ A µ . The extreme RN <strong>black</strong><br />

hole is not only a solution <strong>of</strong> Einstein-Maxwell theory, but also <strong>of</strong> the extended theory,<br />

with ψ A µ = 0. It also possess Killing spinors <strong>and</strong> is thus a supersymmetric solution in<br />

the above sense. Actually the only condition comes from the gravitino variation<br />

δɛψµA = ∇µɛA − 1 −<br />

Fab 4 γaγ b γµεABɛ B = 0, (5.19)<br />

where we suppressed the spinor indices <strong>and</strong> introduced some new notation. We use<br />

four-dimensional Majorana spinors, but project onto one chirality. This is encoded<br />

it the placement <strong>of</strong> the index A = 1, 2: letting γ5 work on ɛ A gives ɛ A , whereas<br />

for a lower index A we pick up a minus sign. Half <strong>of</strong> the components <strong>of</strong> ɛA, ɛ A are<br />

independentFµν is the graviphoton field strength <strong>and</strong> the minus superscript indicates<br />

its antiselfdual part:<br />

F ± µν = 1<br />

2 (Fµν ± i ⋆ Fµν). (5.20)<br />

One can now check (we will not do this) that the extreme RN <strong>black</strong> hole has Killing<br />

spinors which can be split in a part h(�x) which is completely fixed in terms <strong>of</strong> the<br />

coefficients in the metric, <strong>and</strong> a part at infinity<br />

where the part at infinity satisfies an additional equation<br />

ɛA(h(�x)) = h(�x)ɛA(∞), (5.21)<br />

ɛA(∞) + iγ 0 Z<br />

|Z| ɛABɛ B (∞) = 0. (5.22)


5. Extremal Black Holes in String Theory 45<br />

Consequently, half <strong>of</strong> the parameters are fixed in terms <strong>of</strong> others, which means that<br />

we half <strong>of</strong> the maximum, which is 8/2=4, Killing spinors. Furthermore, it can be<br />

shown that the extreme RN <strong>black</strong> hole is part <strong>of</strong> a hypermultiplet. The factor Z/|Z|<br />

is the central charge. In local supersymmetric theories central charge transformations<br />

are local U(1) transformation. The corresponding gauge field is a graviphoton. The<br />

central charge is a linear combination <strong>of</strong> the electric <strong>and</strong> magnetic charge <strong>of</strong> this U(1):<br />

Z = 1<br />

� �<br />

− 2F = p − iq. So |Z| = p2 + q2 = Q = M gives us our relation between<br />

4π<br />

the extreme RN <strong>black</strong> hole <strong>and</strong> the supersymmetric BPS limit. The extreme RN <strong>black</strong><br />

hole has thus all the features we expect from a BPS state:<br />

• It is invariant under half <strong>of</strong> the supersymmetry transformations.<br />

• It sits in a short multiplet.<br />

• It saturates the supersymmetric mass bound.<br />

The strange no-force property that admits that a number <strong>of</strong> extremal <strong>black</strong> holes can<br />

sit next to each other without feeling each others presence (see chapter 2) can now be<br />

understood as a consequence <strong>of</strong> the additional supersymmetry.<br />

Our next goal will be to try to find a <strong>microscopic</strong> description <strong>of</strong> the Reissner-<br />

Nordström <strong>black</strong> hole. It turns out that the simplest example one can consider in<br />

string theory is its five dimensional analogue. We will need to consider low energy<br />

effective field theory describing string theory. This is a generalization <strong>of</strong> Einstein-<br />

Maxwell theory. The number <strong>of</strong> spacetime dimensions involved is ten. There are also<br />

more fields involved: they are precisely the massless modes that arise in (perturbative)<br />

string theory. Higher massive modes decouple <strong>and</strong> do not appear in the low energy<br />

effective theory. We will now study type II supergravity which is the field theory that<br />

describes such massless modes. As the name already implies, supergravity is the union<br />

<strong>of</strong> general relativity (a theory <strong>of</strong> gravity) <strong>and</strong> supersymmetry.


5. Extremal Black Holes in String Theory 46<br />

5.4 Type II supergravity<br />

Supersymmetry is a symmetry between bosons <strong>and</strong> fermions <strong>and</strong> the massless states<br />

in (super)string theory include both bosons <strong>and</strong> fermions. However, since <strong>black</strong> holes<br />

are solutions <strong>of</strong> the classical bosonic field equations, we will only be interested in the<br />

bosonic degrees <strong>of</strong> freedom. Closed string theories have massless (bosonic) modes<br />

associated with a symmetric, traceless, rank-2 tensor (the graviton field Gµν), an an-<br />

tisymmetric rank-2 tensor (the Kalb-Ramond field Bµν), <strong>and</strong> the dilaton 2 field φ. In<br />

type II string theories, these states arise in the NS-NS sector. Before we continue, let<br />

us remind ourselves <strong>of</strong> some string theory basics.<br />

In string theory a two-dimensional non-linear sigma-model with action<br />

�<br />

SW S = 1<br />

4πα ′<br />

d<br />

Σ<br />

2 σ √ −hh αβ (σ)∂αX µ ∂βX ν Gµν(X) (5.23)<br />

is being used to model the motion <strong>of</strong> a string in a curved spacetime background with<br />

metric Gµν(X). The integration is over the world-sheet Σ. There are two coordinates<br />

(σ 0 , σ 1 ) = σ <strong>and</strong> hαβ(σ) is the world-sheet metric. The coordinates <strong>of</strong> the string in<br />

spacetime are X µ (σ). The parameter α ′ is related to the string tension T <strong>and</strong> the string<br />

length scale lS via α ′ = 1<br />

2πT<br />

1 = 2l2 S <strong>and</strong> is the only independent dimensionful parameter<br />

in string theory. One can impose different conditions on the boundaries <strong>of</strong> the-world<br />

sheet. One <strong>of</strong> them is<br />

∂1X µ |∂Σ= 0. (5.24)<br />

This is the Neumann boundary condition <strong>and</strong> it describes open strings. One can also<br />

fix the endpoints <strong>of</strong> the string. This is the Dirichlet boundary condition<br />

∂0X µ |∂Σ= 0. (5.25)<br />

Since the momentum at the end <strong>of</strong> the string is not conserved, Dirichlet boundary con-<br />

ditions require the string to be coupled to another dynamical object. Such objects are<br />

called D-branes. They are non-perturbative objects since they have a tension propor-<br />

tional to 1/gS. D-branes are very important in underst<strong>and</strong>ing the <strong>microscopic</strong> origin<br />

2 The expectation value <strong>of</strong> the dilaton fixes the string coupling constant via gS = 〈e φ 〉.


5. Extremal Black Holes in String Theory 47<br />

<strong>of</strong> <strong>black</strong> hole entropy. At this point let us just state that one can form so-called Dp-<br />

branes by imposing Neumann conditions along time <strong>and</strong> p directions, <strong>and</strong> Dirichlet<br />

conditions among the remaining D − p − 1 directions. (Here D is the dimension <strong>of</strong> the<br />

spacetime.) We will come back to this in the next section.<br />

Type II string theory consists <strong>of</strong> closed oriented strings <strong>and</strong> the world-sheet action<br />

is extended to a (1,1) supersymmetric action by adding world-sheet fermions ψ µ (σ).<br />

For these fermions one can choose the boundary conditions independently to be either<br />

periodic (Ramond) or anti-periodic (Neveu-Schwarz). We thus have four scenarios:<br />

NS-NS, NS-R, R-NS, R-R. The states in the NS-NS <strong>and</strong> R-R sectors are bosonic, the<br />

ones in the R-NS <strong>and</strong> NS-R sectors are fermionic.<br />

As stated above, the massless spectrum <strong>of</strong> the NS-NS sector for both type IIA<br />

<strong>and</strong> IIB string theory consists <strong>of</strong> Gµν, Bµν, φ. They are constructed using the Neveu-<br />

Schwarz (half-integer moded) world-sheet fermion creation operators for both left-<br />

<strong>and</strong> right-movers. In the R-R sector we also have massless bosonic states. They are<br />

constructed using the Ramond (integer moded) world-sheet fermion creation operators<br />

for both left- <strong>and</strong> right-movers. For the R-R sector we have to distinguish between the<br />

IIA <strong>and</strong> IIB theories. Let us look at IIA for the moment. We can do this without loss<br />

<strong>of</strong> generality, since one can relate the (chiral) IIB to the (non-chiral) IIA theory via<br />

T-duality. The R-R states may be represented by form fields C(1) <strong>and</strong> C(3) (for type IIB<br />

these are C(0), C(2) <strong>and</strong> C(4)), where<br />

C(n) ≡ 1<br />

n! Cµ1µ2...µndx µ1 ∧ dx µ2 ∧ ... ∧ dx µn . (5.26)<br />

We could as well have represented these fields by C(7) <strong>and</strong> C(5). They are related by<br />

the 8-dimensional epsilon tensor. It is a known fact [26] that the condition for the tree-<br />

level Weyl invariance <strong>of</strong> the string world-sheet action is preserved in the (quantum)<br />

string theory boils down to the vanishing <strong>of</strong> the associated (renormalization-group)<br />

beta functions for these fields. As a result, one gets the following effective action<br />

(which is a good approximation in the low energy limit α ′ → 0) in the so-called string


frame:<br />

SIIA =<br />

5. Extremal Black Holes in String Theory 48<br />

1<br />

2κ 2 10<br />

�<br />

+<br />

� �<br />

d 10 x(−G) 1/2 e −2φ R(G)<br />

[e −2φ (4dφ ∧ ∗ dφ − 1<br />

2 H(3) ∧ ∗ H(3)) − 1<br />

2 F(2) ∧ ∗ F(2)<br />

− 1<br />

2 ˜ F(4) ∧ ∗ F(4) − 1<br />

2 B(2) ∧ F(4) ∧ F(4)]<br />

�<br />

. (5.27)<br />

We see that all terms from the NS-NS sector are coupled to the dilaton, while terms<br />

from the R-R sector are not. In this expression G is the determinant <strong>of</strong> Gµν (the<br />

metric appearing in the non-linear sigma model (5.23)). κ10 has the dimensions <strong>of</strong><br />

a ten-dimensional gravitational coupling, but can not be be directly identified with the<br />

physical gravitational coupling. The reason for this is simply that a rescaling <strong>of</strong> κ<br />

can be compensated (as can be seen from equation (5.27)) by a shift <strong>of</strong> the dilaton’s<br />

vacuum expectation value. It is clear what the relation between κ10, α ′ <strong>and</strong> gS should<br />

be, namely κ10 = (α ′ ) 2 gS · const. In fact,<br />

2κ 2 10 = (2π) 7 (α ′ ) 4 g 2 S. (5.28)<br />

Furthermore we have the dilaton 1-form dφ, which is simply ∂µφdx µ . The field-<br />

strength forms are related to the potentials by<br />

F(2) = dC(1), F(4) = dC(3) <strong>and</strong> H(3) = dB(2)<br />

where B(2) is the 2-form associated with Bµν, <strong>and</strong><br />

(5.29)<br />

˜F(4) = F(4) + C(1) ∧ H(3). (5.30)<br />

Finally, Hodge duals in D dimensions (so in our case, D = 10) are defined by<br />

∗ C(p) = (−G)1/2<br />

p!(D − p)! ɛν1ν2...νD−pµ1µ2...µpC µ1µ2...µp dx ν1 ∧ dx ν2 ∧ ... ∧ dx νD−p . (5.31)<br />

It is easy to transform this ‘string frame’ action into the ‘Einstein frame’ action, in<br />

which there is no dilaton factor in the curvature term, by plugging in the field redefini-<br />

tion<br />

Gµν = e φ/2 gµν. (5.32)


This gives<br />

SIIA =<br />

1<br />

2κ 2 10<br />

− 1<br />

2<br />

5. Extremal Black Holes in String Theory 49<br />

�<br />

� �<br />

d 10 x(−g) 1/2 R(g)<br />

[dφ ∧ ∗ −φ 1<br />

dφ + e<br />

2 H(3) ∧ ∗ H(3) + e 3φ/2 F(2) ∧ ∗ F(2)<br />

+e φ/2 ˜ F(4) ∧ ∗ ˜ F(4) + B(2) ∧ F(4) ∧ F(4)]<br />

�<br />

(5.33)<br />

for the effective bosonic action in the Einstein frame. Another possibility would be to<br />

absorb the dilaton vacuum expectation value in the prefactor <strong>of</strong> the action, see [30].<br />

We see that type II string theory effective action involves various field strengths. In<br />

the same fashion as an electron is coupled by its charge to the Maxwell gauge potential<br />

Aµ, one would expect that there should be some objects in the underlying string theory<br />

that couple to the associated gauge form fields. Indeed, such extended objects exist:<br />

they are called p-branes, where p st<strong>and</strong>s for the number <strong>of</strong> spatial dimensions. In fact,<br />

p-branes are higher dimensional analogues <strong>of</strong> the extremal Reissner-Nordström <strong>black</strong><br />

hole. Before we come back to this, let us study our form fields a bit more.<br />

5.5 Form fields <strong>and</strong> D-branes<br />

Starting with the gauge field Aµ, which is coupled naturally to the world line X µ (τ)<br />

<strong>of</strong> some charged particle by a term<br />

�<br />

Aµ<br />

dX µ<br />

dτ (5.34)<br />

dτ<br />

in the action, <strong>and</strong> whose vector potential 1-form Aµdx µ ≡ A(1) transforms as A(1) →<br />

A(1) + dΛ0 (where Λ(0) is a function) we get the field strength 2-form F(2) = dA(1).<br />

The Maxwell equations in four dimensions are<br />

dF(2) = 0 <strong>and</strong> d ∗ F(2) = ∗ J(1), (5.35)<br />

where J(1) = Jµdx µ is the current one form. Since we are in four dimensions, ∗J(1) is<br />

a 3-form <strong>and</strong> we can find the electric charge by integrating over some spatial volume<br />

V3 which is enclosed by some surface S2. Using Gauss’ theorem we get<br />

� �<br />

∗<br />

Q = J(1) =<br />

V 3<br />

S 2<br />

∗ F(2). (5.36)


5. Extremal Black Holes in String Theory 50<br />

Now let us look at an analogous thing in string theory.<br />

Analogously to the gauge field/world line coupling above, one has the coupling <strong>of</strong><br />

the Kalb-Ramond field to the string world-sheet by<br />

�<br />

S ∼<br />

Bµνɛ αβ ∂αX µ ∂βX ν d 2 σ. (5.37)<br />

We have dH(3) = ddB(2) = 0. The electric charge Q1 <strong>of</strong> the string in 10 dimensions<br />

is given by<br />

Q1 =<br />

�<br />

S 7<br />

∗H(3). (5.38)<br />

This can be generalized for higher dimensional charges. Obviously, the generalization<br />

<strong>of</strong> (5.34) <strong>and</strong> (5.37) for a general antisymmteric (p + 1) dimensional tensor gauge field<br />

is a term in the action <strong>of</strong> the form<br />

�<br />

Cµ1µ2...µp+1ɛ α0α1...αp ∂α0X µ1 ∂α1X µ2 ...∂αpX µp+1 d p+1 σ. (5.39)<br />

So what this describes, is the coupling <strong>of</strong> the C(p+1) gauge field to the (p + 1) dimen-<br />

sional world volume <strong>of</strong> some extended object. This object has p spatial dimensions <strong>and</strong><br />

is called a p-brane. The electric charge <strong>of</strong> a p-brane that is enclosed by a hypersurface<br />

S8−p is given by<br />

� 8−p<br />

∗<br />

Qp = F(p+2), (5.40)<br />

S<br />

where F(p+2) = dC(p+1) is the field strength, which is invariant under gauge transfor-<br />

mations C(p+1) → C(p+1) + dΛ(p).<br />

In the effective action for type II superstrings (5.27) we see both the field strengths<br />

H3 <strong>and</strong> Fn appear. There is quite a difference between these. The H3 comes from<br />

the NS-NS form field B(2) which is coupled electrically to the string world sheet or its<br />

excitations. The Fn, with n even, are derived from the R-R fields C(n−1) <strong>and</strong> do not<br />

couple electrically. This suggests that R-R field strengths F(n) are naturally associated<br />

with branes having p = n − 2 dimensions.<br />

Since perturbative string states are neutral with respect to the charges associated<br />

with the R-R fields, we have to add non-perturbative dynamical objects that do have R-<br />

R charges to type II string theory. (By R-R charges we mean electric charges associated


5. Extremal Black Holes in String Theory 51<br />

with the R-R gauge form fields.) These objects are called Dp-branes. The D st<strong>and</strong>s<br />

for Dirichlet: there has also to be an open string sector where open strings end on<br />

p-dimensional planes [7]. Thus, the open string world sheet has Dirichlet boundary<br />

conditions in the directions perpendicular to the branes. These Dp-branes do couple<br />

electrically to the associated R-R fields. (Just as the fundamental string is coupled to<br />

the NS-NS field Bµν.) It is precisely this enlargement <strong>of</strong> string theory that lead to the<br />

underst<strong>and</strong>ing <strong>of</strong> <strong>black</strong> hole entropy in terms <strong>of</strong> associated microstates.<br />

But before we construct explicit <strong>black</strong> hole solutions <strong>of</strong> the type II supergravity<br />

field equations (whose entropy we shall be able to explain in terms <strong>of</strong> Dp-branes) let<br />

us briefly discuss dualities in string theory <strong>and</strong> give an example to get to know some<br />

<strong>of</strong> the techniques for doing <strong>microscopic</strong> entropy calculations.<br />

5.6 Dualities<br />

By a duality we mean two different ways to look at the same physics. The first duality<br />

symmetry we will consider, T-duality, is a symmetry that relates compactifications on<br />

a manifold <strong>of</strong> (large) volume V to ones with (small) volume 1/V . Obviously, the<br />

simplest case is compactification on a circle with radius R. It turns out that type IIA<br />

compactified on a circle with radius R is equivalent to type IIB compactified on a circle<br />

with radius 1/R, <strong>and</strong> likewise for heterotic E8 × E8 <strong>and</strong> heterotic SO(32).<br />

Next we consider S-duality. This is a (conjectured) non-perturbative duality. (Non-<br />

perturbative because it acts on the dilaton as gS → 1/gS.) S-duality is a weak ↔ strong<br />

duality: it relates the weak coupling regime <strong>of</strong> one theory to the strong coupling regime<br />

<strong>of</strong> another one. In particular, IIB theory is self-dual under S-duality <strong>and</strong> Heterotic<br />

SO(32) is S-dual to Type I theory. Type IIB string theory is believed to have an exact<br />

non-perturbative SL(2, Z) symmetry. We will have more to say about S-duality in<br />

chapter 7.


5. Extremal Black Holes in String Theory 52<br />

5.7 Counting <strong>of</strong> 1/2 BPS heterotic states on T 4 × T 2<br />

In this section we will count 1/2 BPS states as example to illustrate some st<strong>and</strong>ard<br />

counting techniques. We will consider heterotic string theory on T 4 × T 2 with the<br />

following coordinates:<br />

T 4<br />

����<br />

6789<br />

× ˜ S 1<br />

����<br />

5<br />

× S 1<br />

����<br />

4<br />

For convenience we will set α ′ = 1 in this section.<br />

× M 4<br />

����<br />

0123<br />

. (5.41)<br />

We consider a string around the ”5” direction, winding it ω times <strong>and</strong> with momen-<br />

tum n. The question is how many such states there are, subject to the usual Virasoro<br />

constraints that the left moving Hamiltonian is<br />

<strong>and</strong> the other constraint is<br />

where<br />

2 M<br />

−<br />

2 M<br />

−<br />

4 + NR + p2R 2<br />

4 + NL − 1 + p2L 2<br />

pL,R =<br />

= 0, (5.42)<br />

= 0, (5.43)<br />

� � �<br />

1 n<br />

∓ wR . (5.44)<br />

2 R<br />

We look at the right moving ground state, so NR = 0 an thus, from (5.42) we get that<br />

M = √ 2pR, (5.45)<br />

which looks like a BPS bound. Furthermore, the right moving superstring has a 16<br />

dimensional representation (8v for the NS sector, <strong>and</strong> 8s for the R sector) <strong>and</strong> this is<br />

thus a short representations (a long one would be 256 dimensional). From these two<br />

facts we see that this is BPS state. The question we want to consider is: how many<br />

such BPS states are there?<br />

The second equation says we can have arbitrary left-moving oscillations. The left-<br />

moving string is a 26 dimensional bosonic string. We choose one direction for the<br />

string to carry momentum <strong>and</strong> one for time, so we have 24 left moving oscillator.<br />

We have spacetime bosons X i , <strong>and</strong> internal bosons Y I which parametrize the torus


5. Extremal Black Holes in String Theory 53<br />

<strong>of</strong> E8 × E8. Let us denote the oscillators <strong>of</strong> X i as α i n <strong>and</strong> <strong>of</strong> Y I as β I n. Then the<br />

left-moving oscillations can be written as<br />

NL = �<br />

n>0<br />

(nα i −nα i n + nβ I −nβ I n), (5.46)<br />

where i = 1, ..., 8 because there are 8 transverse dimensions in 10 dimensions <strong>and</strong><br />

I = 1, ..., 16 because the rank <strong>of</strong> E8 × E8 is 16. To take care <strong>of</strong> the zero-point energy<br />

we introduce N = NL − 1 <strong>and</strong> from (5.43) we get<br />

N = NL − 1 = p2 R − p2 L<br />

2<br />

= nω. (5.47)<br />

We are interested in Ω(N), the number <strong>of</strong> states. Basically this is the harmonic oscil-<br />

lator problem with 24 oscillators. This Ω(N) appears in the partition function on the<br />

world-sheet,<br />

Z(β) = Tre −βN = � Ω(N)e −βN , (5.48)<br />

<strong>and</strong> we can us this to find Ω(N) by doing an inverse Laplace transform<br />

Ω(N) = 1<br />

�<br />

2πi<br />

e βN Z(β)dβ. (5.49)<br />

So our problem is reduced to finding this inverse Laplace transform. We would like<br />

to know the asymptotics <strong>of</strong> Ω(N) at large N, or in other words, at large temperature.<br />

Plugging in the expansions for N into (5.48), we get<br />

Z(β) = Tre −β(P (nα i −n αi n+nβ I −n βI n)−1 . (5.50)<br />

By introducing q = e −β <strong>and</strong> using the geometric series for 1 + q n + q 2n + ... we get<br />

Z(β) =<br />

1<br />

q �<br />

n (1 − qn . (5.51)<br />

) 24<br />

The trick is to use the fact that this is a modular form <strong>of</strong> weight 12 (it also appears in<br />

the one loop partition function <strong>of</strong> the bosonic string), namely<br />

Z(β) =<br />

� �12 β<br />

Z<br />

2π<br />

� 4π 2<br />

β<br />

�<br />

. (5.52)


5. Extremal Black Holes in String Theory 54<br />

We can relate high temperature with low temperature partition function by this equa-<br />

tion. For convenience, let’s introduce<br />

So at high temperature we get<br />

<strong>and</strong> thus<br />

˜β ≡ 4π2<br />

β , ˜q ≡ e− ˜ β<br />

(5.53)<br />

β → 0 ⇒ ˜ β → ∞ ⇒ ˜q → 1. (5.54)<br />

Z( ˜ β) =<br />

1<br />

˜q � (1 − ˜q n 1<br />

�<br />

) 24 ˜q<br />

= e 4π2<br />

β (5.55)<br />

for ˜q → 1 The above implies the following behavior at large N, or, at high temperature:<br />

Ω(N) = 1<br />

�<br />

2πi<br />

e βN e 4π2<br />

� �12 β<br />

β dβ<br />

2π<br />

(5.56)<br />

In our case we have a central charge c = 24. The more general expression is<br />

Ω(N) = 1<br />

�<br />

2πi<br />

e βN e 4π2<br />

� �12 c β<br />

β 24 dβ<br />

2π<br />

(5.57)<br />

We now want to do a saddle point approximation, so let us spend a few words on<br />

this. Basically, we want to find where the exponent has its peak. Let us do this a bit<br />

more general: suppose (<strong>and</strong> this is indeed what we want) that we want to calculate an<br />

integral<br />

� ∞<br />

A = dx e −ξf(x)<br />

−∞<br />

(5.58)<br />

in the limit <strong>of</strong> large ξ. Assuming f(x) has a global minimum at some x0 we can Taylor<br />

exp<strong>and</strong> f(x) <strong>and</strong> get to quadratic order<br />

f(x) ≈ f(x0) + 1<br />

2 f ′′ (x0)(x − x0) 2 . (5.59)<br />

Assuming that our global minimum is sufficiently separated from other local minima<br />

<strong>and</strong> whose value is sufficiently smaller than the value <strong>of</strong> those, the dominant contri-<br />

bution to the integral comes from a small region around x0. Our integral turns into a<br />

st<strong>and</strong>ard Gaussian integral <strong>and</strong> the result for ξ goes to infinity is simply<br />

A ≈ e −ξf(x0)<br />

�<br />

2π<br />

ξf ′′ �1/2 . (5.60)<br />

(x0)


5. Extremal Black Holes in String Theory 55<br />

Let us now use this method, called the saddle point approximation, for (5.57). We<br />

assume the second derivative (which is a 1/β 3 term) to be small. Then we can easily<br />

find the maximum,<br />

Plugging this value <strong>of</strong> β in give<br />

f(β) = βN + 4π2<br />

β<br />

f ′ (β) = 0 at β = β0 = 2π<br />

√ N<br />

(5.61)<br />

(5.62)<br />

f(β0) = 4π √ N = 4π √ nω (5.63)<br />

So now we simply apply the above saddle point approximation <strong>and</strong> get (in our case<br />

c = 24)<br />

Ω(N) ∼ e 4π√ N = e 4π √ nω . (5.64)<br />

Clearly, for some general central charge c, we get what is called Cardy’s formula,<br />

�<br />

cN<br />

f(β0) = 2π<br />

(5.65)<br />

6<br />

<strong>and</strong> the number <strong>of</strong> states is<br />

Ω(N) ∼ e 2π<br />

√<br />

cN<br />

6 . (5.66)<br />

This kind <strong>of</strong> manipulations were already studies by Ramanujan <strong>and</strong> Cardy because<br />

what we have been doing is basically equivalent to counting the number <strong>of</strong> ways to<br />

partition an integer N using integers <strong>of</strong> 24 colors. Physically, what we have been<br />

solving is the problem <strong>of</strong> distributing integer energy among 24 oscillators.<br />

5.8 D=5 Reissner-Nordström <strong>black</strong> holes<br />

In this important section we will investigate the simplest example <strong>of</strong> a <strong>black</strong> hole in<br />

string theory for which a <strong>microscopic</strong> description can be found. For this we will need<br />

the five dimensional analogue <strong>of</strong> the charged <strong>black</strong> hole solution (2.15) that we already<br />

studies in chapter 2:<br />

ds 2 �<br />

= 1 − 2M Q2<br />

+<br />

r2 r4 �<br />

dt 2 �<br />

− 1 − 2M Q2<br />

+<br />

r2 r4 �−1 dr 2 − r 2 dΩ 2 3. (5.67)


5. Extremal Black Holes in String Theory 56<br />

The gauge potential 1-form has an 1/r 2 dependence because we are in four spatial<br />

dimensions<br />

A = Q<br />

dt. (5.68)<br />

r2 Analogously to the four dimensional case, this may, in the extremal case M = Q, be<br />

cast into an isotropic form by introducing a new coordinate ρ ≡ � r 2 − Q. We get<br />

ds 2 = H −2 dt 2 − H(dρ 2 + ρdΩ 2 3), (5.69)<br />

A = (1 − H −1 )dt, H = 1 + Q<br />

. (5.70)<br />

ρ2 We will now try to connect this 5D <strong>black</strong> hole with string theory via the effective field<br />

theory describing string theory in the low-energy limit.<br />

5.8.1 Derivation <strong>of</strong> <strong>macroscopic</strong> entropy<br />

From our discussion in sections 5.4 <strong>and</strong> 5.5 it follows that a solution <strong>of</strong> the field equa-<br />

tions that one gets from the effective action (5.33) for the metric gµν, with non-zero<br />

field strength F(n), gives the gravitational field associated with an (n − 2)-brane hav-<br />

ing the associated NS-NS or R-R charge. This motivates us to spilt our coordinates<br />

into two groups. One for the n − 1 = p + 1 dimensions <strong>of</strong> the p-brane world volume<br />

with Poincaré invariance (‘longitudinal’) <strong>and</strong> one for the rotational invariance <strong>of</strong> the<br />

remaining 10−(p+1) = 9−p directions (‘transversal’). So let us split our coordinates<br />

x µ as x a , with a = 0, 1, ..., p for the longitudinal ones, <strong>and</strong> y i , with i = (p + 1), ..., 9<br />

for the transversal ones. We would like to know the metric. It turns out that the line<br />

element <strong>of</strong> this configuration in the string frame is <strong>of</strong> the form<br />

ds 2 = H(r) −1/2<br />

�<br />

(dx 0 ) 2 −<br />

p�<br />

a=1<br />

(dx a ) 2<br />

�<br />

− H(r) 1/2<br />

Let us say briefly where this expression comes from.<br />

9�<br />

(dy i ) 2 . (5.71)<br />

i=p+1


5. Extremal Black Holes in String Theory 57<br />

Ansatz<br />

A logical ansatz would be something like the 5D Reissner-Nordström <strong>black</strong> hole, so<br />

we assume the metric to have the form<br />

ds 2 = e 2A(r) ηabdx a dx b − e 2B(r) δijdy i dy j , (5.72)<br />

where r is the radial coordinate, r = � y i y i . For the dilaton <strong>and</strong> antisymmetric tensor,<br />

our ansatz is<br />

φ = f(r), (5.73)<br />

C012...p = e C(r) − 1, (5.74)<br />

Now the claim is that there is a solution for the classical field equations following from<br />

(5.33) in which all <strong>of</strong> the functions A, B, C, f are determined by a single harmonic<br />

function H(r) [33], as follows<br />

ds 2 = Hp(r) −(7−p)/8 ηabdx a dx b − Hp(r) (p+1)/8 δijdy i dy j , (5.75)<br />

e φ = Hp(r) (3−p)/4 , (5.76)<br />

C(p+1) = (Hp(r) −1 − 1)dx 0 ∧ dx 1 ∧ ... ∧ dx p . (5.77)<br />

where the harmonic function Hp(r) is given by<br />

Hp(r) = 1 + L7−p p<br />

, (5.78)<br />

r7−p with<br />

L 7−p<br />

p = 2κ10<br />

√<br />

πqp<br />

(2π<br />

(7 − p)Ω8−p<br />

√ α ′ ) 3−p , (5.79)<br />

where qp is an integer <strong>and</strong> Ωn is the volume <strong>of</strong> the unit n-sphere. This looks so com-<br />

plicated because it is for general p. For example for p = 5 this reduces (using (5.28))<br />

to<br />

L 2 5 = q5gsα ′ . (5.80)<br />

By transforming (5.75) using (5.32) <strong>and</strong> (5.76) we indeed get a line element that in the<br />

string frame looks like<br />

ds 2 = Hp(r) −1/2 ηabdx a dx b − Hp(r) 1/2 δijdy i dy j . (5.81)


5. Extremal Black Holes in String Theory 58<br />

Do any <strong>of</strong> these solutions give the metric <strong>of</strong> a <strong>black</strong> hole? Well, maybe. At this point<br />

it is clear that there are two problems:<br />

• The warp factors Hp(r) are wrong.<br />

• The dilaton is singular at the event horizon.<br />

Extremality <strong>and</strong> supersymmetry<br />

The above solutions (5.81) are extreme solitons, as one can check comparing the mass<br />

that can be read <strong>of</strong> from Hp(r), <strong>and</strong> the charge associated with the C(p+1) form gauge<br />

potential that one gets by integrating ∗ F(p+2), namely<br />

2π<br />

Mp = Qp = qp<br />

gs<br />

�<br />

2π √ α ′<br />

�−(1+p) . (5.82)<br />

(Note that this diverges as gs → 0, as expected for non-perturbative objects.) It is<br />

important to note that the solitons have qp units <strong>of</strong> some fundamental Dp-brane charge.<br />

We would like to get our 5D Reissner-Nordtröm metric back, <strong>and</strong> by looking at<br />

(5.78) this seems to have the best chances if we compactify the x 1 , ..., x 5 coordinates<br />

on a five torus, <strong>and</strong> take p = 5.<br />

Indeed, this is a good choice in the light <strong>of</strong> a result that follows from supersym-<br />

metry considerations. The point is that the solution is invariant under 16 supercharges<br />

Qα + P ˜ Qα, where Q acts on the rightmovers, ˜ Q on the leftmovers, <strong>and</strong> P represents<br />

the operator that reflects the directions y i transverse to the brane. By looking at two<br />

reflection operators P1 <strong>and</strong> P2 we find that the preserved supersymmetry charges sat-<br />

isfy<br />

Qα + P1 ˜ Qα = Qα + P2 ˜ Qα = Qα + P1(P −1 P2) ˜ Qα. (5.83)<br />

This can be shown to imply that the number <strong>of</strong> directions that are transverse one brane<br />

<strong>and</strong> parallel to the other is a multiple <strong>of</strong> four [34]. So because <strong>of</strong> the above extremality<br />

<strong>and</strong> supersymmetry argument we may combine the p = 5 solutions with the p = 1<br />

solutions. We call the coordinates on the p = 1 soliton x 0 <strong>and</strong> x 1 . The remaining four<br />

coordinates x a are transverse to the 1-brane but parallel to the 5-brane.


5. Extremal Black Holes in String Theory 59<br />

Compactification <strong>and</strong> ‘smearing’<br />

The directions x 2 , ..., x 5 are compactified on T 4 ⊂ T 5 <strong>of</strong> volume V4 which leads to a<br />

smearing <strong>of</strong> the D1-branes. This is analogous to the situation <strong>of</strong> the ‘method <strong>of</strong> images’<br />

used in Kaluza-Klein theory. Let us for example look only at the compactification in<br />

the x 2 direction on a circle <strong>of</strong> radius R. The D1-branes are then replaced by an infinite<br />

array <strong>of</strong> parallel branes a distance 2πR apart. This gives an effective harmonic function<br />

˜H1 for the array. For r ≫ R this function varies as<br />

� ˜r 2 − (x 2 ) 2 � −5/2 . (5.84)<br />

It is as if the D1-branes also wrap x2 √<br />

. The effective number <strong>of</strong> D2-branes is q1 α ′ /R.<br />

This case is easily generalized to the case where all coordinates are compactified on<br />

T 4 . The effective number ˜q5 <strong>of</strong> D5-branes is given by<br />

˜q5 = (2π√α ′ ) 4<br />

q1. (5.85)<br />

This is the number that should be plugged into (5.79) which then gives ˜ H1 from (5.78)<br />

with p = 5. Now by the ‘harmonic function sum rule’ [35] we can simply take products<br />

<strong>of</strong> the warp factors <strong>and</strong> get for the combined system<br />

ds 2 =H −1/4<br />

5<br />

− H 3/4<br />

5<br />

V4<br />

˜H −3/4<br />

1 [(dx 0 ) 2 − (dx 1 ) 2 ] − H −1/4 ˜H 5<br />

1/4<br />

1 [(dz 2 ) 2 + ... + (dx 5 ) 2 ]<br />

˜H 1/4<br />

1 [(dy 6 ) 2 + ... + (dy 9 ) 2 ], (5.86)<br />

e φ = H −1/2<br />

5<br />

˜H 1/2<br />

1 , (5.87)<br />

C(5) = [H −1<br />

5 − 1]dx 0 ∧ dx 1 ∧ ... ∧ dx 5 , C(1) = [ ˜ H −1<br />

1 − 1]dx 0 ∧ dx 1 . (5.88)<br />

This solves one <strong>of</strong> our problems. The dilaton is now finite at the event horizon r = 0.<br />

Our final mission is to repair the warp factor.


5. Extremal Black Holes in String Theory 60<br />

Adding momentum in the x 1 -direction<br />

The solution is to add momentum along the x 1 -direction. This leads [35] to the final<br />

form <strong>of</strong> the metric<br />

ds 2 =H −1/4<br />

5<br />

− H −1/4<br />

5<br />

˜H −3/4<br />

1 [(dx 0 ) 2 − (dx 1 ) 2 + HP (dx 0 − dx 1 ) 2 ]<br />

˜H 1/4<br />

1 [(dx 2 ) 2 + ... + (dx 5 ) 2 ] − H 3/4 ˜H 5<br />

1/4<br />

1 [dr 2 + r 2 dΩ 2 3]. (5.89)<br />

The harmonic functions may be written in the form<br />

Using (5.28), (5.79) <strong>and</strong> (5.85),<br />

˜H1 = 1 + r2 1<br />

r 2 , H5 = 1 + L2 5<br />

r2 , HP = r2 P<br />

r<br />

2 . (5.90)<br />

r 2 1 = q1gsα ′ 2π√α ′ ) 4<br />

, L<br />

V4<br />

2 5 = q5gsα ′ , r 2 P = ng 2 sα ′ 2π√α ′ ) 4 α<br />

V4<br />

′<br />

, (5.91)<br />

R2 with R the radius <strong>of</strong> the circle upon which the x 1 coordinate is compactified <strong>and</strong> n an<br />

integer specifying the momentum P = n/R on the circle.<br />

The entropy <strong>of</strong> this solution<br />

Taking the (eight-dimensional) volume <strong>of</strong> the time slice at the horizon, we get the<br />

‘area’ AH <strong>of</strong> the event horizon, which is, taking the limit r → 0 (from above) equal to<br />

AH = � (r −3/4<br />

1<br />

L −1/4<br />

rP )2πR �� (r 1/4<br />

5<br />

1 L −1/4<br />

5 ) 4 V4<br />

�� (r 1/4<br />

1 L 3/4<br />

5 )2π 2�<br />

= 4π 3 RV4r1L5rP . (5.92)<br />

The surface gravity is zero. Thus the temperature <strong>of</strong> this extreme <strong>black</strong> hole is zero.<br />

This is not surprising since this in an extreme <strong>black</strong> hole. It does however have an<br />

entropy. The entropy associated with this <strong>black</strong> hole simplifies to<br />

Sbh = AH<br />

4G10<br />

= 2π √ q1q5n (5.93)<br />

<strong>and</strong> is thus independent <strong>of</strong> both the string coupling gS <strong>and</strong> the size <strong>of</strong> the compact<br />

dimension. It only depends on the integers q1, q5 <strong>and</strong> n.<br />

Next, we will show how one gets this formula on the <strong>microscopic</strong> side.


5. Extremal Black Holes in String Theory 61<br />

5.8.2 Derivation <strong>of</strong> <strong>microscopic</strong> entropy<br />

The above constructed <strong>black</strong> hole solution can be interpreted as a bound state <strong>of</strong> q5<br />

D5-branes <strong>and</strong> q1 D1-branes with some momentum n/R. The reason behind this is<br />

that, as we have shown in (5.82), the soliton solution has qp units <strong>of</strong> p-brane charge<br />

(we also argued that Dp-branes are sources <strong>of</strong> R-R charges).<br />

lows<br />

Now, our configuration (5.89) breaks the 10-dimensional Lorentz symmetry as fol-<br />

SO(1, 9) → SO(1, 1) × SO(4)� × SO(4)⊥. (5.94)<br />

The SO(1, 1) acts on the D1-brane world sheet (x 0 , x 1 ). The SO(4)� acts on the<br />

rest <strong>of</strong> the D5-brane world volume (x 2 , ..., x 5 ). The final factor SO(4)⊥ acts on the<br />

remaining dimensions (y 6 , ..., y 9 ) that are transverse to both. Since the rigid branes<br />

don’t carry linear nor angular momentum, it is an important question to ask, what<br />

degrees <strong>of</strong> freedom do carry the momentum n/R. Well, an obvious possibility is the<br />

these degrees are the massless states <strong>of</strong> open strings that begin <strong>and</strong> end on D-branes.<br />

(The massive excitations <strong>of</strong> the D-branes themselves do not play any role when the<br />

coupling is weak, since they have masses proportional to g −1<br />

S .) Now there are four<br />

possibilities: 1-1 states, 5-5 states, 1-5 states <strong>and</strong> 5-1 states (by this notation we mean<br />

that in an n-m state the open string begins on a Dn-brane <strong>and</strong> ends on a Dm-brane).<br />

For the 1-1 <strong>and</strong> 5-5 state we are taken away from the maximal degeneracy state, which<br />

is the <strong>black</strong> hole state, by the separation <strong>of</strong> the individual D1- <strong>and</strong> D5-branes from<br />

each other. We thus only consider the 1-5 <strong>and</strong> 5-1 states. These give a total <strong>of</strong> 4q1q5<br />

bosons <strong>and</strong> an equal number <strong>of</strong> fermions.<br />

What we want to do, is to distribute the momentum n/R among the 4q1q5 bosons<br />

<strong>and</strong> fermions. This is basically the same problem as we have seen before, in section<br />

5.1 where we estimated the number <strong>of</strong> bosonic string states at mass level n. There is,<br />

however, a small complication, because now also have fermions.<br />

As before, we solve the problem by using a generating function (this is the partition<br />

function)<br />

Z(w) ≡ tr w N =<br />

∞�<br />

dnw n , (5.95)<br />

n=0


5. Extremal Black Holes in String Theory 62<br />

where N = P R <strong>and</strong> dn is the number <strong>of</strong> states with eigenvalue n <strong>of</strong> N (so it has<br />

momentum n/R). Doing analogous steps to the one in section 5.1 we get<br />

tr w N =<br />

∞�<br />

� m 1 + w<br />

1 − wn �4q1q5 . (5.96)<br />

m=1<br />

The fermionic contribution in the numerator is even simpler than the bosonic. In the<br />

one dimensional bosonic case we have many multiparticle states, α−n|0〉, (α−n) 2 |0〉,<br />

etc., which, as we have seen lead to<br />

tr w P ∞<br />

n=1 α−nαn =<br />

∞�<br />

tr w α−nαn =<br />

n=1<br />

∞�<br />

∞�<br />

n=1 i=1<br />

(w n ) i =<br />

∞�<br />

n=1<br />

(1 − w n ) −1 . (5.97)<br />

In the case that our modes were fermions, ψ−n, we would, by Pauli’s exclusion<br />

principle, not be able to make multiparticle states (ψ−n) 2 |0〉, so our partition function<br />

is simply<br />

result<br />

or<br />

tr w P ∞<br />

n=1 ψ−nψn =<br />

∞�<br />

tr w ψ−nψn =<br />

n=1<br />

∞�<br />

(1 + w n ). (5.98)<br />

Finally, we may estimate dn for large values <strong>of</strong> n, as we did in section 5.1 with the<br />

n=1<br />

dn ∼ e 2π√ q1q5n , (5.99)<br />

Sbh = 2π √ q1q5n (5.100)<br />

Basically, this is the Strominger-Vafa result [8]. There are however some differences,<br />

which we explain in the next section. Before we do that let us note that we can permute<br />

the three charges among themselves in all possible ways. Note that by doing four T-<br />

dualities along the T 4 we can exchange the D1 <strong>and</strong> D5 branes with each other, leaving<br />

the momentum P unchanged. Another set <strong>of</strong> dualities can map D1-D5-P to P-D1-D5.<br />

Via S-duality this then can be mapped to P-NS1-NS5. (Therefore, keeping only the<br />

first two charges the D1-D5 bound state is dual to the P-NS1 bound state, which is<br />

simply an elementary string carrying vibrations.) For more details see section 2.5 <strong>of</strong><br />

[36].


5. Extremal Black Holes in String Theory 63<br />

5.9 Link with the Strominger-Vafa result<br />

The case treated by Strominger <strong>and</strong> Vafa in their break-through article [8] was only a<br />

little bit different. They took (type IIB string theory) on the K3 manifold instead <strong>of</strong><br />

our T 4 . So in their case there are<br />

Q5 D5−branes wrapping K3 × S 1 , <strong>and</strong><br />

Q1 D1−branes wrapping S 1 .<br />

Again, the D1-brane can move in four directions transverse direction inside the D5-<br />

brane, <strong>and</strong> we have 4Q1Q5 bosons <strong>and</strong> as many fermions. Now bosons have central<br />

charge 1, whereas the fermions have central charge 1/2. This gives a central charge <strong>of</strong><br />

the corresponding conformal field theory for this situation<br />

The Cardy formula, equation (5.66), gives us the answer<br />

(1 + 1/2)4Q1Q5 = 6Q1Q5. (5.101)<br />

Ω(N) = e 2π<br />

√<br />

cN<br />

6 = e 2πQ1Q5N<br />

, (5.102)<br />

where N is the momentum. In fact, Strominger <strong>and</strong> Vafa use axion charge QH <strong>and</strong><br />

electric charge QF <strong>of</strong> the fundamental string. By the axion charge QH we mean the flux<br />

<strong>of</strong> the H-field; it st<strong>and</strong>s for the charge <strong>of</strong> the NS5-brane. By applying S-duality QH<br />

<strong>and</strong> QF become equal to the Q1 <strong>and</strong> Q5 charges <strong>of</strong> the D1- <strong>and</strong> D5-branes above. So<br />

this is the link between what we did in the previous section, <strong>and</strong> what Strominger <strong>and</strong><br />

Vafa do in their paper. They get for the degeneracy <strong>of</strong> the BPS solitons (or equivalently<br />

the growth <strong>of</strong> the so-called elliptic genus, which they introduce in their paper) for large<br />

charge QH<br />

S = 2π<br />

�<br />

QH( 1<br />

2 Q2 F + 1), (5.103)<br />

which agrees to leading order with the Bekenstein-Hawking entropy for large Q 2 F ,<br />

�<br />

QHQ<br />

SBH = 2π<br />

2 F<br />

. (5.104)<br />

2


5. Extremal Black Holes in String Theory 64<br />

Strominger <strong>and</strong> Vafa speculate that the elliptic genus is a more relevant physical<br />

quantity than the actual number <strong>of</strong> BPS states, which may depend on the moduli <strong>of</strong> the<br />

K3. The point is that the elliptic genus is an appropriately weighted sum (with weights<br />

±1) <strong>and</strong> is moduli independent. However, either quantity will give the same leading<br />

degeneracy as a function <strong>of</strong> the charges, <strong>and</strong> therefore this subtlety is unimportant<br />

for their paper. Nevertheless, for many other purposes, like our counting <strong>of</strong> dyons in<br />

N = 4 string theory in chapter 7, the elliptic genus turned out to be very important.


6. MATHEMATICAL TOOLS<br />

6.1 Modular forms<br />

A function <strong>of</strong> rational character on the plane can have two (or more?) complex periods.<br />

A nice example <strong>of</strong> such a function is Jacobi’s sinus amplitudinus [37]. Let f be such a<br />

(nonconstant, single-valued) function with periods ω: f(x + ω) = f(x). These form a<br />

module L over Z. (Any integer combination <strong>of</strong> periods n1ω1 + n2ω2 is also a period.)<br />

For ‘most’ functions L is trivial (most functions are not periodic), for some functions<br />

(like the sinus) it is <strong>of</strong> rank one (L = 2πZ), <strong>and</strong> for some (like this so-called sinus<br />

amplitudinus) L is <strong>of</strong> rank two. Jacobi [37] proved that there are no other possibilities.<br />

In the third case, we call L a period lattice, <strong>and</strong> f an elliptic function.<br />

The question one can ask now is: What are all the possible pairs <strong>of</strong> primitive peri-<br />

ods for a fixed lattice L = ω1Z ⊕ ω2Z? (For example, ω1 + 6ω2 <strong>and</strong> ω2 will do just as<br />

well as ω1 <strong>and</strong> ω2.) One can write one pair in terms <strong>of</strong> another, like<br />

ω ′ 2 = aω2 + bω1, (6.1)<br />

ω ′ 1 = cω2 + dω1, (6.2)<br />

where a, b, c, d ∈ Z. It is easily seen that the determinant <strong>of</strong> the transformation matrix<br />

should be −1 or +1. Without further comment, we choose it (we can do this 1 ) to be<br />

+1. So what we learn from this, is that two pairs <strong>of</strong> primitive periods are related by a<br />

linear substitution � a b<br />

c d<br />

�<br />

∈ SL(2, Z). (6.3)<br />

Conversely, any such substitution produces new primitive periods from old ones. The<br />

corresponding period ratios ω = ω2/ω1 are related by the associated fractional linear<br />

1 This can be done by imposing the condition that the period ratio ω = ω2/ω1 should have positive<br />

imaginary part.


substitution by<br />

6. Mathematical Tools 67<br />

ω ↦→<br />

aω + b<br />

cω + d<br />

(6.4)<br />

from the projective group P SL(2, Z) = SL(2, Z)/(±1). This is called the modular<br />

group <strong>of</strong> first level Γ1. Two tori (tori can be written as C/L) are (conformally) equiv-<br />

alent if <strong>and</strong> only if their periods are related by a substitution <strong>of</strong> this kind.<br />

tions<br />

One easily checks that an element from P SL(2, Z) is generated by the transforma-<br />

T : ω → ω + 1, S : ω → −1/ω. (6.5)<br />

The change ω → ω ′ = (aω + b)/(cω + d) is called a modular transformation. The<br />

generators T <strong>and</strong> S satisfy S 2 = (ST ) 3 .<br />

A modular form <strong>of</strong> weight 2n ∈ N is an analytic function f on the open half-plane<br />

H that transforms under the modular group Γ1 according to<br />

� �<br />

aω + b<br />

f = (cω + d)<br />

cω + d<br />

2n f(ω). (6.6)<br />

Such a form is <strong>of</strong> period 1 <strong>and</strong> so can be exp<strong>and</strong>ed in whole powers <strong>of</strong> the parameter<br />

q = e 2πiω in a neighborhood <strong>of</strong> q = 0 (so at ω = i∞ so to speak). It is furthermore<br />

required that this expansion contains no negative powers <strong>of</strong> q. Also, f is to be pole-free<br />

at ∞. When f vanishes at infinity, we call it a cusp form. If we do not require f to be<br />

pole-free at infinity, f is said to be a modular function. It has a Fourier expansion <strong>of</strong><br />

the form<br />

f(ω) = �<br />

n≥−N<br />

anq n<br />

(6.7)<br />

which converges for 0 < q < ɛ ≤ 1 for some ɛ. (If N > 0 then f is said to have a pole<br />

<strong>of</strong> order N at infinity. (So one speaks <strong>of</strong> a modular form precisely when f converges<br />

for all |q| < 1, which implies N ≤ 0; a0 is it’s value an infinity.)<br />

A typical example <strong>of</strong> a modular function is the Eisenstein series. A typical example<br />

<strong>of</strong> a modular form is the Dedekind eta-function. Or perhaps one should say that raised<br />

to a certain power this becomes a real modular form.


6. Mathematical Tools 68<br />

6.2 (Weak) Jacobi forms<br />

The definition <strong>of</strong> a Jacobi function that we give here is a generalization <strong>of</strong> the one<br />

in [38]. We include half integer index, in order to incorporate elliptic genera <strong>of</strong> odd-<br />

dimensional Calabi-Yau Manifolds. Let 0 ≤ k ∈ Z <strong>and</strong> 0 ≤ m ∈ Z/2. We call a<br />

function φ from the upper half-plane H × C → C a Jacobi function <strong>of</strong> weight k <strong>and</strong><br />

index r if it transforms like<br />

�<br />

aω + b<br />

φ<br />

cω + d ,<br />

�<br />

z<br />

= (cω + d)<br />

cω + d<br />

k rcz2<br />

2πi<br />

e cω+d φ(ω, z), (6.8)<br />

φ(ω, z + λω + µ) = (−1) 2r(λ+µ) e −2πir(λ2 ω+2λz) φ(ω, z), (6.9)<br />

where � a b<br />

c d<br />

<strong>and</strong> which has Fourier expansion <strong>of</strong> the form<br />

φ(ω, z) =<br />

�<br />

∈ SL(2, Z), λ, µ ∈ Z (6.10)<br />

�<br />

m≥−M,l∈Z+r<br />

c(n, l)q m y l<br />

(6.11)<br />

(where q = e 2πiω , y = e 2πiz , <strong>and</strong> M ∈ Z), which converges for all 0 < |q| < ɛ ≤ 1<br />

<strong>and</strong> z ∈ C. One easily checks that the product <strong>of</strong> two Jacobi functions with weights<br />

k, k ′ <strong>and</strong> indices r, r ′ is a Jacobi function <strong>of</strong> weight k + k ′ <strong>and</strong> index r + r ′ , so the class<br />

just defined is closed under multiplication. We see that a modular function is a special<br />

kind <strong>of</strong> Jacobi function: namely a Jacobi function <strong>of</strong> index r = 0. (φ(ω, 0) transforms<br />

as a modular function <strong>of</strong> weight k. Also, for r = 0 we have c(m, l) = 0 if l �= 0.)<br />

An important fact, which is essentially proven in [38], is that the coefficients c(m, l)<br />

depend only on 4rm − l 2 <strong>and</strong> on l mod 2r. A weak Jacobi form is a special kind <strong>of</strong><br />

Jacobi function, namely one with a Fourier expansion <strong>of</strong> the form<br />

φ(ω, z) = �<br />

m≥0,l∈Z+r<br />

c(m, l)q m y l , (6.12)<br />

which converges for all |q| < 1. The reason why we are interested in this weak Jacobi<br />

form, is that the elliptic genus <strong>of</strong> a Calabi-Yau manifold <strong>of</strong> complex dimensions d is<br />

given by a weak Jacobi form <strong>of</strong> weight zero <strong>and</strong> index m = d/2, [39].


6. Mathematical Tools 69<br />

In physics it seems we must use weak Jacobi forms <strong>and</strong> not Jacobi forms because<br />

L0 − c/24 > 0 in the Ramond sector <strong>of</strong> a unitary theory. Let us now give the definition<br />

<strong>of</strong> the elliptic genus, <strong>and</strong> study some <strong>of</strong> its properties.<br />

6.3 Elliptic genus<br />

Let us now give a proper definition <strong>of</strong> the elliptic genus for a Kähler manifold M <strong>of</strong><br />

complex dimensions d, study some properties <strong>of</strong> its properties <strong>and</strong> give a geometrical<br />

description [40, 41, 42].<br />

6.3.1 Definition <strong>and</strong> properties<br />

Lets start with an elliptic curve E with modulus τ <strong>and</strong> a line bundle labeled by z ∈<br />

Jac(E) ∼ = E. Recall that a an elliptic curve is basically a type <strong>of</strong> a cubic curve whose<br />

solutions are confined to a region <strong>of</strong> space that is topologically equivalent to a torus.<br />

In fact, an elliptic curve or (complex) torus, provided with a complex <strong>structure</strong>, can be<br />

defined by the quotient<br />

Eτ = C/(Z + τZ). (6.13)<br />

The complex parameter τ determines the complex <strong>structure</strong> <strong>of</strong> the torus, <strong>and</strong> is called<br />

the modulus <strong>of</strong> the torus. One can find the modulus <strong>of</strong> a torus by considering so-called<br />

periods. Each torus has, up to scalings, a unique holomorphic one-form λ = adx,<br />

where a is an arbitrary scaling factor, <strong>and</strong> x is a global complex coordinate. Now<br />

the first homology <strong>of</strong> the torus is spanned by two fundamental 1-cycles, which can be<br />

chosen to be the paths joining the origin with the points τ, <strong>and</strong> 1 (all in C). These are<br />

closed curves on the torus <strong>and</strong> we call them γ1 <strong>and</strong> γ2 respectively. The modulus can<br />

than be found from the quotient <strong>of</strong> the periods<br />

�<br />

γ1<br />

τ =<br />

λ<br />

�<br />

γ2 λ.<br />

(6.14)<br />

Different fundamental cycles give values for τ which are related by modular transfor-<br />

mations.


6. Mathematical Tools 70<br />

Next let us introduce q = e 2πiτ , y = e 2πiz . The elliptic genus is then defined as<br />

χ(M; q, y) = TrH(M)(−1) F y FL q L0− d<br />

8 ¯q ¯ L0− d<br />

8 , (6.15)<br />

where F = FL + FR <strong>and</strong> H(M) is the Hilbert space <strong>of</strong> the N = 2 supersymmetric<br />

field theory with (d-dimensional) target space M. Physically, the elliptic genus count<br />

the number <strong>of</strong> string states with ¯ L0 = 0. It is a genus because it satisfies the following<br />

relations. It is additive under disjoint union,<br />

χ(M ⊔ M ′ ; q, y) = χ(M; q, y) + χ(M ′ ; q, y). (6.16)<br />

For any product space M × M ′ we have<br />

And, finally,<br />

in the sense <strong>of</strong> a complex bordism.<br />

χ(M × M ′ ; q, y) = χ(M; q, y) · χ(M ′ ; q, y). (6.17)<br />

χ(M = ∂N; q, y) = 0, (6.18)<br />

For q = 0 the elliptic genus reduces to a weighted sum over the Hodge numbers,<br />

which is essentially the Hirzeburch χy-genus,<br />

χ(M; 0, y) = �<br />

(−1) p+q y<br />

p,q<br />

p− d<br />

2 h p,q (M). (6.19)<br />

Recall that the Hodge numbers are the dimensions <strong>of</strong> the Hodge cohomology,<br />

h p,q = dim H p,q (M) = dim ker ¯ ∂p,q<br />

im ¯ , (6.20)<br />

∂p,q−1<br />

where ¯ ∂p,q indicates the anti-holomorphic differential ¯ ∂ (which maps a (p, q)-form to<br />

a (p, q + 1)-form) restricted to (p, q)-forms.<br />

For y = 1 the elliptic genus simply reduces the Euler number <strong>of</strong> M,<br />

χ(M; q, 1) = χ(M) = �<br />

(−1) n bn, (6.21)<br />

where bn are the Betti numbers. The second equality is the Euler-Poincaré theorem.<br />

For any topological space, the n-th Betti number bn is defined as the rank <strong>of</strong> the n-th<br />

homology group. The relation between Betti numbers <strong>and</strong> Hodge numbers is<br />

bn = �<br />

h p,q . (6.22)<br />

p+q=n<br />

n


6. Mathematical Tools 71<br />

6.3.2 A geometrical description<br />

First, let us recall the notions <strong>of</strong> a Chern character <strong>and</strong> a Todd class. By the splitting<br />

principle <strong>and</strong> the Whitney product formula, we can write the Chern class <strong>of</strong> a vector<br />

bundle E, denoted by c(E), as [43]<br />

c(E) =<br />

r�<br />

(1 + xi), (6.23)<br />

i=1<br />

where xi is the first Chern class <strong>of</strong> the complex line bundle Li in the Whitney sum; xi<br />

is called a Chern root. Let x1, ..., xn be Chern roots <strong>of</strong> a rank r vector bundle E → M.<br />

Then the Chern character <strong>of</strong> the vector bundle is defined as<br />

r�<br />

ch(E) = e xi . (6.24)<br />

Here e xi is to be regarded as a formal power series. (And in fact it is a polynomial,<br />

since the Chern roots are expressed in symmetric functions which are cohomology<br />

classes that vanish for degree > r.) The Todd class is defined as<br />

i=1<br />

td(E) = � xi 1<br />

i = 1r = 1 + −xi 1 − e 2 c1 + 1<br />

12 (c21 + c2) + 1<br />

24 c1c2 + ... (6.25)<br />

If we take as bundle E the (holomorphic) tangent bundle <strong>of</strong> the some variety X then<br />

we just write td(X).<br />

We defined these two concepts in order to introduce the Hirzebruch-Riemann-Roch<br />

formula, which states that<br />

n�<br />

i=0<br />

(−1) i dim H i (X, E)) = deg((ch(E) · td(X))n), (6.26)<br />

with X a non-singular, projective variety <strong>of</strong> dimension n, <strong>and</strong> E a holomorphic vector<br />

bundle over X. On the right h<strong>and</strong> side <strong>of</strong> (6.26), ch(E) · td(X) is to be calculated in<br />

the so-called Chow ring CH(X), <strong>and</strong> then the n-subscript indicates that we have to<br />

take the part <strong>of</strong> the result which lies in CH n (X). We will not do anything with Chow<br />

rings here, so we will not give the definition; this can be found in Appendix A <strong>of</strong> [44].<br />

For smooth manifolds M, this makes an alternative geometrical definition <strong>of</strong> the<br />

elliptic genus possible. For any vector bundle V define the formal sums<br />

�<br />

V = �<br />

q k<br />

k�<br />

V, SqV = �<br />

q<br />

k≥0<br />

k≥0<br />

q k S k V, (6.27)


6. Mathematical Tools 72<br />

where ∧ k <strong>and</strong> S k denote the k’th exterior <strong>and</strong> symmetric product respectively. Now<br />

let TM <strong>and</strong> ¯ TM be holomorphic <strong>and</strong> anti-holomorphic tangent bundles <strong>of</strong> M. Then one<br />

can define the bundle<br />

�<br />

�<br />

�<br />

d<br />

−<br />

Eq,y = y 2<br />

n≥1<br />

−q n−1 y<br />

TM ⊗ �<br />

−q n y −1<br />

¯TM ⊗ Sq nTM ⊗ Sq n ¯ TM<br />

�<br />

. (6.28)<br />

The elliptic genus can then we defined via the Hirzebruch-Riemann-Roch formula as<br />

�<br />

χ(M; q, y) = ch(Eq,y)td(M). (6.29)<br />

M


7. COUNTING N = 4 DYONS VIA AUTOMORPHIC FORMS<br />

In this final chapter we will focus on the dyonic spectrum <strong>of</strong> N = 4 string theory in<br />

four dimensions. Dyonic states are states that are both electric <strong>and</strong> magnetic at the<br />

same time. These preserve only 1/4 <strong>of</strong> the supersymmetries <strong>and</strong> are quite mysterious<br />

since they can not be treated perturbatively in any description. This in contrast with<br />

purely electric (or magnetic) states which preserve 1/2 <strong>of</strong> the supercharges <strong>and</strong> whose<br />

degeneracies are easily determined (they correspond to the heterotic string states that<br />

are in the right-moving ground state). We will study a proposal <strong>of</strong> Dijkgraaf, Verlinde<br />

<strong>and</strong> Verlinde [45] for the exact degeneracy <strong>of</strong> these dyonic BPS states. Recently, their<br />

proposal was confirmed [46].<br />

Let us do some preparatory work first.<br />

7.1 Electric-magnetic duality, S-duality <strong>and</strong> dyons <strong>of</strong> charge eθ/2π<br />

S-duality is a (conjectured) equivalence between weakly coupled <strong>and</strong> strongly cou-<br />

pled string theory; it relates the perturbative expansion around zero to an expansion<br />

around infinite coupling. It is thus not an equivalence that holds order-by-order in<br />

string perturbation theory. S-duality is a far-reaching equivalence that involves non-<br />

trivial relations among all orders <strong>of</strong> perturbation theory.<br />

Let us begin with considering the free Maxwell equations (so we take J = 0 in<br />

(5.35)), dF = 0 <strong>and</strong> d(∗F ) = 0, or, written out with indices<br />

where F µν<br />

d<br />

∂µF µν = 0, (7.1)<br />

∂µF µν<br />

d<br />

= 0, (7.2)<br />

= 1<br />

2 ɛµναβ Fαβ. One observes that these stay perfectly valid under the trans-


7. Counting N = 4 Dyons via Automorphic Forms 75<br />

formation F ↔ ∗F , or equivalently,<br />

F µν ↔ F µν<br />

d , (7.3)<br />

which is called electric-magnetic duality (or “abelian S-duality”) since it interchanges<br />

the E <strong>and</strong> B field. Note that the above Maxwell equations are different in nature: the<br />

former is an equation <strong>of</strong> motion, derived from the action, while the latter is a Bianchi<br />

identity.<br />

This can be extended to the quantum theory provided the Dirac quantization con-<br />

dition is satisfied [47]. This means that if a particle <strong>of</strong> electric <strong>and</strong> magnetic charge<br />

(Qe, Qm) exists, <strong>and</strong> another <strong>of</strong> charges (Q ′ e, Q ′ m), then<br />

QeQ ′ m − QmQ ′ e ∈ 2πZ. (7.4)<br />

The simplest solution is that the electric charge is a multiple <strong>of</strong> e, <strong>and</strong> the magnetic<br />

charge is a multiple <strong>of</strong> 2π/e, so we get a lattice<br />

Qe = en1, Qm = 2π<br />

e n2, (7.5)<br />

with n1, n2 integers. Note that at this moment something <strong>of</strong> the strong ↔ weak cou-<br />

pling (or S-duality) becomes apparent, since a theory with an electrically charged field<br />

(e, 0) <strong>and</strong> a magnetically charged field (0, 2π/e) is invariant under (7.3) if we also take<br />

e → 2π/e, interchanging strong <strong>and</strong> weak coupling.<br />

We can get a general solution to the Dirac quantization condition by including a<br />

θ-parameter for the gauge field; the electric charges shift by an amount proportional to<br />

the magnetic charge [48],<br />

At this point it is useful to form the combination<br />

θ<br />

Qe = en1 + en2<br />

2π , Qm = 2π<br />

e n2. (7.6)<br />

τ = θ<br />

2π<br />

i<br />

+ . (7.7)<br />

e2 The solution <strong>of</strong> the Dirac quantization condition (7.6) is invariant under θ → θ + 2π,<br />

with n1 → n1 − n2, which can be rephrased as<br />

τ → τ + 1, n1 → n1 − n2. (7.8)


7. Counting N = 4 Dyons via Automorphic Forms 76<br />

Furthermore, one can check that under electric-magnetic duality <strong>and</strong> θ → θ + 2π we<br />

have<br />

τ → − 1<br />

τ , n1 ↔ n2. (7.9)<br />

So these generate the familiar SL(2, Z).<br />

We will use electromagnetism with an angle θ, as above, <strong>and</strong> the special combina-<br />

tion which we called τ to show how one gets the entropy formula [49]<br />

SBH = π � q 2 eq 2 m − (qe · qm) 2 , (7.10)<br />

for the Bekenstein-Hawking entropy for 1/4-BPS <strong>black</strong> holes that we will encounter in<br />

the following.<br />

7.2 Making use <strong>of</strong> the attractor mechanism<br />

From (7.5) the ratio <strong>of</strong> qe <strong>and</strong> qm is equal to e 2 up to some constant <strong>and</strong> we can write<br />

(in this section we will write p <strong>and</strong> q for qe <strong>and</strong> qm respectively)<br />

√ pq ∼ |ep + iq<br />

e |2 = e 2 |p + iq<br />

e 2 |2 . (7.11)<br />

We can write this in a more general form by using τ as in the previous section<br />

τ = θ<br />

2π<br />

i<br />

+ . (7.12)<br />

e2 The angle θ can vary <strong>and</strong> for θ = 0 we get our equation (7.11) back; the generalized<br />

expression is<br />

1<br />

|�p + τ�q| 2 . (7.13)<br />

τ2<br />

Here τ2 is the imaginary part <strong>of</strong> τ <strong>and</strong> one can check that for θ = 0 we indeed get back<br />

(7.11) for θ = 0, since (7.13) is just<br />

e 2 |�p + �qθ i�q<br />

+<br />

2π e2 |2 . (7.14)<br />

The thing we would like to do is to minimize this function (7.13) <strong>of</strong> τ. The so-<br />

called attractor mechanism [50] explains why this we should minimize this at the hori-<br />

zon <strong>of</strong> the <strong>black</strong> hole. So let us minimize<br />

1<br />

|�p + τ�q| 2 = 1<br />

�<br />

|�p| 2 + (τ 2 1 + τ 2 2 )|�q| 2 + 2τ1(�p · �q) 2<br />

�<br />

τ2<br />

τ2<br />

(7.15)


7. Counting N = 4 Dyons via Automorphic Forms 77<br />

first with respect to τ1 <strong>and</strong> plug the result into the minimization with respect to τ2.<br />

Minimizing with respect to τ1 gives<br />

Minimizing with respect to τ2 gives<br />

− |�p|2<br />

τ 2 2<br />

τ1 = −<br />

(�p · �q)<br />

. (7.16)<br />

|�q| 2<br />

− τ 2 1<br />

τ 2 |�q|<br />

2<br />

2 + |�q| 2 − 2τ1(�p · �q)<br />

τ 2 2<br />

By plugging in the τ1 that we found in (7.16) we get<br />

So<br />

<strong>and</strong> we get out desired formula<br />

− 1<br />

τ 2 � 2 (�p · �q)<br />

|�p| +<br />

2<br />

2<br />

|�q| 2 − |�q|2 τ 2 2 −<br />

= 0. (7.17)<br />

2(�p · �q)2<br />

|�q| 2<br />

�<br />

= 0 (7.18)<br />

(�p · �q) 2 − |�p| 2 |�q| 2 = −|�q| 4 τ 2 2 , (7.19)<br />

|�q| 2 τ2 = � |�p| 2 |�q| 2 − (�p · �q) 2 . (7.20)<br />

7.3 The Dijkgraaf-Verlinde-Verlinde degeneracy formula<br />

In this section, we shall recall an interesting conjecture, due to R. Dijkgraaf, E. Ver-<br />

linde <strong>and</strong> H. Verlinde (DVV), which relates the exact degeneracies <strong>of</strong> 1/4 BPS states<br />

in N = 4 string theory, to Fourier coefficients <strong>of</strong> certain automorphic forms [45].<br />

N = 4 string theory has two perturbative formulations, one in terms <strong>of</strong> the heterotic<br />

string compactified on T 6 , <strong>and</strong> a dual one in terms <strong>of</strong> type II strings compactified on<br />

K3 × T 2 . The moduli space factorizes into<br />

SL(2, R)<br />

U(1)<br />

× SO(6, 22, R)<br />

SO(6) × SO(22)<br />

<strong>and</strong> points related by an action <strong>of</strong> the duality group<br />

(7.21)<br />

Γ = SL(2, R) × SO(6, 22, R) (7.22)<br />

are conjectured to be equivalent under non-perturbative dualities. In the point <strong>of</strong> view<br />

<strong>of</strong> the heterotic string we have 28 electric charges qe <strong>and</strong> 28 magnetic charges qm which


7. Counting N = 4 Dyons via Automorphic Forms 78<br />

both lie on the Γ 22,6 lattice. We see that the Bekenstein-Hawking entropy for 1/4-BPS<br />

<strong>black</strong> holes, which is given by 1<br />

SBH = π � q 2 eq 2 m − (qe · qm) 2 (7.23)<br />

is invariant under the duality group (7.22). This entropy equation stays invariant un-<br />

der qe ↔ (−)qm which, <strong>of</strong> course, is related to the first factor <strong>of</strong> (7.22), being an<br />

electromagnetic S-duality which acts on electric charges qeΛ <strong>and</strong> magnetic charges q Λ m,<br />

Λ = 0, ..., 27 transforming in the 28 <strong>of</strong> the second factor.<br />

DVV proposed that the exact degeneracies, or, to be more precise, the number<br />

<strong>of</strong> bosonic minus the number <strong>of</strong> fermionic BPS-multiplets for a given electric <strong>and</strong><br />

magnetic charge, should be given by<br />

�<br />

d(qe, qm) =<br />

dρdσdν eiπ(q2 mρ+q2 eσ+2qe·qmν) . (7.24)<br />

Φ(ρ, σ, ν)<br />

The integral is over the contour 0 ≤ ρ, ν, σ ≤ 2π. The function Ω(ρ, ν, σ) is manifestly<br />

SL(2, Z) invariant, which suggest that it can be written as a modular form in<br />

� �<br />

ρ ν<br />

Ω =<br />

ν σ<br />

(7.25)<br />

It is natural to identify the matrix Ω with the period matrix <strong>of</strong> a genus two Riemann<br />

surface. This period matrix Ω transforms under Sp(2, Z), which can be written as a<br />

4 × 4 matrix decomposed onto four real 2 × 2 blocks, A, B, C, D,<br />

� �<br />

A B<br />

C D<br />

with<br />

Then Ω transforms as follows under Sp(2, Z),<br />

(7.26)<br />

A T D − C T B = DA T − CB T = 12, (7.27)<br />

A T C = C T A, B T D = D T B. (7.28)<br />

Ω → Ω ′ = (AΩ + B)(CΩ + D) −1<br />

1 See section 7.2 for a quantitative explanation <strong>of</strong> where this expression comes from.<br />

(7.29)


7. Counting N = 4 Dyons via Automorphic Forms 79<br />

In fact, Ω is the unique automorphic form <strong>of</strong> weight 10 <strong>of</strong> the modular group Sp(2, Z).<br />

Combining the electric <strong>and</strong> magnetic charges into a vector<br />

� �<br />

qm<br />

q =<br />

the degeneracy formula takes the form<br />

�<br />

d(qe, qm) =<br />

qe<br />

(7.30)<br />

dΩ eiπq·Ω·q<br />

. (7.31)<br />

Φ(Ω)<br />

This can be seen as a ‘logical’ generalization <strong>of</strong> the 1/2-BPS purely electric case, in<br />

which case the number <strong>of</strong> multiplets is given by<br />

�<br />

d(qe) =<br />

dσ eiπσq2 e<br />

, (7.32)<br />

η(σ) 24<br />

where the integral over σ is from 0 to 1 <strong>and</strong> η(σ) is the Dedekind η-function.<br />

DVV carried out several consistency checks, some <strong>of</strong> which we will briefly discuss.<br />

7.3.1 Consistency checks<br />

One consistency check one can do, that was done by DVV, is to extract the large<br />

charge behavior <strong>of</strong> d(qe, qm) by computing the contour integral in (7.24) by residues.<br />

Of course, what one expects to find, is agreement with (7.23), <strong>and</strong> indeed it can be<br />

shown that in the large charge limit<br />

d(qe, qm) ∼ e π<br />

√<br />

q2 eq2 m−(qe·qm)2 . (7.33)<br />

This check can be done by picking the residue at the divisor D = ρσ − ν 2 + ν <strong>and</strong><br />

using<br />

where<br />

See [45], [51] <strong>and</strong> [52] for some more details.<br />

Φ = D 2 η 24 (ρ ′ )η 24 (σ ′ )/det 12 (Ω), (7.34)<br />

ρ ′ = − σ<br />

ρσ − ν2 , σ′ = − ρ<br />

. (7.35)<br />

ρσ − ν2 As another check, <strong>and</strong> continuing our discussion <strong>of</strong> Sp(2, Z) above, let us note that<br />

the SL(2, Z) duality transformations are identified with the subgroup <strong>of</strong> Sp(2, Z) that


7. Counting N = 4 Dyons via Automorphic Forms 80<br />

leave the genus two modular form Φ(Ω) invariant, which leads us to the conclusion<br />

that the presented degeneracies are manifestly duality symmetric. (Using the modular<br />

invariance <strong>of</strong> Φ, one can cancel the action <strong>of</strong> SL(2, Z) by a change <strong>of</strong> the integration<br />

contour γ → γ ′ , <strong>and</strong> deform γ ′ back to γ while avoiding singularities.)<br />

7.4 State counting <strong>and</strong> the elliptic genus<br />

The way in which the elliptic genus makes its entrance in the counting <strong>of</strong> N = 4 dyons<br />

is via the cusp form Φ, which was an infinite product representation<br />

Φ(ρ, σ, ν) = e 2πi(ρ+σ+ν)<br />

�<br />

�<br />

1 − e 2πi(kρ+lσ+mν)<br />

�c(4kl−m2 )<br />

, (7.36)<br />

(k,l,m)>0<br />

where (k, l, m) > 0 means that k, l ≥ 0 <strong>and</strong> m ∈ Z, m < 0 for k = l = 0. The<br />

coefficients c(k) are the Fourier coefficients <strong>of</strong> the elliptic genus <strong>of</strong> K3,<br />

χ(K3; ρ, ν) = �<br />

c(4h − m 2 )e 2πi(hρ+mz) . (7.37)<br />

h≥0,m∈Z<br />

It is known [41] that the elliptic genus <strong>of</strong> K3, which is defined by<br />

χ(K3, τ, z) = Tr(−1) F c<br />

2πi(τ(L0−<br />

e 24 )+zFL)<br />

(7.38)<br />

plays an import role in the counting <strong>of</strong> certain string states. In the above expression<br />

the trace is over the RR-sector <strong>of</strong> a superconformal sigma-model <strong>and</strong> F = FL + FR is<br />

the sum <strong>of</strong> the left- <strong>and</strong> right- moving fermion number. One can think <strong>of</strong> the elliptic<br />

genus as a partition function. An interesting identity for the elliptic genus that we will<br />

just quote (a pro<strong>of</strong> can be found in [42]) is a relation between orbifold elliptic genera<br />

<strong>of</strong> the symmetric product manifolds in terms <strong>of</strong> that <strong>of</strong> a Kähler manifold M is<br />

�<br />

p N χ(S N M; q, y) = �<br />

(7.39)<br />

N=0<br />

n>0,m≥0,l<br />

where the coefficients c(m, l) are defined via the expansion<br />

χ(M; q, y) = �<br />

m≥0,l<br />

1<br />

(1 − p n q m y l ) c(nm,l)<br />

c(m, l)q m y l . (7.40)<br />

There is a nice physical interpretation <strong>of</strong> this identity in terms <strong>of</strong> second quantized<br />

string theory [42].


8. OUTLOOK<br />

At the end <strong>of</strong> this thesis we come to the starting point <strong>of</strong> recent research work on <strong>black</strong><br />

holes in the context <strong>of</strong> string theory. Much can be said at this point, but let us only<br />

mention briefly (quite arbitratily): 1. Near-extremal <strong>black</strong> holes, 2. Mathur’s fuzzball<br />

proposal, <strong>and</strong> 3. Topological strings <strong>and</strong> the OSV conjecture. All three could be a<br />

subject for a thesis <strong>of</strong> its own, so we will just scratch the surfaces.<br />

Near-extremal <strong>black</strong> holes<br />

The Strominger-Vafa result can be generalized to near-extremal <strong>black</strong> holes, as was<br />

soon realized by Callan <strong>and</strong> Maldacena [53]. In the near-extremal case the entropy<br />

becomes a function <strong>of</strong> the mass <strong>of</strong> the <strong>black</strong> hole as well as its four charges. Probably<br />

the most prominent application <strong>of</strong> going away from the BPS limit <strong>and</strong> studying near-<br />

BPS states is the derivation <strong>of</strong> Hawking radiation from the D-brane perspective. By<br />

adding a small amount <strong>of</strong> right-moving momentum to a purely left-moving BPS state<br />

one gets a near-BPS state. Intuitively one can imagine that what happens in the D-<br />

brane picture is that left- <strong>and</strong> right-moving open strings interact to form closed string<br />

states <strong>and</strong> leave the brane. This is indicated in picture 8.1. A nice review <strong>of</strong> various<br />

aspects <strong>of</strong> near extremal <strong>black</strong> holes is Maldacena’s PhD thesis [27].<br />

Note that extremal <strong>black</strong> holes do not Hawking radiate (as we already discussed).<br />

This is clear from the picture: you get Hawking radiation only when a left mover<br />

attaches to a right mover <strong>and</strong> they leave the brane as a closed string (this is the near-<br />

extremal case).


8. Outlook 82<br />

Fig. 8.1: In the D-brane picture left- <strong>and</strong> right-moving open strings can interact, form closed<br />

string states <strong>and</strong> leave the brane as Hawking radiation.<br />

Mathur’s fuzzball proposal<br />

Mathur et al, see e.g. [36], set up a whole program in which the aim is to give all mi-<br />

croscopic states <strong>of</strong> a <strong>black</strong> hole an own gravitational solution. This so-called ‘fuzzball’<br />

approach looks on the one h<strong>and</strong> very promising, but is on the other h<strong>and</strong> quite con-<br />

troversial. For example, in the recent article on the Farey tail expansion for N = 2<br />

attractor <strong>black</strong> holes [54], quite the contrary seems to be the case. The major part <strong>of</strong><br />

the <strong>black</strong> hole entropy is carried by one particular <strong>black</strong> hole background, with no one<br />

to one correspondence between microstates <strong>and</strong> gravitational backgrounds.<br />

Topological strings <strong>and</strong> the OSV conjecture<br />

The claim <strong>of</strong> the OSV conjecture [55] is that there is a direct relation between <strong>black</strong><br />

hole entropy <strong>and</strong> the topological string partition function, |Ztop| 2 = ZBH. An excellent<br />

review is [52].


APPENDIX<br />

Appendix A: Geodesic Deviation <strong>and</strong> the Raychaudhuri Equation<br />

As is well known, the defining property <strong>of</strong> flat geometry is the postulate that initially<br />

parallel lines always remain parallel. In curved space this is not true (think for example<br />

<strong>of</strong> a sphere). To study this ‘curved behavior’ we consider a family <strong>of</strong> non-crossing<br />

geodesics, γs(λ). Here for each s ∈ R, γs(λ) is a geodesic parameterized by the affine<br />

parameter λ. On the two-dimensional surfaces that these geodesics define we pick<br />

coordinates s <strong>and</strong> t. We can think <strong>of</strong> our surface as a set <strong>of</strong> points x µ (s, t). It is natural<br />

to define vector fields T µ (tangent vectors) <strong>and</strong> S µ (deviation vectors) as indicated in<br />

figure A.1, namely by<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

t<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

T µ = ∂xµ<br />

∂t <strong>and</strong> Sµ = ∂xµ<br />

∂s<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¦<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

¥<br />

Geodesics<br />

Tangent vector<br />

¢¡ £<br />

¡ £<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

£<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

¤<br />

Deviation vector<br />

s<br />

(8.1)<br />

Fig. 8.2: A set <strong>of</strong> geodesics with tangent vectors <strong>and</strong> deviation vectors. The vector field consisting<br />

<strong>of</strong> the deviation vectors measures the deviation between nearby geodesics.<br />

We want to look at how S µ changes in the T direction, so we define the “relative


velocity <strong>of</strong> geodesics” as<br />

8. Outlook 84<br />

V µ = (∇T S) µ = T ρ ∇ρS µ<br />

<strong>and</strong> also the “relative acceleration <strong>of</strong> geodesics”<br />

(8.2)<br />

A µ = (∇T V ) µ = T ρ ∇ρV µ . (8.3)<br />

A straightforward calculation (see [10], page 146) shows that the relative acceleration<br />

between two neighboring geodesics is proportional to the curvature:<br />

A µ = R µ νρσT ν T ρ S σ<br />

This equation is called the geodesic deviation equation.<br />

(8.4)<br />

Instead <strong>of</strong> looking at a one-parameter family <strong>of</strong> geodesics we will now consider an<br />

entire congruence <strong>of</strong> geodesics. A congruence in O ⊂ M, where O is an open subset<br />

<strong>of</strong> a manifold M, is a family <strong>of</strong> curves such that through each point p ∈ O there passes<br />

exactly one curve in this family. Clearly, we get a vector field for each congruence <strong>and</strong><br />

the other way around. A congruence is said to be smooth if the corresponding vector<br />

field is smooth.<br />

Our goal in the rest <strong>of</strong> this section will be to define the expansion, shear, <strong>and</strong> twist<br />

<strong>of</strong> timelike geodesic congruences in a spacetime (M, gµnu) <strong>and</strong> derive an equation (the<br />

Raychaudhuri equation) for their rate <strong>of</strong> change as one moves along the curves in the<br />

congruence. One can imagine setting up a set <strong>of</strong> three normal vectors orthogonal to our<br />

timelike geodesic <strong>and</strong> looking at their failure to be parallel transported as we follow<br />

their evolution along the geodesics.<br />

cated.<br />

We will also briefly discuss the null geodesic case which is slightly more compli-<br />

Let T µ = dx µ /dτ be the tangent vector field to a four-dimensional timelike geodesic<br />

congruence; T µ is normalized, TµT µ = −1, <strong>and</strong> obeys the geodesic equation T λ ∇λT µ =<br />

0. We define a tensor field Bµν by<br />

Bµν = ∇νTµ<br />

(8.5)


8. Outlook 85<br />

Physically, Bµν measures the failure <strong>of</strong> S µ to be parallel transported along the congru-<br />

ence. Next we define a “spatial metric” 1<br />

The expansion θ is defined by<br />

hµν = gµν + TµTν<br />

θ = B µν hµν<br />

(8.6)<br />

(8.7)<br />

<strong>and</strong> it describes the change <strong>of</strong> a volume <strong>of</strong> a sphere <strong>of</strong> test particles centered at some<br />

point when moving along their central geodesic. Next we define the shear σµν, which<br />

represents a distortion in the shape <strong>of</strong> our collection <strong>of</strong> test particles from a sphere into<br />

an ellipsoid, by<br />

Finally, we define the twist or rotation<br />

Thus, we can decompose Bµν as<br />

σµν = B(µν) − 1<br />

3 θhµν<br />

(8.8)<br />

ωµν = B[µν]. (8.9)<br />

Bµν = 1<br />

3 θhµν + σµν + ωµν. (8.10)<br />

Now the geodesic deviation equation (8.4) yields equations for the rate <strong>of</strong> change<br />

<strong>of</strong> θ, σµν, <strong>and</strong> ωµν:<br />

dθ<br />

dτ<br />

= −1<br />

3 θ2 − σµνσ µν + ωµνω µν − RµνT µ T ν<br />

(8.11)<br />

This is the timelike version <strong>of</strong> the Raychaudhuri equation. It gives a relation be-<br />

tween the rate <strong>of</strong> change <strong>of</strong> the expansion, <strong>and</strong> the shear <strong>and</strong> twist <strong>of</strong> timelike geodesics<br />

as one moves along the curves in the congruence.<br />

One can also consider null geodesics. The derivation <strong>of</strong> the null version <strong>of</strong> Ray-<br />

chaudhuri’s equation is slightly more complicated (see for example [10] Appendix F,<br />

or [11] chapter 9.2)] <strong>and</strong> we will not treat it here. The main difference is that for null<br />

1 One can think <strong>of</strong> this as some kind <strong>of</strong> projection: any vector from the tangent space TpM can<br />

be projected onto a subspace corresponding to vectors normal to T µ by the ‘projection tensor’ h µ ν =<br />

g µρ hρν = δ µ ν + T µ Tν.


8. Outlook 86<br />

geodesics studying the evolution <strong>of</strong> vectors in a three-dimensional subspace normal to<br />

the tangent field doesn’t make sense, since tangent vectors to null curves are normal<br />

to themselves. Instead, we can study a two-dimensional subspace <strong>of</strong> spatial vectors<br />

normal to the null tangent vector field k µ = dx µ /dλ. The difficulty is that observers in<br />

different Lorentz frames will have different notions <strong>of</strong> what constitutes a spatial vec-<br />

tor, so there is no unique way <strong>of</strong> defining this subspace. For more details we refer the<br />

reader to Appendix F <strong>of</strong> [10]. This difference between the three <strong>and</strong> two dimensional<br />

subspaces in respectively the timelike <strong>and</strong> null geodesic case results in a pre-factor <strong>of</strong><br />

1/2 instead <strong>of</strong> 1/3 in the null Raychauhuri equation, which reads<br />

where<br />

dθ<br />

dλ<br />

= −1<br />

2 θ2 − ˆσµν ˆσ µν + ˆωµν ˆω µν − Rµνk µ k ν , (8.12)<br />

θ = ˆ h µν Bµν<br />

ˆ (8.13)<br />

ˆσµν = ˆ B(µν) − 1<br />

2 θˆ hµν<br />

(8.14)<br />

ˆωµν = ˆ B[µν] (8.15)<br />

are again to be interpreted as the expansion, shear, <strong>and</strong> twist <strong>of</strong> the congruence. The<br />

hatted tensor field ˆ Bµν is given by<br />

ˆBµν = 1<br />

2 θˆ hµν + ˆσµν + ˆωµν, (8.16)<br />

<strong>and</strong> the hatted metric ˆ h µν is the metric on our two-dimensional subspace. Let us be a<br />

little bit more precise about what is meant by this. The tangent vectors in the tangent<br />

space Vp at a point p ∈ M which are orthogonal to k µ form a three-dimensional<br />

subspace ˜ Vp. Now define ˆ Vp to be the vector subspace <strong>of</strong> equivalence classes <strong>of</strong> vectors<br />

in ˜ Vp, with the equivalence relation<br />

x µ ∼ y µ ⇐⇒ x µ − y µ = ck µ<br />

(8.17)<br />

for some c ∈ R. The spacetime metric gµν gives now rise to a ‘hatted metric’ on<br />

ˆVp. This hatted metric is denoted by ˆ hµν. The reason why we put no hat on θ is that<br />

ˆh µν ˆ Bµν = h µν Bµν as one can easily verify.


8. Outlook 87<br />

Appendix B: Samenvatting in het Nederl<strong>and</strong>s<br />

Dit stukje is — in gewijzigde vorm — verschenen op http://www.natuurkunde.nl/<br />

Van zwarte gaten heeft iedereen wel eens gehoord. “Een zwart gat is een bijzonder<br />

soort object in het heelal, waar als gevolg van het sterke zwaartekrachtsveld geen licht<br />

<strong>of</strong> materie kan ontsnappen”, is wat Wikipedia ons vertelt. Maar dat is nog niet het hele<br />

verhaal. Zwarte gaten zijn namelijk zo zwart nog niet. Dat komt door een vreemde<br />

eigenschap van het vacuüm. Je zou denken dat vacuüm ‘niks’ is. Maar dat is niet zo:<br />

uit het ‘niks’ worden aan één stuk door deeltjes en antideeltjes gevormd die vervolgens<br />

weer snel verdwijnen, althans, tenzij er ééntje in een zwart gat valt.<br />

Zwarte gaten<br />

In dit stukje zal ik ervan uitgaan dat u in ieder geval een vaag beeld hebt van wat een<br />

zwart gat is, maar meer dan dat hoeft ook niet. Dat zwarte gaten zwart zijn klinkt<br />

ergens wel logisch. Maar als u zich er wat meer in verdiept blijkt al snel dat zwarte<br />

gaten eigenlijk ‘grijs’ zijn. Ze slokken wel alles op, dat is waar, maar er komen ook<br />

deeltjes uit! Wel altijd deeltjes van ‘hetzelfde soort’. Het maakt niet uit wat u in<br />

het zwarte gat hebt gegooid, <strong>of</strong> het nou een fles was waar nog statiegeld op zat (dat<br />

statiegeld kunt u dan trouwens wel vergeten) <strong>of</strong> een tafel <strong>of</strong> een stoel, er komt altijd<br />

hetzelfde uit.<br />

Tja, zult u zich afvragen, hoe kan dat eigenlijk? Zwarte gaten zijn toch van die<br />

zware sterren waar de ontsnappingssnelheid groter is dan de lichtsnelheid. En niks<br />

kan sneller dan het licht. Dus niks kan ontsnappen, zelfs licht niet. Hoe kunnen die<br />

deeltjes dan wèl ontsnappen?<br />

Vacuüm is niet niks<br />

Helemaal niks bestaat niet in de kwantum mechanica. De onzekerheidsrelaties van<br />

Heisenberg vertellen ons dat dat niet kan. In de kwantum mechanica is namelijk niks<br />

zeker, en als het vacuüm echt helemaal leeg zou zijn, zou u zeker weten dat het exact


8. Outlook 88<br />

leeg is, maar aangezien niks zeker is, kan vacuüm dus simpelweg niet helemaal leeg<br />

zijn. Volgt u het nog?<br />

De natuur lost dit probleem op een wonderbaarlijke manier op. Er worden constant<br />

paren deeltjes en antideeltjes uit het niks gevormd die snel weer verdwijnen door tegen<br />

elkaar aan te botsen (zie figuur 8.3). Zulke deeltjes heten ‘virtuele deeltjes’.<br />

Fig. 8.3: Het vacuüm is niet niks. Er worden constant paren deeltjes en antideeltjes uit het<br />

niks gevormd die snel weer verdwijnen door tegen elkaar aan te botsen. Deze deeltjes<br />

noemen we virtuele deeltjes.<br />

Maar om op het verhaal van de zwarte gaten terug te komen: stelt u zich eens voor<br />

dat er zo een deeltje-antideeltje paar wordt gevormd net naast een zwart gat. Het kan<br />

dan gebeuren dat het antideeltje het zwarte gat invalt, terwijl het deeltje de <strong>and</strong>ere kant<br />

op gaat. Ze kunnen dan niet meer tegen elkaar aanbotsen en verdwijnen. Voor ons lijkt<br />

het er dan op <strong>of</strong> er een deeltje door het zwarte gat is uitgezonden (zie figuur 8.4).<br />

De energie van het zwarte gat neemt op deze manier af. Denk maar aan energiebe-<br />

houd. Het deeltje wat onze kant op is gevlogen heeft een zekere energie en omdat<br />

energie behouden is moet het zwarte gat wel een zelfde hoeveelheid energie zijn ver-<br />

loren. Nu is energie via E = mc 2 equivalent is aan massa. We zien dus dat een zwart<br />

gat op deze manier massa kwijtraakt. Dit is het Hawking effect.


8. Outlook 89<br />

Fig. 8.4: Virtuele deeltjes in de buurt van een zwart gat.<br />

Kwantum zwarte gaten<br />

Waarom is het eigenlijk interessant om zwarte gaten te bestuderen? Daar zijn meerdere<br />

redenen voor. De belangrijkste is dat zwarte gaten zowel erg zwaar als erg klein zijn.<br />

Daarom kan men natuurkundige theorieën die gebruikt worden om zwarte gaten te<br />

beschrijven ook gebruiken om de Big Bang te bestuderen. Toen was immers ook alle<br />

materie heel dicht op elkaar gepakt.<br />

Voor de meeste verschijnselen in de natuur volstaat het om <strong>of</strong>wel kwantum me-<br />

chanica <strong>of</strong>wel relativiteitstheorie te gebruiken. Maar voor het bestuderen van een zwart<br />

gat zijn allebei de theorieën tegelijk nodig en dat brengt een lastig probleem met zich<br />

mee. Het punt is dat kwantum mechanica en relativiteitstheorie elkaar niet zo mogen.<br />

Zodra men berekeningen gaat doen waarbij beide theorieën tegelijk gebruikt worden,<br />

komt er steeds een vervelend ‘oneindig’ uit. Er is dus een nieuwe theorie - zeg maar<br />

een theorie van ‘kwantum gravitatie’ - nodig om verder te komen. Momenteel is de<br />

enige serieuze k<strong>and</strong>idaat voor zo een theorie de snaartheorie. Deze scriptie gaat voor<br />

een groot gedeelte over hoe men gebruik makend van snaartheorie een goed begrip van<br />

zwarte gaten kan krijgen. Aan de <strong>and</strong>ere kant kan een goed begrip van zwarte gaten de<br />

snaartheorie, die nog steeds een theorie in wording is, wellicht een stap verder brengen.


BIBLIOGRAPHY<br />

[1] J. M. Bardeen, B. Carter, <strong>and</strong> S. W. Hawking, “The Four laws <strong>of</strong> <strong>black</strong> hole<br />

mechanics”, Commun. Math. Phys. 31, 161–170, 1973.<br />

[2] S. W. Hawking, “Particle creation by <strong>black</strong> holes”, Commun. Math. Phys. 43,<br />

199–220, 1975.<br />

[3] S. W. Hawking, “Breakdown <strong>of</strong> predictablility in gravitational collapse”, Phys.<br />

Rev. D14, 2460–2473, 1976.<br />

[4] L. Susskind, “Some speculations about <strong>black</strong> hole entropy in string theory”,<br />

1993, [arXiv:hep-th/9309145].<br />

[5] G. T. Horowitz <strong>and</strong> J. Polchinski, “A correspondence principle for <strong>black</strong> holes<br />

<strong>and</strong> strings”, Phys. Rev. D55, 6189–6197, 1997, [arXiv:hep-th/9612146].<br />

[6] G. W. Gibbons, “Supersymmetric soliton states in extended supergravity the-<br />

ories”, In: Proceedings, Unified theories <strong>of</strong> elementary particles, 145-151,<br />

München 1981.<br />

[7] J. Polchinski, “Dirichlet-branes <strong>and</strong> Ramond-Ramond charges”, Phys. Rev. Lett.<br />

75, 4724–4727, 1995, [arXiv:hep-th/9510017].<br />

[8] A. Strominger <strong>and</strong> C. Vafa, “Microscopic origin <strong>of</strong> the Bekenstein-Hawking en-<br />

tropy”, Phys. Lett. B379, 99–104, 1996, [arXiv:hep-th/9601029].<br />

[9] J. B. Hartle <strong>and</strong> S. W. Hawking, “Solutions <strong>of</strong> the Einstein-Maxwell equations<br />

with many <strong>black</strong> holes”, Commun. Math. Phys. 26, 87–101, 1972.


Bibliography 91<br />

[10] S. M. Carroll, Spacetime <strong>and</strong> geometry: An introduction to general relativity,<br />

Addison-Wesley, 2004.<br />

[11] R. M. Wald, General Relativity, Chicago University Press, 1984.<br />

[12] S. W. Hawking <strong>and</strong> G. F. R. Ellis, The large scale <strong>structure</strong> <strong>of</strong> space-time, Cam-<br />

bridge University Press, 1973.<br />

[13] R. P. Kerr, “Gravitational field <strong>of</strong> a spinning mass as an example <strong>of</strong> algebraically<br />

special metrics”, Phys. Rev. Lett. 11, 237–238, 1963.<br />

[14] E. T. Newman, E. Couch, K. Chinnapared, A. Exon, A. Prakash, <strong>and</strong> R. Torrence,<br />

“Metric <strong>of</strong> a Rotating Charged Mass”, J. Math. Phys. 6, 918–919, 1965.<br />

[15] J. D. Bekenstein, “Generalized second law <strong>of</strong> thermodynamics in <strong>black</strong> hole<br />

physics”, Phys. Rev. D9, 3292–3300, 1974.<br />

[16] G. ’t Ho<strong>of</strong>t, “Dimensional reduction in quantum gravity”, 1993, [arXiv:gr-<br />

qc/9310026].<br />

[17] J. M. Maldacena, “The large N limit <strong>of</strong> superconformal field theories <strong>and</strong> super-<br />

gravity”’, Adv. Theor. Math. Phys. 2, 231–252, 1998, [arXiv:hep-th/9711200].<br />

[18] T. Jacobson, “Thermodynamics <strong>of</strong> space-time: The Einstein equation <strong>of</strong> state”,<br />

Phys. Rev. Lett. 75, 1260–1263, 1995, [arXiv:gr-qc/9504004].<br />

[19] W. G. Unruh, “Notes on <strong>black</strong> hole evaporation”, Phys. Rev. D14, 870, 1976.<br />

[20] R. M. Wald, “Black hole entropy in the Noether charge”, Phys. Rev. D48, 3427–<br />

3431, 1993, [arXiv:gr-qc/9307038].<br />

[21] G. Lopes Cardoso, B. de Wit, <strong>and</strong> T. Mohaupt, “Corrections to <strong>macroscopic</strong> su-<br />

persymmetric <strong>black</strong>-hole entropy”, Phys. Lett. B451, 309–316, 1999, [arXiv:hep-<br />

th/9812082].<br />

[22] G. ’t Ho<strong>of</strong>t, “Quantum gravity <strong>and</strong> <strong>black</strong> holes”, In: Proceedings, Nonperturba-<br />

tive quatum field theory, 201-226, Cargese 1987.


Bibliography 92<br />

[23] N. D. Birrell <strong>and</strong> P. C. W. Davies, Quantum fields in curved space, Cambridge<br />

University Press, 1982.<br />

[24] L. H. Ryder, Quantum field theory, Cambridge University Press, 1985.<br />

[25] L. Smarr, “Mass formula for Kerr <strong>black</strong> holes”, Phys. Rev. Lett. 30, 71–73, 1973.<br />

[26] J. Polchinski, String Theory (two volumes), Cambridge University Press, 1998.<br />

[27] J. M. Maldacena, Black holes in string theory, Ph.D. thesis, Princeton University,<br />

1996, [arXiv:hep-th/9607235].<br />

[28] K. Skenderis, “Black holes <strong>and</strong> branes in string theory”, Lect. Notes Phys. 541,<br />

325–364, 2000, [arXiv:hep-th/9901050].<br />

[29] A. W. Peet, “TASI lectures on <strong>black</strong> holes in string theory”, 2000, [arXiv:hep-<br />

th/0008241].<br />

[30] T. Mohaupt, “Black holes in supergravity <strong>and</strong> string theory”, Class. Quant. Grav.<br />

17, 3429–3482, 2000, [arXiv:hep-th/0004098].<br />

[31] D. Bailin <strong>and</strong> A. Love, Cosmology in gauge field theory <strong>and</strong> string theory, Bris-<br />

tol, IOP, 2004.<br />

[32] C. V. Johnson, D-branes, Cambridge University Press, 2003.<br />

[33] G. T. Horowitz <strong>and</strong> A. Strominger, “Black strings <strong>and</strong> p-branes”, Nucl. Phys.<br />

B360, 197–209, 1991.<br />

[34] J. Polchinski, S. Chaudhuri, <strong>and</strong> C. V. Johnson, “Notes on D-Branes”, 1996,<br />

[arXiv:hep-th/9602052].<br />

[35] A. A. Tseytlin, “Harmonic superpositions <strong>of</strong> M-branes”, Nucl. Phys. B475, 149–<br />

163, 1996, [arXiv:hep-th/9604035].<br />

[36] S. D. Mathur, “The fuzzball proposal for <strong>black</strong> holes: An elementary review”,<br />

Fortsch. Phys. 53, 793–827, 2005, [arXiv:hep-th/0502050].


Bibliography 93<br />

[37] C. G. J. Jacobi, Fundamenta Nova Theoriae Functionarum Ellipticarum,<br />

Bornträger, Köninsberg, 1829, reprinted in Gesammelte Werke, vol. 1, 49-239.<br />

[38] M. Eichler <strong>and</strong> D. Zagier, The Theory <strong>of</strong> Jacobi Forms, Birkhäuser, 1985.<br />

[39] C. D. D. Neumann, “The elliptic genus <strong>of</strong> Calabi-Yau 3-folds <strong>and</strong> 4-folds, product<br />

formulae <strong>and</strong> generalized Kac-Moody algebras”, J. Geom. Phys. 29, 5–12, 1999,<br />

[arXiv:hep-th/9607029].<br />

[40] P. L<strong>and</strong>weber (Ed.), Elliptic curves <strong>and</strong> modular forms in algebraic topology,<br />

Lecture Notes in Mathamatics 1326, Springer-Verlag, 1986.<br />

[41] T. Kawai, Y. Yamada, <strong>and</strong> S. Yang, “Elliptic genera <strong>and</strong> N=2 superconformal<br />

field theory”, Nucl. Phys. B414, 191–212, 1994, [arXiv:hep-th/9306096].<br />

[42] R. Dijkgraaf, G. W. Moore, E. P. Verlinde, <strong>and</strong> H. L. Verlinde, “Elliptic genera <strong>of</strong><br />

symmetric products <strong>and</strong> second quantized strings”, Commun. Math. Phys. 185,<br />

197–209, 1997, [arXiv:hep-th/9608096].<br />

[43] M. Nakahara, Geometry, topology <strong>and</strong> physics, Bristol, IOP, 1990.<br />

[44] R. Hartshorne, Algebraic Geometry, Springer, 1977.<br />

[45] R. Dijkgraaf, E. P. Verlinde, <strong>and</strong> H. L. Verlinde, “Counting dyons in N = 4 string<br />

theory”, Nucl. Phys. B484, 543–561, 1997, [arXiv:hep-th/9607026].<br />

[46] D. Shih, A. Strominger, <strong>and</strong> X. Yin, “Recounting dyons in N = 4 string theory”,<br />

2005, [arXiv:hep-th/0505094].<br />

[47] P. A. M. Dirac, “Quantised singularities in the electromagnetic field”, Proc. Roy.<br />

Soc. Lond. A133, 60–72, 1931.<br />

[48] E. Witten, “Dyons <strong>of</strong> charge eθ/2π”, Phys. Lett. B86, 283–287, 1979.<br />

[49] M. Cvetic <strong>and</strong> A. A. Tseytlin, “Solitonic strings <strong>and</strong> BPS saturated dyonic <strong>black</strong><br />

holes”, Phys. Rev. D53, 5619–5633, 1996, [arXiv:hep-th/9512031].


Bibliography 94<br />

[50] S. Ferrara <strong>and</strong> R. Kallosh, “Supersymmetry <strong>and</strong> Attractors”, Phys. Rev. D54,<br />

1514–1524, 1996, [arXiv:hep-th/9602136].<br />

[51] G. Lopes Cardoso, B. de Wit, J. Kappeli, <strong>and</strong> T. Mohaupt, “Asymptotic degener-<br />

acy <strong>of</strong> dyonic N = 4 string states <strong>and</strong> <strong>black</strong> hole entropy”, JHEP 12, 075, 2004,<br />

[arXiv:hep-th/0412287].<br />

[52] B. Pioline, “Lectures on <strong>black</strong> holes, topological strings <strong>and</strong> quantum attractors”,<br />

2006, [arXiv:hep-th/0607227].<br />

[53] J. Callan, C. G. <strong>and</strong> J. M. Maldacena, “D-brane Approach to Black Hole Quan-<br />

tum Mechanics”, Nucl. Phys. B472, 591–610, 1996, [arXiv:hep-th/9602043].<br />

[54] J. de Boer, C. N. Cheng, R. Dijkgraaf, J. Manschot, <strong>and</strong> E. Verlinde, “A Farey<br />

tail for attractor <strong>black</strong> holes”, 2006, [arXiv:hep-th/0608059].<br />

[55] H. Ooguri, A. Strominger, <strong>and</strong> C. Vafa, “Black hole attractors <strong>and</strong> the topological<br />

string”, Phys. Rev. D70, 106007, 2004, [arXiv:hep-th/0405146].

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!