Group 2
Group 2
1
analyze in detail how black holes can behave as information mirrors in one theorized resolution to the BHIP,
as shown in [1].
2
To avoid the coordinate singularity at r = rS , we change variables to ρ2 = 4rS (r − rS ) (this singularity
is not a physical singularity; the only physical singularity is located at r = 0). Then, in the approximation
ρ2 2 2
that the photon is close to the horizon, the metric becomes ds2 = 4r 2 2 2
2 c dt − dρ − rS dΩ , where Ω is the
S
solid angle. We now use the trick of rotating to imaginary time by defining t = −itE . Performing this
c2 dt2
manipulation, the metric appears Euclidean: ds = − dρ + ρ2 4r2E + rS2 dΩ2 . Here, imaginary time takes
2 2
S
the role of an angular coordinate, in analogy with Euclidean polar coordinates in 4 dimensions. We now
ctE
impose a period of 2π on the coordinate 2r S
to avoid a conical singularity. From this, we conclude that the
h̄c
temperature is 4πrS . We are thus led to guess that the black hole has a temperature TH , called the Hawking
temperature,
h̄c3
TH =
8πGM
We can use this fact to give a rough argument of what the entropy of a black hole should be. From the
thermodynamic law dU = T dS, applied to a black hole, we have that d(M c2 ) = TH dS (here we use U = M c2
because all of the energy of a Schwarzschild black hole comes from its mass). This gives the Bekenstein-
Hawking entropy:
A
S= 2 (1)
4lP
q
where lp = h̄G c3 is the Planck length and A is the surface area of the black hole horizon. We can argue
that such an entropy may be somehow related to the microscopic description of the black hole. Nonetheless,
we are unable to correctly identify these microscopic degrees of freedom due to the lack of a full theory of
quantum gravity.
Γ
hN (ω)i = h̄ω . (3)
e kB TH
−1
In this equation Γ is the greybody factor, which quantifies the lower intensity of radiation emitted by the
black hole [4] with respect to the radiation emitted by an ideal black body at temperature TH . We present a
derivation of this relation in Appendix A. This equation is nearly the same as equation (2); its interpretation
is that the number of particles per mode seen by the asymptotic observer is given by a blackbody spectrum,
with temperature TH . We therefore conclude that an observer far away from the black hole perceives the
black hole to emit thermal radiation akin to an object at temperature TH .
Furthermore, one can demonstrate (see Appendix A) that black hole radiation is similar to pair pro-
duction. In particular, for every outgoing particle of radiation, there is another infalling particle, traveling
toward the singularity, which has negative energy. Such a pair of radiation particles is known as a Hawking
pair. This process is not unphysical, because, as we have already stated, energy inside the horizon becomes
a momentum. In fact, an analysis of this situation indicates that each photon belonging to the Hawking
3
radiation outside the horizon is entangled with a corresponding particle inside the horizon. This property
will be fundamental to understanding the black hole information problem.
A distant observer will thus view particles emitted from a black hole with nonzero energy. Since the
black hole is the only source of mass-energy in this scenario, it must be that these particles carry energy
away from the black hole, and its mass decreases as radiation is emitted. As demonstrated in [4], this is
indeed the case: Hawking radiation causes black holes to lose energy and mass, ultimately resulting in their
“evaporation”, at which they cease to exist. For typical astrophysical black holes, this process is incredibly
slow. It is often the case that TH 1, and so these black holes are quite cold. Over humanly time scales, the
energy carried away from these black holes in their Hawking radiation is negligible compared to their total
mass-energy. For instance, even the smallest astrophysical black holes will take at least a few billion years
to evaporate, and larger black holes take even longer. Nevertheless, the fact that black holes do evaporate
has profound implications on our interpretation of information and quantum physics.
The thought experiment previously outlined is useful for understanding the paradox, but it is not really
necessary. In fact, the same result can be achieved by taking into account only Hawking radiation. Let us
consider again a black hole that is initially in a pure state. As we have pointed out in Section 2.2 (see also
Appendix A), when a Hawking pair is generated, it exists in a pure, entangled state. One of the modes, ã,
exists inside the horizon, while the other one, a, exists outside and contributes to Hawking radiation. In
effect, this means that Hawking radiation is entangled with the black hole. However, since this “new part”
of Hawking radiation is produced in a pure state |Ψaã i, the state outside the horizon must have the form
|Ψa i ⊗ |ΨE i where |ΨE i is the state describing all the previously emitted Hawking radiation. In other words,
there cannot be entanglement between the newly emitted Hawking radiation and earlier emitted radiation.
This implies that the entanglement entropy of the Hawking radiation always increases. This fact is quite
intuitive: the earlier Hawking radiation is entangled with modes behind the horizon of the black hole. But
those modes cannot exit the horizon and influence the state of the early Hawking radiation, since this process
would violate causality. Therefore, we cannot “go back” to a pure state outside the horizon. This can be
formalized by using the strong subadditivity of entanglement entropy, that we will not demonstrate, but
which guarantees:
Saã + SaE ≥ Sa + SaãE . (6)
4
It can be shown that ordinary subadditivity provides the opposite inequality in this scenario, and thus
equality must hold in this relation. Since |Ψaã i is a pure state, then Saã = 0, SaãE = SE , and we obtain the
expected result:
SaE = Sa + SE . (7)
Note that this behavior is very different from an ordinary system. For instance, let us consider a piece of
burning coal and suppose that it is initially in a pure state. The photons first emitted can be thought as
entangled with the particles still in the coal. Then we can imagine that, initially, the entanglement entropy
increases linearly, just as in the black hole case. But when almost half of the coal is already burnt, we can
think of the newly generated photons as encoding the information contained in the coal particles that are
entangled with the radiation emitted earlier. Therefore, pure states will be “recomposed” in the emitted
radiation, and entanglement entropy will start decreasing. When the coal has burnt up, we end up with a
pure state and the entanglement entropy will be 0. This behavior of the entanglement entropy is described
by the Page curve (see Figure 1). The difference between this system and a black hole is that the presence
of the event horizon prevents the information stored in the modes behind the horizon from influencing the
state of the Hawking radiation. Therefore, as we proved, the entanglement entropy of the Hawking radiation,
i.e. of the outside of the black hole, follows the Hawking curve, increasing monotonically until the black
hole disappears. If we want to recover a pure state in the end, an exotic, quantum gravity-driven process
should allow for an extremely steep drop of the entanglement entropy in the last instants of the black hole
evaporation, when the Hawking’s argument is not valid anymore.
The Page curve for an evaporating black hole:
S
Hawking result
SBekenstein-Hawking
Page curve
t
In order for the Hawking radiation to be pure, we
Figure 1: Hawking curve (Von Neumann entropy of the black hole and of the radiation, i.e. entanglement entropy
must deviate from the Hawking calculation
of the two), Bekenstein-Hawking entropy and Page curve. Figure from [8].
already
From Figure around
1, we note thethemidpoint:
also that when evaporation ofan O(1)
the black holeeffect.
is at the halfway point, the
Bekenstein-Hawking entropy becomes smaller than the entanglement entropy of the black hole. This is
another warning of a possible breakdown of ordinary physics. Indeed, the Bekenstein-Hawking entropy is a
thermodynamic entropy, and can therefore be regarded as a “coarse-grained” version of the entanglement
entropy [3]. Since entropy can be viewed as a count of the microstates associated to the black hole macrostate,
logically the Bekenstein-Hawking entropy should be always larger than the entanglement entropy. Therefore,
the situation after the middle point of the evaporation is clearly of difficult interpretation. We remark that,
for the reasons we explained earlier, the Page curve follows the Hawking curve for the first half of the
evaporation process, and then it drops following the Bekenstein-Hawking curve. At the end of the process,
the difference between an ordinary system and a black hole is visually highlighted by the discrepancy between
the Hawking curve and the Page curve.
The fact that entanglement entropy of the Hawking radiation always increases leads again to the paradox
that we pointed out in our first thought experiment: when the evaporation process ends and the black hole
5
disappears, we are left only with Hawking radiation, a mixed state which now represents the whole system.
Therefore, we again started with a pure state and ended up with a mixed state: information has been lost
during the evaporation process.
6
evolution of such a state will always be unitary. Therefore, the evolution of the black hole must be unitary
as well, since it is completely determined by the dual evolution of the CFT state. It is remarkable how the
AdS/CFT correspondence suggests that the black hole information paradox is solved and unitarity is safe,
but it does not explain how this can happen in the gravitational side of the story. All we know is that the
full quantum gravitational theory, whatever it is, must preserve unitarity because the dual CFT does it. It
is a solution without an explanation.
This argument is certainly not satisfying, and its exposition is not detailed, but it is enough to understand
that there are significant hints of the non-violation of unitarity in the black hole evaporation process.
1
Inf(R) = S( I) − S(R)
dR
1 1
= S( I) − S( I)
dR dB
= log dR − log dB
= 2 log dR − log(N )
where N = dB dR = Dimension of the total Hilbert Space containing the black hole and its radiation.
This preliminary analysis suggests that for any information to be retrieved from the Hawking radiation,
one must wait till the half-way point, and from this point onwards, the information increases to its maximum
value log N once the evaporation is complete (no information loss). We will return to this result in the next
section where, relying on some sort of solution of the black hole information paradox, we will assume unitarity
to be preserved.
7
Figure 2: The cloning problem. Figure 3: Experiment to verify cloning.
First of all, we see that if Bob waits for too long outside before jumping in, Alice’s message will not
reach Bob. It is a calculation in general relativity to find a bound on ∆t so that he is just in time to receive
Alice’s message. This bound is ∆t ≤ O(rS log rS ). Next, we ask how long ∆t should be so that Bob can
actually recover |φi from the Hawking Radiation. If BHC is consistent, then we want the experiment to fail,
i.e. ∆t ≥ rS log rS .
Looking at the Page curve discussed in Section 3, it seems that for any information to be available at all
from the Hawking radiation, we need to wait for half the black hole to evaporate, i.e. ∆t is of the order of
the lifetime of the black hole, which goes as O(rS3 ) and is enough to protect BHC. However, by posing the
question more sharply, Preskill and Hayden found that in some scenarios, the relevant time-scale can shift
from the black-hole lifetime to the black-hole thermalization time.
This can be rephrased in terms of an equivalent condition. For one message M that is maximally
entangled with a reference N (which remains outside the black hole), if Alice drops M into the black hole,
Bob should be able to recover a k-qubit subsystem that is maximally entangled with the reference N .
Now, we will argue that this is in turn equivalent to the condition that the post-evaporation black hole
state B 0 is completely uncorrelated with the reference N (Figure 6). To see this, first we note that the joint
state of N and B 0 , σN B 0 , is mixed because of the Hawking Radiation. So, the s qubit system R (the Hawking
Radiation) can beused to purify σN B 0 . We break up R into a k-qubit system M 0 and an s − k-qubit system
Q. So, let trM 0 ,Q |ξN B 0 M 0 Q ihξN B 0 M 0 Q | = σN B 0 . Additionally if we succeed in decoupling N and B 0 , then
8
σN B 0 = σN ⊗ σB 0 . Now, we can find a purifier for σN in M 0 and for σB 0 in Q.
Therefore if N and B 0 are decoupled, then Bob can use the unitary operation given by the equivalence of
purifications to recover a system maximally entangled with N .
We have thus reduced the problem to finding how close σN B 0 is to σN ⊗σB 0 , i.e. finding kσN B 0 − σN ⊗ σB 0 k1 .
Preskill and Hayden calculate this quantity averaged over unitary matrices V chosen uniformly from a Haar
measure on 2n -dimensional unitaries. They make an additional simplification by assuming that the post-
evaporation black hole state σB 0 is maximally mixed. The result that we need to bound this average distance
is Z
dN dBM 2
dV kσN B 0 − σN ⊗ σB 0 k1 ≤ 2 tr (ωBM N ) (8)
dR
where dN = 2k is the dimension of N , and so on, and ωBM N is the joint stateof B, M and
N . Since B was
2
in a pure state to start with, and M N is pure too, ωBM N is pure. Hence, tr (ωBM N ) = 1. Substituting
k n
the numbers into the RHS of the inequality in Equation (8), we get 222s
2 1
= 22s−n−k . This means that if Bob
collects a constant number of qubits more than (n + k)/2, he can recover the message. This is in agreement
with the Page analysis which says that Bob needs to wait for half of the black hole to evaporate.
While we haven’t proved the inequality in Equation (8) here, we would like to present an argument by
Patrick Hayden [10] that motivates it. Going through this argument will lead us to a scenario where the
results are drastically different from the Page analysis!
3. S2 (ρ) = − log tr ρ2
Now, we (naively) think of dimension as a rank and use the above definitions to rewrite the RHS of
Inequality (8) as
2S0 (N )+S0 (BM )−S2 (BM N ) 2−2s (9)
Now, we (very roughly) think of S0 (N ) + S0 (BM ) − S2 (BM N ) as S(N ) + S(BM ) − S(BM N ), which is
the mutual information I(N : BM ) between N and BM . Then Inequality 8 seems to say that the average
9
distance of σN B 0 from being a product state is ≤ 2I(N :BM )−2s . It is easier to make sense of this – the
average distance is a measure of correlation between N and B 0 . All the initial correlation between N and
the black hole is given by I(N : BM ) and this correlation slowly decreases as s increases, i.e. as we throw
away more and more qubits into the Hawking Radiation.
However, we see that something is not right with the above numbers – the mutual information between
N and BM shouldn’t depend on the size of the black hole, but the RHS of Inequality (8) clearly does! This
begs the question – can we modify the scenario so that the RHS is indeed independent of n?
2. Meanwhile Bob collected the early Hawking radiation E that was emitted before Alice threw her qubits
in.
3. Now, Bob tries to recover Alice’s state from E and the new Hawking radiation R.
Figure 7: Modified scenario: Bob collected the early radiation E apart from the later Hawking Radiation R.
Since Alice waited for half the black-hole to evaporate, we can assume that
B is maximally entangled
2
with the early radiation, and hence ρB maximally mixed. Thus, tr (ωBM N ) = 2−(n−k) . Next, we observe
that the rank calculations done above are now correct, i.e. rank ρBM = 2n . Now, we can plug the numbers
back into the RHS of Inequality 8 to get 2k 2n 2−(n−k) 2−2s = 2−2(s−k) .
Z
1
dV kσN B 0 − σN ⊗ σB 0 k1 ≤ 2(s−k) (10)
2
This is drastically different from the Page analysis! This means that just by collecting O(1) qubits, i.e. a
constant number of qubits more than k, Bob can recover Alice’s state. So in this model for a black hole, the
black hole acts as a mirror for any quantum information that is thrown into it after half of it got evaporated!
10
in Section 4 of [1] do this by assuming that the black hole’s qubits are distributed uniformly just outside
(one Planck length away) the horizon. They estimate the thermalization time to be log(rS ).
If this calculation is correct, then the estimate for ∆t is rS log rS , which means that BHC could be safe,
but only precariously so!
5 Conclusions
In this work, we have developed the Black Hole Information Paradox (BHIP) and discussed how black holes
behave as information mirrors in a possible resolution to the BHIP. We saw that there is a thermodynamics
analogy to classical black hole phenomena. Using quantum field theory in curved spacetime of a black hole,
h̄c3
we found that black holes are indeed characterized by the temperature TH = 8πGM . This indicates that black
holes emit Hawking radiation, leading to their ultimate evaporation. When this phenomenon is analyzed
closer, we discover that it takes pure states to mixed states, a violation of unitarity, a fundamental property
of quantum physics.
Resolutions to this paradox have been developed, and numerous are still being explored. One resolution
is provided by the AdS/CFT correspondence, which dictates that unitarity is preserved because the black
hole can be described by a dual conformal field theory. However, if we assume that unitarity is preserved,
we are still in trouble because we find that black holes can effectively clone quantum states. The Black Hole
Complementarity hypothesis tries to resolve this by arguing that cloning is unverifiable. Hayden and Preskill
investigated this issue to check if cloning is actually unverifiable. Using a model in which black holes act as
instantaneous quantum randomizers, they found that black holes act as a mirrors for quantum information.
However, using estimates of black hole thermalization time, they argue how this result is not inconsistent
with black hole complementarity.
Another important problem in this context is the Firewall Paradox. Here, there is an apparent violation of
the Principle of Monogamy of Entanglement, provided that a certain task, called the HH decoding task, can
be done. This task amounts to separating a Bell Pair out of a system, by applying a unitary transformation
just on the part of the system that doesn’t contain the first qubit of the Bell Pair. Harlow and Hayden
showed [11] that the computational hardness of this task is related to a problem in complexity theory on
whether SZK ⊆ BQP (SZK = Statistical Zero Knowledge class). This has opened up research directions
at the interface of BHIP and computer science [12]. In case the HH-decoding task can be performed in
polynomial time, then to prevent the paradox of Monogamy violation, one has to theorize the existence of
firewalls at the horizon – where a smooth spacetime at the horizon is replaced by Planck energy particles!
This would be a marked departure from our current understanding of the event horizon.
In spite of much progress in this field, many questions have yet to be answered. For example, what is
the process which leads to preservation of unitarity in the evolution of black holes? What precisely is the
thermalization time of a black hole? We expect that the physics of quantum information scrambling, often
used to study condensed matter systems, could provide insights into calculating this. To say the least, BHIP
has opened doors to new concepts in general relativity, quantum information and computer science!
References
[1] Patrick Hayden and John Preskill. Black holes as mirrors: quantum information in random subsystems.
http://iopscience.iop.org/article/10.1088/1126-6708/2007/09/120, 2007.
[2] Ted Jacobson. Introductory lectures on black hole thermodynamics. "https://www.physics.umd.edu/
grt/taj/776b/lectures.pdf, 1996.
[3] Joseph Polchinski. The black hole information problem. In New Frontiers in Fields and Strings: TASI
2015 Proceedings of the 2015 Theoretical Advanced Study Institute in Elementary Particle Physics, pages
353–397. 2017.
[4] Ted Jacobson. Introduction to quantum fields in curved spacetime and the hawking effect. https:
//arxiv.org/abs/gr-qc/0308048, 2003.
11
[5] Mark Van Raamsdonk. Lectures on gravity and entanglement. In New Frontiers in Fields and Strings:
TASI 2015 Proceedings of the 2015 Theoretical Advanced Study Institute in Elementary Particle Physics,
pages 297–351. World Scientific, 2017.
[6] Anthony Zee. Einstein gravity in a nutshell, 2013.
[7] Sean Carroll. Spacetime and geometry: An introduction to general relativity, 2004.
[8] Joseph Polchinski. https://physics.princeton.edu/strings2014/slides/polchinski.pdf, 2014.
[9] Leonard Susskind, Larus Thorlacius, and John Uglum. The stretched horizon and black hole comple-
mentarity. Physical Review D, 48(8):3743, 1993.
[10] Patrick Hayden. Lecture at us-india advanced studies institute on thermalization: From glasses to black
holes. https://www.youtube.com/watch?v=qom1jAPlycQ&t=4292s.
[11] Daniel Harlow and Patrick Hayden. Quantum computation vs. firewalls. Journal of High Energy Physics,
2013(6):85, 2013.
[12] Scott Aaronson. The complexity of quantum states and transformations: from quantum money to black
holes. arXiv preprint arXiv:1607.05256, 2016.
[13] Michael Peskin and Daniel Schroeder. An introduction to quantum field theory, 1995.
12
Next, in order to build a QFT, we quantize our classical field theory by promoting φ and π to operators
acting on a Hilbert space that obey the commutation relation [φ(~x, t), π(~x0 , t)] = ih̄δ(~x − ~x0 ). φ(~x, t) and
π(~x, t) are now operators acting on the quantum state of the system, which we denote by |Ψi.
In order to study the action of these operators on a state, one expands φ(~x) into its Fourier modes
(wave-like solutions to the equations of motion) with coefficients ap~ and a†p~ :
d3 p
Z
1
ap~ e−i(ωp~ t−~p·~x) + a†p~ ei(ωp~ t−~p·~x) ,
p
φ(~x, t) = 3
p ωp~ = p|2 + m2 ,
|~ (12)
(2π) 2ωp~
d3 p
Z r
ωp~
ap~ e−i(ωp~ t−~p·~x) − ap†~ ei(ωp~ t−~p·~x) .
π(~x, t) = (13)
(2π)3 2
Once this procedure is performed, one can use the canonical commutation relations to find that ap†~ and ap~ ,
act as creation and annihilation operators, respectively. In particular, these operators act on the state of a
system by creating or annihilating one particle with momentum p~. In addition, the operator that measures
the number of particles with momentum p~ is the number operator Np~ = a†p~ ap~ . The expected number of
particles in the mode p~ is then
hNp~ i = hΨ| Np~ |Ψi . (14)
If we use this procedure to count the number of photons of the quantized electromagnetic field surrounding a
body at temperature T , we discover that hNp~ i is given by equation 2, which describes the radiation emitted
by the body. Therefore, in order to analyze the possibility of black hole radiation, we aim to apply quantum
field theory to the curved spacetime of a black hole, and determine hNp~ i in the quantum state of the black
hole.
d3 p 0
Z
αp~p~0 ap~0 − βp~p~0 bp†~0
bp~ = (16)
(2π)3
13
d3 p0
Z
αp∗~0 p~ ap~0 + βp~∗0 p~ b†p~0
ap~ = 3
(17)
(2π)
Here, αp~p~0 and βp~p~0 are functions that relate the sets of operators. Their specific form is specified by the
system under analysis.
Since different observers have different sets of creation/annihilation operators that are related to each
other by such a transformation, observers in different spacetime regions perceive differing quantum envi-
ronments. This is the essence of the Hawking effect: an infalling observer near the black hole horizon will
perceive the quantum state to be a vacuum, whereas observers far from the black hole perceive the quantum
state to contain particles, emitted as radiation from the black hole.
Γ
hN (ω)i = hΨ| a†p~ ap~ |Ψi = h̄ω . (18)
e k B TH
−1
In this equation Γ is the greybody factor, which quantifies the lower intensity of radiation emitted by the
black hole [4] with respect to the radiation emitted by an ideal black body at temperature TH . This is nearly
the same as equation (2): the number of particles per mode seen by the asymptotic observer is given by a
blackbody spectrum, with temperature TH . We therefore conclude that an observer far away from the black
hole perceives the black hole to emit thermal radiation akin to an object at temperature TH .
Furthermore, one can demonstrate that black hole radiation is similar to pair production, as we noted
previously. For every outgoing particle of radiation, there is another infalling particle, traveling toward the
singularity, which has negative energy and is produced by a creation operator ã†p~ . Such a pair of radiation
particles is known as a Hawking pair. This process is not unphysical, because, as we have already stated,
energy inside the horizon becomes a momentum. ã†p~ is defined as the equivalent of a†p~ , but acting inside the
horizon. It can be shown that the vacuum state of the infalling observer can be written as:
Z ∞
dω − 2Tω † †
|Ψi = N exp e H a ã
p
~ p ~ |0ia ≡ |Ψaã i (19)
0 2π
where |0ia is the vacuum for the asymptotic observer. |Ψaã i is an entangled state. Therefore, each photon
belonging to the Hawking radiation outside the horizon is entangled with a corresponding particle inside the
horizon. This property was fundamental to development of the black hole information problem above.
14