Statistical Mechanics of Strange Metals and Black Holes: Arxiv:2205.02285
Statistical Mechanics of Strange Metals and Black Holes: Arxiv:2205.02285
Statistical Mechanics of Strange Metals and Black Holes: Arxiv:2205.02285
02285
Subir Sachdev
Department of Physics, Harvard University, Cambridge MA-02138, USA
School of Natural Sciences, Institute for Advanced Study, Princeton, NJ-08540, USA and
International Centre for Theoretical Sciences,
Tata Institute of Fundamental Research, Bengaluru 560 089, India
(Dated: October 7, 2022)
arXiv:2205.02285v8 [hep-th] 3 Dec 2022
Abstract
A colloquium style review of the connections between the Sachdev-Ye-Kitaev model and strange metals
without quasiparticles, and between the SYK model and the quantum properties of black holes. Along
with other insights, this connection has led to an understanding of the universal form of the low energy
density of states of charged black holes in asymptotically 3+1 dimensional Minkowski space.
1
The meeting of the American Physical Society in New York in 1987 celebrated the discovery
of high temperature superconductivity in the cuprate series of compounds. I was fortunate to be
present as a young postdoc at this ‘Woodstock’ meeting. Physicists were confidently applying the
till-then highly successful methods of condensed matter physics to these materials, expecting a
rapid explanation of their remarkable properties. But the ensuing decades have shown that new
paradigms of the collective quantum behavior of electrons would be needed, which would take
years to develop. The biggest mystery, as became evident early on, was the unusual metallic
state of these materials, above the superconducting critical temperature. This ‘strange metal’
as it has since come to be called, displayed unusual temperature and frequency dependencies
in its properties, which indicated that the strange metal was an entangled many-body quantum
state without ‘quasiparticles’. Almost all of condensed matter physics is built on the idea of
quasiparticles: it allows us to account for the Coulomb interactions between electrons, by assuming
their main effect is to renormalize each electron with a cloud of electron-hole pairs, after which we
can treat each electron as a nearly independent quasiparticle. This decomposition of the excitations
of a many-body system into a composite of simple quasiparticle excitations is an assumption so
deeply engrained in the theoretical framework that it is usually left unstated.
Complex many-particle quantum entanglement is also a central theme in another major puzzle in
theoretical physics. In 1974, Stephen Hawking [1] combined heretofore distinct pillars of physics:
the quantum theory of microscopic particles like the electron, and Einstein’s theory of general
relativity which applies on astrophysical scales. He argued that the application of quantum theory
to the black hole solutions of general relativity led to the remarkable conclusion that each black
hole had a non-zero temperature, and an associated entropy (which had been postulated earlier by
Bekenstein [2]). Hawking’s arguments were based upon semi-classical methods, related to the old
quantum theory of Bohr and Sommerfeld. It was not at all clear whether Hawking’s results were
compatible with the microscopic quantum theory of Schrödinger and Heisenberg, which is known
today to apply without change to all the microscopic constituents of our universe. Indeed, Hawking
famously stated in the early days that perhaps the Schrödinger-Heisenberg quantum theory broke
down near black holes. Today, there is an emerging consensus that the Schrödinger-Heisenberg
quantum theory is indeed compatible with general relativity and black holes, and complex and
chaotic many-particle quantum entanglement is the key to resolving the difficulties in the semi-
classical description.
The last few decades have seen much progress in our understanding of the remarkable physical
consequences of many-particle quantum entanglement, coming from a synthesis of ideas from
quantum condensed matter theory, quantum information science, quantum field theory, string
theory, and also from modern mathematical ideas built on category theory. Here, I will discuss
some of the insights that have emerged from a study of the Sachdev-Ye-Kitaev (SYK) model. I
2
proposed a closely related model with the same physical properties in 1992, some of which were
described in a paper with my first graduate student, Jinwu Ye [3]. Alexei Kitaev [4] proposed a
modification in 2015 which simplified its solution, and enabled important insights from a more
refined analysis [5–16]. My motivation in 1992 was to write down the simplest model of a metal
without quasiparticles, as a starting point towards addressing the strange metal problem of the
cuprates. Additional properties of the SYK model were described by Olivier Parcollet and Antoine
Georges in Refs. [17, 18] in 1999-2001, and in 2010 I pointed out [19, 20] that the SYK model also
provided a remarkable description of the low temperature properties of certain black holes [21].
This connection has since undergone rapid development and has been made quite precise. The
SYK model shows that the quantum entanglement responsible for the absence of quasiparticles
in strange metals is closely connected to that needed for a microscopic quantum theory of black
holes.
I. FOUNDATIONS BY BOLTZMANN
where S(E) is the thermodynamic entropy in the microcanonical ensemble with extensive energy
E.
Second, in 1872, Boltzmann’s equation gave a correct description of the time evolution of the
observable properties of a dilute gas in response to external forces. He applied Newton’s laws
of motion to individual molecules, and obtained an equation for fp , the density of particles with
momentum p. In a spatially uniform situation, Boltzmann’s equation takes the form
∂fp
+ F · ∇p fp = C[f ] , (3)
∂t
3
FIG. 1: Collision between two molecules. The collision term in the Boltzmann equation is
proportional to the absolute square of the T -matrix.
where t is time, and F is the external force. The left-hand-side of (3) is just a restatement of
Newton’s laws for individual molecules. Boltzmann’s innovation was the right-hand-side, which
describes collisions between the molecules. Boltzmann introduced the concept of ‘molecular chaos’,
which asserted that in a sufficiently dilute gas successive collisions were statistically independent.
With this assumption, Boltzmann showed that
Z
C[f ] ∝ · · · [fp fp1 − fp2 fp3 ] (4)
p1,2,3
for a collision between molecules as shown in Fig. 1. The statistical independence of collisions is
reflected in the products of the densities in (4), and the second term represents the time-reversed
collision.
The remarkable fact is that Boltzmann’s equation also applies, with relatively minor modifica-
tions, in situations very different from the dilute classical gas: it also applies to the dense quantum
gas of electrons found in ordinary metals. Now collisions become rare because of Pauli’s exclusion
principle, and the statistical independence of collisions is assumed to continue to apply. The main
modification is that the collision term in (4) is replaced by
Z
C[f ] ∝ · · · [fp fp1 (1 − fp2 )(1 − fp3 ) − fp2 fp3 (1 − fp )(1 − fp1 )] , (5)
p1,2,3
where the additional (1 − f ) factors ensure that the final states of collisions are not occupied.
Now the fp measure the distribution of electronic quasiparticles, and the cloud of particle-hole
pairs around each electron only renormalizes the microscopic scattering cross-section. Such a
quantum Boltzmann equation is the foundation of the quasiparticle theory of the electron gas
in metals, superconductors, semiconductors, and insulators, and indeed almost all of condensed
matter physics before the 1980’s. One of its important predictions is that as temperature T → 0,
the typical time between collisions, tc , diverges as tc ∼ 1/T 2 .
4
One can now ask, how short can we make tc before we cannot ignore the quantum interference
between successive collisions, and the concept of quasiparticles does not make sense? An energy-
time uncertainty-principle argument indicates that any many-body quantum system should have
a relaxation time [22]
~
tr ≥ α , T → 0, (6)
kB T
where α is a dimensionless, T -independent constant. For systems with quasiparticles, we expect
tc ∼ tr , and we have introduced a general relaxation time tr to allow a more general discussion in
systems without quasiparticles. From studies of various quantum critical systems, it was argued
[22] that the inequality in (6) becomes an equality when quasiparticles are absent, as in strange
metals. Recent experiments [23] on the strange metal in cuprate superconductors have measured a
particular relaxation time by connecting it to the angle dependence of the resistivity in a magnetic
field, and indeed found it obeys (6) as an equality, with α ≈ 1.2. This is often stated as the strange
metal exhibiting ‘Planckian time’ dynamics [24].
We can write the quantum theory of black holes as a Feynman path integral over the spacetime
metric gµν , and the electromagnetic gauge field aµ : this involves computing the partition function
Z ~/(kB T ) !
√
Z Z
1
Z = Dgµν Daµ exp − dd x dτ g Ld [gµν , aµ ] (7)
~ 0
In imaginary time, the spacetime geometry outside a black hole is that of a ‘cigar’ as shown in
Fig. 2. Note that all dependence of (7) on ~ and T is explicit, and there is no ~ or T in the
Lagrangian Ld .
Formally, the path integral in (7) is pathological because it includes infinities that cannot be
controlled by the usual renormalization tricks of quantum field theory. Nevertheless, Gibbons and
Hawking [25] boldly decided to evaluate it in the semi-classical limit, where one only includes
5
FIG. 2: Spacetime geometry outside a black hole. Only the radial direction and the imaginary
time direction τ are shown, and the two angular directions are not shown.
the contribution of the cigar saddle point in Fig. 2. For a black hole, they also imposed the
requirement that spacetime was smooth at the horizon in imaginary time. From this relatively
simple computation, they were able to obtain the thermodynamic properties of a black hole,
including its temperature and entropy. For a neutral black hole of mass M in d = 3 they found
S c3 A kB T c3
= , = (9)
kB 4~G ~ 8πGM
where c is the velocity of light, G is Newton’s gravitational constant, and A = 4πR2 is the area of
the black hole horizon with R = 2GM/c2 the horizon radius.
The revolutionary results in (9) raised many more questions than they answered. Is this semi-
classical computation of thermodynamics compatible with Boltzmann’s fundamental statistical
interpretation of entropy in (2)? How does a computation in imaginary time outside a black hole
know about the entropy of quantum degrees of freedom inside a black hole? Can one compute the
energy eigenvalues of a quantum Hamiltonian describing the inside of a black hole whose density
of states D(E) yields a S(E) that is consistent with (9), and the partition function Z in (7)? With
the energy E shifted so that E = 0 for the ground state, Z is related to D(E) by
Z ∞
E
Z= dED(E) exp − . (10)
0− kB T
Many other questions are raised when one considers the fate of the black hole as it evaporates while
emitting blackbody radiation at the temperature in (9), and computes the entanglement entropy
of the Hawking radiation.
A remarkable feature of the entropy in (9) is that it is proportional to the surface area of the
black hole. This contrasts with extensive volume proportionality of the entropy, mentioned below
(1), obeyed by all other quantum systems. Attempts to understand this feature led to the idea
6
FIG. 3: Holography: the number of qubits required for the quantum simulation of a black hole is
proportional to its surface area.
of holography [26–28] illustrated in Fig. 3. Let us try to build a quantum simulation of a black
hole by simple two-level systems i.e. qubits. How many qubits will we need? As N qubits can
describe 2N linearly-independent quantum states, the formula in (2) tells us that the number of
qubits is proportional to the entropy, and hence the area of the black hole. So the qubits realize
a many-body quantum system which we can imagine residing on its surface i.e. the qubits are a
faithful (d − 1)-dimensional hologram of the complete quantum gravitational theory of the black
hole in d spatial dimensions.
A specific realization of a quantum simulation was found in string theory for ‘extremal’ black
holes (defined below) with low energy supersymmetry [29]. This realization has a ground state
with an exponentially large degeneracy, and special features of the supersymmetry were employed
to compute this degeneracy, yielding
where . . . refers to a continuum above an energy gap. The value of S(0) was found to be precisely
that in the Hawking formula in (9) [30]. However, the zero energy delta-function in (11) is known
[16, 31] to be a special feature of theories with low energy supersymmetry, and is not a property
of the generic semi-classical path integral over Einstein gravity in (7), as we will discuss below (see
Fig. 5).
To move beyond supersymmetric string theory, we ask if there are any general constraints
that must be obeyed by the many-body system realized by the interactions between the qubits.
An important constraint comes from an earlier result by Vishveshwara [32]. He computed the
relaxation time, tr of quasi-normal modes of black holes in Einstein’s classical theory; this is
7
the time in which a black hole relaxes exponentially back to a spherical shape after it has been
perturbed by another body:
8πGM
tr = α0 , (12)
c3
where α0 is a numerical constant of order unity dependent upon the precise quasi-normal mode.
Comparing Vishveshwara’s result in (12) with Hawking’s result in (9), we can write
~
tr = α0 (13)
kB T
which is exactly of the form in (6) for many body quantum systems without quasiparticles! This is
a key hint that the holographic quantum simulation of a black hole must involve a quantum system
without quasiparticle excitations, if it is reproduce basic known features of black hole dynamics.
At this point, it is interesting to note that measurements of tr in binary black hole mergers by
LIGO-Virgo [33] do indeed fall around the value of ~/(kB T ).
The Hamiltonian of a version of a SYK model is illustrated in Fig. 4. We take a system with
fermions ψi , i = 1 . . . N states. Depending upon physical realizations, the label i could be position
or an orbital, and it is best to just think of it as an abstract label of a fermionic qubit with the
two states |0i and ψi† |0i. We now place QN fermions in these states, so that a density Q ≈ 1/2
is occupied, as shown in Fig. 4. The quantum dynamics is restricted to only have a ‘collision’
term between the fermions, analogous to the right-hand-side of the Boltzmann equation. However,
FIG. 4: The SYK model: fermions undergo the transition (‘collision’) shown with quantum
amplitude Uij;k` .
8
in stark contrast to the Boltzmann equation, we will not make the assumption of statistically
independent collisions, and will account for quantum interference between successive collisions:
this is the key to building up a many-body state with non-trivial entanglement. So a collision in
which fermions move from sites i and j to sites k and ` is characterized not by a probability, but
by a quantum amplitude Uij;k` , which is a complex number.
The model so defined has a Hilbert space of order 2N states, and a Hamiltonian determined by
order N 4 numbers Uij;k` . Determining the spectrum or dynamics of such a Hamiltonian for large
N seems like an impossibly formidable task. But if we now make the assumption that the Uij;k` are
statistically independent random numbers, remarkable progress is possible. Note that we are not
considering an ensemble of SYK models with different Uij;k` , but a single fixed set of Uij;k` . Most
physical properties of this model are self-averaging at large N , and so as a technical tool, we can
rapidly obtain them by computations on an ensemble of random Uij;k` . In any case, the analytic
results we now describe have been checked by numerical computations on a computer for a fixed
set of Uij;k` . We recall that even for the Boltzmann equation, there was an ensemble average over
the initial positions and momenta of the molecules that was implicitly performed.
Using these methods, key properties of the SYK model have been established (for complete
references to the literature, please see the review in Ref. [34]):
• There are no quasiparticle excitations, and it exhibits quantum dynamics with a Planckian
relaxation time obeying (13) at T U , where U/N 3/2 is the root-mean-square value of the
Uijk` . In particular, the relaxation time is independent of U , a feature not present in any
ordinary metal with quasiparticles.
The limit limT →0 limN →∞ S(T )/(kB N ) = s0 is non-zero, implying an energy level spacing
exponentially small in N near the ground state. This is very different from systems with
quasiparticle excitations, whose energy level spacing vanishes with a positive power of 1/N
9
FIG. 5: (a) Plot of the 65536 many-body eigenvalues of a N = 32 Majorana SYK Hamiltonian;
however, the analytical results quoted here are for the SYK model with complex fermions which
has a similar spectrum. The coarse-grained low energy and low temperature behavior is described
by (14) and (15). The lower energy part of this density of states is argued to apply to the
semi-classical path integral in (7) for charged black holes at low T (i.e. extremal black holes). (b)
Schematic of the lower energy density of states of the supersymmetric string theory description of
extremal black holes [16] in (11). The energy gap ∆ is proportional to the inverse of S(E = 0).
near the ground state. However, there is no exponentially large degeneracy of the ground
state itself in the SYK model, unlike the ground states of the string theory solutions leading
to (11), and the ground states in Pauling’s model of ice [36]. Obtaining the ground state
degeneracy requires the opposite order of limits between T and N , and numerical studies show
that the entropy density does vanish in such a limit for the SYK model. The density of states
(14) implies that any small energy interval near the ground state contains an exponentially
large number of energy eigenstates with an exponentially small spacing in energy (see Fig. 5a).
The wavefunctions of these eigenstates in Fock space change chaotically from one state to the
next, providing a realization of maximal many-body quantum chaos [37] in a precise sense.
This structure of eigenstates is very different from systems with quasiparticles, for which the
lowest energy eigenstates differ only by adding and removing a few quasiparticles.
• The E dependence of the density of states in (14) is associated with a time reparameterization
mode, and (14) shows that its effects are important when E ∼ 1/N . We can express the low
energy quantum fluctuations in terms of a path integral which reparameterizes imaginary
time τ → f (τ ), in a manner analogous to the quantum theory of gravity being expressed in
10
terms of the fluctuations of the spacetime metric. There are also quantum fluctuations of a
phase mode φ(τ ), whose time derivative is the charge density, and so we have the partition
function !
Z Z ~/(kB T )
1
ZSY K = eN s0 Df Dφ exp − dτ LSY K [f, φ] (16)
~ 0
The Lagrangian LSY K is known, and involves a Schwarzian of f (τ ). Remarkably, despite its
non-quadratic Lagrangian, the path integral in (16) can be performed exactly [12], and leads
to (14).
Can we use insights from the path integral over time reparameterizations of the SYK model in
(16) to evaluate the path integral over spacetime metrics of black holes in (7)? Remarkably, for a
black hole with a non-zero fixed total charge Q, the answer is yes (for complete references to the
literature in the discussion below, please see the review in Ref. [34]).
The saddle-point solution of the Einstein-Maxwell action for a charged black hole has the form
shown in Fig. 6: while the spacetime is 3+1 dimensional flat Minkowski far from the black hole, it
factorizes into a 1+1 dimensional spacetime involving the radial direction ζ, and a 2-dimensional
space of non-zero angular momentum modes around the spherical black hole. As the black hole
temperature T → 0 (also known as the extremal limit), the non-zero angular momentum modes
become unimportant, and we can write the partition function of the charged black hole purely as a
theory of quantum gravity in 1+1 spacetime dimensions, which is an extension of a theory known
as Jackiw-Teitelboim (JT) gravity; then (7) reduces to
!
~/(kB T )
√
Z Z Z
A0 c3 /(4~G) 1
ZJT = e Dgµν Daµ exp − dζ dτ g L1 [gµν , aµ ] , (17)
~ 0
where A0 = 2GQ2 /c4 is the area of the black hole horizon at T = 0. The (1 + 1)-dimensional
spacetime saddle point of ZJT has a uniform negative curvature: it is the anti-de Sitter space
AdS2 , noted in Fig. 6. Quantum gravity in 1+1 dimensions is especially simple because there is
no graviton, and it is possible to make an explicit holographic mapping to a quantum system in
0+1 dimensions. It turns out that the holographic quantum realization of the 1+1 dimensional
theory ZJT in (17) is exactly the 0 + 1 dimensional SYK model partition function in ZSY K in
(16). The fluctuations of the metric in the boundary region between the 1+1 dimensional and
3+1 dimensional spacetimes (denoted ‘boundary graviton’ in Fig. 6) are described by the time
reparameterization f (τ ), and the boundary value of aµ becomes the phase field φ(τ ). This powerful
connection enables us to proceed beyond the semi-classical results of Hawking in (9) for black holes
11
FIG. 6: Two views of the spacetime outside a charged black hole. S 2 is the sphere, and AdS2 is
anti-de Sitter space with metric ds2 = (dζ 2 + dτ 2 )/ζ 2 .
with non-zero charge Q. Applying this mapping from ZJT to ZSY K , we obtain a density of states
D(E) with precisely the E dependence in (14), which also corresponds to that shown for the SYK
model in Fig. 5a. This should be contrasted with the supersymmetric result in (11) and Fig. 5b
[15, 16]. The parameters in the black hole D(E) can be deduced by comparing the SYK entropy in
(15) with the low T limit of the entropy of a charged black hole in asymptotically 3+1 dimensional
Minkowski spacetime, which is
c3 √ 3/2 kB T
5
(~c /G)1/2
S(T ) 3
= A0 + 2 πA0 − ln + ... . (18)
kB 4~G ~c 2 kB T
The first two terms in (18) correspond to Hawking’s area law in (9), with an horizon area A which
increases linearly with T from its T = 0 value A0 . The linear T dependence can also be obtained
from the saddle-point action of ZJT in (17). These first two terms in (18) are in one-to-one
correspondence with the first two terms in (15) for the SYK model. Indeed the correspondence
extends also to the third terms in (18) and (15), both of which are obtained by performing the
equivalent path integrals in ZJT and ZSY K . These terms are obtained in an expansion in small
~G for the black hole, in place of large N for the SYK model. There is also a T -independent term,
(−559/180) ln(A0 c3 /(~G)), arising from other massless modes of the 3 + 1 dimensional black hole,
not included in (18) [15, 38, 39]; this analogous to the ln N term in (15), but the co-efficients of
the logarithms differ because the matter content of the two theories differ. In this manner, we
12
obtain the expression for the universal form of the low energy density of states of charged black
holes in asymptotically 3+1 dimensional Minkowski spacetime, which is the analog of (14) for the
SYK model −347/90 1/2 !
A 0 c3 A0 c3 √ 3/2 c3 E
D(E) ∼ exp sinh πA0 . (19)
~G 4~G ~G ~c
This is very different from the result in (11) obtained from string theory. Recall that A0 is the
area of the black hole horizon at T = 0; all other parameters in (19) are fundamental constants
of nature. The result in (19) is a rare formula which combines Planck’s constant ~ with Newton’s
gravitational constant G: the exponential prefactor was obtained by Hawking, and the sinh follows
from developments ensuing from the solution of the SYK model. In particular, the time reparam-
eterization mode is important when the sinh in (19) becomes of order unity, and such corrections
address [15, 16] issues raised in Ref. [40].
We have now answered one of the questions raised by Hawking’s semi-classical computation of
black hole entropy, at least for the case of a charged black hole at low T . There is a perfectly
well-defined quantum system, with precise and discrete energy levels, whose many body density
of states D(E) is described at low energies by a gravitational theory expressed in terms of the
semi-classical fluctuations of a spacetime metric: this is the SYK model. This connection has
enabled computation of the logarithmic correction to black hole entropy in (18) and the density of
states in (19), and has led to some understanding of the key role played by many-body quantum
chaos in the black hole microstates [41]. The random couplings in the SYK model mainly play
the role of a powerful computational tool for accessing the physics of chaotic behavior, as was
also the case in single-particle problems such as ‘quantum billiards’ [42]. The quantities being
studied here self-average in a single realization of the random couplings, and are not sensitive to
the particular member chosen from the random ensemble. It is also worth noting here that the
Boltzmann equation also involves an implicit average over an ensemble of initial conditions for the
quasiparticles.
The connection between the SYK model and black holes does not imply that the ultimate
high energy and short distance physics is described by the SYK model. That likely requires
string theory, in which supersymmetry is restored at high energies. Nevertheless, the SYK model
provides a much simpler quantum simulation of the low energy physics in certain cases, including
the complex quantum entanglement and the maximal many-body quantum chaos. The semi-
classical path integral over Einstein gravity can capture certain coarse-grained properties of the
underlying Schrödinger-Heisenberg quantum theory, and is not sensitive to all the microscopic
details at the smallest length scales. Recent work [43, 44] has shown that the path integral over
Einstein gravity can also consistently describe the time evolution of the entanglement entropy of
an evaporating black hole, provided we also include the contributions of ‘wormhole’ solutions of
13
Einstein gravity. Such wormholes can also be realized by the quantum states of the SYK model.
The idea that Einstein quantum gravity is a coarse-grained description of the microscopics has also
led to many theoretical advances by employing averages over ensembles of consistent, microscopic,
quantum theories [45].
To conclude, we return to the original motivation for studying the SYK model, to the strange
metal phase of the cuprate superconductors and related compounds. Realistic models of the lattice
scale physics of these materials involve mobile electrons with strong interactions. These models
can be solved by methods closely related to those of the SYK model, after they are extended to
large spatial dimension d, and upon including random spin exchange interactions. These large d
models display strange metal phases with many similarities to observations [34]. In particular,
such strange metals do have a time reparameterization low energy mode, and this mode leads to a
linear-in-temperature resistivity at the smallest T [46], as is observed in the strange metals of the
cuprates.
Other works have focused on the strange metal directly in d = 2 dimensions, which is the
physically relevant dimension for the cuprates. Rather than departing from the SYK model, phys-
ically realistic models of such strange metals can be built from ‘Yukawa-SYK’ models of fermions
and bosons with Yukawa couplings [34]. An important consideration is the role of the Fermi sur-
face: for free electrons, this is the surface which divides occupied and empty electron states in
momentum space. A suitably defined Fermi surface is also present when there are interactions
between the electrons in an ordinary metal, and also when quasiparticles are no longer present in
a strange metal. An interesting and much studied strange metal is obtained when the interactions
between the Fermi surface excitations are mediated by the exchange of a boson with a gapless
energy spectrum. Such a boson may be the order parameter of a symmetry-breaking quantum
phase transition, or an emergent gauge field in a correlated metal. However, this strange metal has
a zero resistance because the low energy theory has a continuous translation symmetry. But recent
studies [47–50] have shown that adding a SYK-like randomness to the boson-electron Yukawa cou-
pling does lead to a linear-in-temperature resistivity as T → 0. Such randomness can arise from
impurities in the strange metal, and quantifying the role of randomness is an important direction
for the further study of strange metals.
14
Acknowledgements
I thank Matthew Heydeman, G. Joaquin Turiaci and Spenta Wadia for important comments
on the manuscript. This research has been supported by the U.S. National Science Foundation,
most recently under grant No. DMR-2002850.
[1] S. W. Hawking, Particle Creation by Black Holes, Commun. Math. Phys. 43, 199 (1975).
[2] J. D. Bekenstein, Black holes and entropy, Phys. Rev. D 7, 2333 (1973).
[3] S. Sachdev and J. Ye, Gapless spin-fluid ground state in a random quantum Heisenberg magnet, Phys.
Rev. Lett. 70, 3339 (1993), arXiv:cond-mat/9212030 [cond-mat] .
[4] A. Kitaev, A simple model of quantum holography, talk given at KITP program: entanglement in
Strongly-Correlated Quantum Matter, University of California, Santa Barbara (2015).
[5] S. Sachdev, Bekenstein-Hawking Entropy and Strange Metals, Phys. Rev. X 5, 041025 (2015),
arXiv:1506.05111 [hep-th] .
[6] A. Kitaev and S. J. Suh, The soft mode in the Sachdev-Ye-Kitaev model and its gravity dual, Journal
of High Energy Physics 05, 183 (2018), arXiv:1711.08467 [hep-th] .
[7] J. Maldacena and D. Stanford, Remarks on the Sachdev-Ye-Kitaev model, Phys. Rev. D 94, 106002
(2016), arXiv:1604.07818 [hep-th] .
[8] J. Maldacena, D. Stanford, and Z. Yang, Conformal symmetry and its breaking in two dimensional
Nearly Anti-de-Sitter space, Prog. Theor. Exp. Phys. 2016, 12C104 (2016), arXiv:1606.01857 [hep-th]
.
[9] W. Fu, D. Gaiotto, J. Maldacena, and S. Sachdev, Supersymmetric Sachdev-Ye-Kitaev models, Phys.
Rev. D 95, 026009 (2017), [Addendum: Phys.Rev.D 95, 069904 (2017)], arXiv:1610.08917 [hep-th] .
[10] J. S. Cotler, G. Gur-Ari, M. Hanada, J. Polchinski, P. Saad, S. H. Shenker, D. Stanford, A. Streicher,
and M. Tezuka, Black Holes and Random Matrices, JHEP 05, 118, [Erratum: JHEP 09, 002 (2018)],
arXiv:1611.04650 [hep-th] .
[11] D. Bagrets, A. Altland, and A. Kamenev, Power-law out of time order correlation functions in the
SYK model, Nucl. Phys. B 921, 727 (2017), arXiv:1702.08902 [cond-mat.str-el] .
[12] D. Stanford and E. Witten, Fermionic Localization of the Schwarzian Theory, Journal of High Energy
Physics 10, 008 (2017), arXiv:1703.04612 [hep-th] .
[13] U. Moitra, S. P. Trivedi, and V. Vishal, Extremal and near-extremal black holes and near-CFT1 ,
Journal of High Energy Physics 07, 055 (2019), arXiv:1808.08239 [hep-th] .
[14] S. Sachdev, Universal low temperature theory of charged black holes with AdS2 horizons, J. Math.
Phys. 60, 052303 (2019), arXiv:1902.04078 [hep-th] .
15
[15] L. V. Iliesiu and G. J. Turiaci, The statistical mechanics of near-extremal black holes, Journal of High
Energy Physics 05, 145 (2021), arXiv:2003.02860 [hep-th] .
[16] M. Heydeman, L. V. Iliesiu, G. J. Turiaci, and W. Zhao, The statistical mechanics of near-BPS black
holes, J. Phys. A 55, 014004 (2022), arXiv:2011.01953 [hep-th] .
[17] O. Parcollet and A. Georges, Non-Fermi-liquid regime of a doped Mott insulator, Phys. Rev. B 59,
5341 (1999), arXiv:cond-mat/9806119 [cond-mat.str-el] .
[18] A. Georges, O. Parcollet, and S. Sachdev, Quantum fluctuations of a nearly critical Heisenberg spin
glass, Phys. Rev. B 63, 134406 (2001), arXiv:cond-mat/0009388 [cond-mat.str-el] .
[19] S. Sachdev, Holographic metals and the fractionalized Fermi liquid, Phys. Rev. Lett. 105, 151602
(2010), arXiv:1006.3794 [hep-th] .
[20] S. Sachdev, Strange metals and the AdS/CFT correspondence, J. Stat. Mech. 1011, P11022 (2010),
arXiv:1010.0682 [cond-mat.str-el] .
[21] J. McGreevy, In pursuit of a nameless metal, Physics 3, 83 (2010).
[22] S. Sachdev, Quantum Phase Transitions, 1st ed. (Cambridge University Press, Cambridge, UK, 1999).
[23] G. Grissonnanche, Y. Fang, A. Legros, S. Verret, F. Laliberté, C. Collignon, J. Zhou, D. Graf,
P. A. Goddard, L. Taillefer, and B. J. Ramshaw, Linear-in temperature resistivity from an isotropic
Planckian scattering rate, Nature 595, 667 (2021), arXiv:2011.13054 [cond-mat.str-el] .
[24] S. A. Hartnoll and A. P. Mackenzie, Planckian Dissipation in Metals, (2021), arXiv:2107.07802
[cond-mat.str-el] .
[25] G. W. Gibbons and S. W. Hawking, Action integrals and partition functions in quantum gravity,
Phys. Rev. D 15, 2752 (1977).
[26] G. ’t Hooft, Dimensional reduction in quantum gravity, Conf. Proc. C 930308, 284 (1993), arXiv:gr-
qc/9310026 .
[27] L. Susskind, L. Thorlacius, and J. Uglum, The Stretched horizon and black hole complementarity,
Phys. Rev. D 48, 3743 (1993), arXiv:hep-th/9306069 .
[28] J. M. Maldacena, The Large N limit of superconformal field theories and supergravity, Adv. Theor.
Math. Phys. 2, 231 (1998), arXiv:hep-th/9711200 .
[29] J. R. David, G. Mandal, and S. R. Wadia, Microscopic formulation of black holes in string theory,
Physics Reports 369, 549 (2002), arXiv:hep-th/0203048 .
[30] A. Strominger and C. Vafa, Microscopic origin of the Bekenstein-Hawking entropy, Phys. Lett. B
379, 99 (1996), arXiv:hep-th/9601029 .
[31] L. V. Iliesiu, S. Murthy, and G. J. Turiaci, Black hole microstate counting from the gravitational path
integral, (2022), arXiv:2209.13602 [hep-th] .
[32] C. V. Vishveshwara, Scattering of Gravitational Radiation by a Schwarzschild Black-hole, Nature
227, 936 (1970).
16
[33] G. Carullo, D. Laghi, J. Veitch, and W. Del Pozzo, Bekenstein-Hod Universal Bound on Information
Emission Rate Is Obeyed by LIGO-Virgo Binary Black Hole Remnants, Phys. Rev. Lett. 126, 161102
(2021).
[34] D. Chowdhury, A. Georges, O. Parcollet, and S. Sachdev, Sachdev-Ye-Kitaev models and beyond:
Window into non-Fermi liquids, Rev. Mod. Phys. 94, 035004 (2022), arXiv:2109.05037 [cond-mat.str-
el] .
[35] Y. Gu, A. Kitaev, S. Sachdev, and G. Tarnopolsky, Notes on the complex Sachdev-Ye-Kitaev model,
Journal of High Energy Physics 02, 157 (2020), arXiv:1910.14099 [hep-th] .
[36] L. Pauling, The Structure and Entropy of Ice and of Other Crystals with Some Randomness of Atomic
Arrangement, Journal of the American Chemical Society 57, 2680 (1935).
[37] J. Maldacena, S. H. Shenker, and D. Stanford, A bound on chaos, Journal of High Energy Physics
2016, 106 (2016).
[38] A. Sen, Logarithmic Corrections to N = 2 Black Hole Entropy: An Infrared Window into the Mi-
crostates, Gen. Rel. Grav. 44, 1207 (2012), arXiv:1108.3842 [hep-th] .
[39] L. V. Iliesiu, S. Murthy, and G. J. Turiaci, Revisiting the Logarithmic Corrections to the Black Hole
Entropy, (2022), arXiv:2209.13608 [hep-th] .
[40] J. Preskill, P. Schwarz, A. Shapere, S. Trivedi, and F. Wilczek, Limitations on the statistical descrip-
tion of black holes, Modern Physics Letters A 06, 2353 (1991).
[41] S. H. Shenker and D. Stanford, Black holes and the butterfly effect, Journal of High Energy Physics
03, 067 (2014), arXiv:1306.0622 [hep-th] .
[42] O. Bohigas, M. J. Giannoni, and C. Schmit, Characterization of chaotic quantum spectra and uni-
versality of level fluctuation laws, Phys. Rev. Lett. 52, 1 (1984).
[43] A. Almheiri, T. Hartman, J. Maldacena, E. Shaghoulian, and A. Tajdini, The entropy of Hawking
radiation, Rev. Mod. Phys. 93, 035002 (2021), arXiv:2006.06872 [hep-th] .
[44] R. Bousso, X. Dong, N. Engelhardt, T. Faulkner, T. Hartman, S. H. Shenker, and D. Stanford,
Snowmass White Paper: Quantum Aspects of Black Holes and the Emergence of Spacetime, (2022),
arXiv:2201.03096 [hep-th] .
[45] J. Chandra, S. Collier, T. Hartman, and A. Maloney, Semiclassical 3D gravity as an average of large-c
CFTs, (2022), arXiv:2203.06511 [hep-th] .
[46] H. Guo, Y. Gu, and S. Sachdev, Linear in temperature resistivity in the limit of zero temperature
from the time reparameterization soft mode, Annals Phys. 418, 168202 (2020), arXiv:2004.05182
[cond-mat.str-el] .
[47] E. E. Aldape, T. Cookmeyer, A. A. Patel, and E. Altman, Solvable Theory of a Strange Metal at the
Breakdown of a Heavy Fermi Liquid, (2020), arXiv:2012.00763 [cond-mat.str-el] .
[48] I. Esterlis, H. Guo, A. A. Patel, and S. Sachdev, Large N theory of critical Fermi surfaces, Phys.
17
Rev. B 103, 235129 (2021), arXiv:2103.08615 [cond-mat.str-el] .
[49] H. Guo, A. A. Patel, I. Esterlis, and S. Sachdev, Large-N theory of critical Fermi surfaces. II.
Conductivity, Phys. Rev. B 106, 115151 (2022), arXiv:2207.08841 [cond-mat.str-el] .
[50] A. A. Patel, H. Guo, I. Esterlis, and S. Sachdev, Universal T -linear resistivity in two-dimensional
quantum-critical metals from spatially random interactions, (2022), arXiv:2203.04990 [cond-mat.str-
el] .
18