128
Entanglement, holography, and the quantum phases of matter HARVARD Bethe Colloquium, Bonn, January 14, 2013 Subir Sachdev Wednesday, January 16, 13

Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Entanglement,holography,

andthe quantum phases of matter

HARVARD

Bethe Colloquium, Bonn, January 14, 2013

Subir Sachdev

Wednesday, January 16, 13

Page 2: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Sommerfeld-Bloch theory of metals, insulators, and superconductors:many-electron quantum states are adiabatically

connected to independent electron states

Wednesday, January 16, 13

Page 3: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Modern phases of quantum matterNot adiabatically connected

to independent electron states:many-particle

quantum entanglement

Wednesday, January 16, 13

Page 4: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

| i ) Ground state of entire system,

⇢ = | ih |

⇢A = TrB⇢ = density matrix of region A

Entanglement entropy SE = �Tr (⇢A ln ⇢A)

B

Entanglement entropy

A P

Wednesday, January 16, 13

Page 5: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

| i ) Ground state of entire system,

⇢ = | ih |

Take | i = 1p2(|"iA |#iB � |#iA |"iB)

Then ⇢A = TrB⇢ = density matrix of region A=

12 (|"iA h"|A + |#iA h#|A)

Entanglement entropy SE = �Tr (⇢A ln ⇢A)= ln 2

Entanglement entropy

Wednesday, January 16, 13

Page 6: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Gapped quantum matter Spin liquids, quantum Hall states

Conformal quantum matter Quantum critical points in antiferromagnets, superconductors, and ultracold atoms; graphene

Compressible quantum matter Strange metals in high temperature superconductors, Bose metals

“Complex entangled” states of quantum matter,

not adiabatically connected to independent particle states

S. Sachdev, 100th anniversary Solvay conference,arXiv:1203.4565Wednesday, January 16, 13

Page 7: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Gapped quantum matter Spin liquids, quantum Hall states

Conformal quantum matter Quantum critical points in antiferromagnets, superconductors, and ultracold atoms; graphene

Compressible quantum matter Strange metals in high temperature superconductors, Bose metals

S. Sachdev, 100th anniversary Solvay conference,arXiv:1203.4565

“Complex entangled” states of quantum matter,

not adiabatically connected to independent particle states

Wednesday, January 16, 13

Page 8: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Band insulators

E

MetalMetal

carryinga current

InsulatorSuperconductor

kAn even number of electrons per unit cell

Entanglement entropy of a band insulator

Wednesday, January 16, 13

Page 9: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

SE = aP � b exp(�cP )

where P is the surface area (perimeter)

of the boundary between A and B.

B

Entanglement entropy of a band insulator

A P

Wednesday, January 16, 13

Page 10: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

H = J�

�ij⇥

⌅Si · ⌅Sj =

Mott insulator: Kagome antiferromagnet

P. Fazekas and P. W. Anderson, Philos. Mag. 30, 23 (1974).

Wednesday, January 16, 13

Page 11: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

H = J�

�ij⇥

⌅Si · ⌅Sj =

Mott insulator: Kagome antiferromagnet

Kekule (1865)

Wednesday, January 16, 13

Page 12: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

H = J�

�ij⇥

⌅Si · ⌅Sj =

Mott insulator: Kagome antiferromagnet

Wednesday, January 16, 13

Page 13: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

H = J�

�ij⇥

⌅Si · ⌅Sj =

Mott insulator: Kagome antiferromagnet

Wednesday, January 16, 13

Page 14: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

H = J�

�ij⇥

⌅Si · ⌅Sj =

Mott insulator: Kagome antiferromagnet

Wednesday, January 16, 13

Page 15: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

H = J�

�ij⇥

⌅Si · ⌅Sj =

Mott insulator: Kagome antiferromagnet

Wednesday, January 16, 13

Page 16: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

H = J�

�ij⇥

⌅Si · ⌅Sj =

Mott insulator: Kagome antiferromagnet

Wednesday, January 16, 13

Page 17: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Mott insulator: Kagome antiferromagnet

Alternative view Pick a reference configuration

Wednesday, January 16, 13

Page 18: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Mott insulator: Kagome antiferromagnet

Alternative view A nearby configuration

Wednesday, January 16, 13

Page 19: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Mott insulator: Kagome antiferromagnet

Alternative view Difference: a closed loop

Wednesday, January 16, 13

Page 20: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Mott insulator: Kagome antiferromagnet

Alternative view Ground state: sum over closed loops

Wednesday, January 16, 13

Page 21: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Mott insulator: Kagome antiferromagnet

Alternative view Ground state: sum over closed loops

Wednesday, January 16, 13

Page 22: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Mott insulator: Kagome antiferromagnet

Alternative view Ground state: sum over closed loops

Wednesday, January 16, 13

Page 23: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Mott insulator: Kagome antiferromagnet

Alternative view Ground state: sum over closed loops

Wednesday, January 16, 13

Page 24: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

B

Entanglement in the Z2 spin liquid

A

The sum over closed loops is characteristic of the Z2 spin liquid, introduced in N. Read and S. Sachdev, Phys. Rev. Lett. 66, 1773 (1991), X.-G. Wen, Phys. Rev. B 44, 2664 (1991)

P

Wednesday, January 16, 13

Page 25: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

B

Entanglement in the Z2 spin liquid

A

Sum over closed loops: only an even number of links cross the boundary between A and B

The sum over closed loops is characteristic of the Z2 spin liquid, introduced in N. Read and S. Sachdev, Phys. Rev. Lett. 66, 1773 (1991), X.-G. Wen, Phys. Rev. B 44, 2664 (1991)

P

Wednesday, January 16, 13

Page 26: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

SE = aP � ln(2)

where P is the surface area (perimeter)

of the boundary between A and B.

B

Entanglement in the Z2 spin liquid

A P

A. Hamma, R. Ionicioiu, and P. Zanardi, Phys. Rev. A 71, 022315 (2005) M. Levin and X.-G. Wen, Phys. Rev. Lett. 96, 110405 (2006); A. Kitaev and J. Preskill, Phys. Rev. Lett. 96, 110404 (2006)

Y. Zhang, T. Grover, and A. Vishwanath, Phys. Rev. B 84, 075128 (2011)Wednesday, January 16, 13

Page 27: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Simeng Yan, D. A. Huse, and S. R. White, Science 332, 1173 (2011).

Mott insulator: Kagome antiferromagnet

S. Depenbrock, I. P. McCulloch, and U. Schollwoeck, Physical Review Letters 109, 067201 (2012)

Hong-Chen Jiang, Z. Wang, and L. Balents, Nature Physics 8, 902 (2012)

Strong numerical evidence for a Z2 spin liquid

Wednesday, January 16, 13

Page 28: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Nature 492, 406 (2012)

LETTERdoi:10.1038/nature11659

Fractionalized excitations in the spin-liquid state of akagome-lattice antiferromagnetTian-Heng Han1, Joel S. Helton2, Shaoyan Chu3, Daniel G. Nocera4, Jose A. Rodriguez-Rivera2,5, Collin Broholm2,6 & Young S. Lee1

The experimental realization of quantum spin liquids is a long-sought goal in physics, as they represent new states of matter. Quan-tum spin liquids cannot be described by the broken symmetriesassociated with conventional ground states. In fact, the interactingmagnetic moments in these systems do not order, but are highlyentangled with one another over long ranges1. Spin liquids have aprominent role in theories describing high-transition-temperaturesuperconductors2,3, and the topological properties of these statesmay have applications in quantum information4. A key feature ofspin liquids is that they support exotic spin excitations carryingfractional quantum numbers. However, detailed measurements ofthese ‘fractionalized excitations’ have been lacking. Here we reportneutron scattering measurements on single-crystal samples of thespin-1/2 kagome-lattice antiferromagnet ZnCu3(OD)6Cl2 (also calledherbertsmithite), which provide striking evidence for this characte-ristic feature of spin liquids. At low temperatures, we find that thespin excitations form a continuum, in contrast to the conventionalspin waves expected in ordered antiferromagnets. The observation ofsuch a continuum is noteworthy because, so far, this signature offractional spin excitations has been observed only in one-dimensionalsystems. The results also serve as a hallmark of the quantum spin-liquid state in herbertsmithite.

In a spin liquid, the atomic magnetic moments are strongly corre-lated but do not order or freeze even in the limit as the temperature, T,goes to zero. Although many types of quantum spin-liquid states existin theory, a feature that is expected to be common to all is the presenceof deconfined spinons as an elementary excitation from the groundstate1. Spinons are spin-half (S 5 1/2) quantum excitations into whichconventional spin-wave excitations with S 5 1 fractionalize. In onedimension, this phenomenon is well established for the S 5 1/2Heisenberg antiferromagnetic chain, where spinons may be thoughtof as magnetic domain boundaries that disrupt Neel order and are freeto propagate away from each other. In the one-dimensional compoundKCuF3, a continuum of spinon excitations has been well characterizedusing neutron scattering5. In two dimensions, the nature of the spinonexcitations is less clear. First, the existence of two-dimensional magnetswith a quantum spin-liquid ground state is still a matter of great debate.Second, the various spin-liquid states which are proposed in theorygive rise to a variety of spinon excitation spectra, which may be eithergapped or gapless.

The S 5 1/2 kagome-lattice Heisenberg antiferromagnet has longbeen recognized as a promising system in which to search for quantumspin-liquid states, because the kagome network of corner-sharing tri-angles frustrates long-range magnetic order6–8. We have devised syn-thetic methods to produce herbertsmithite (ZnCu3(OH)6Cl2) in whichthe S 5 1/2 Cu21 moments are arranged on a structurally perfectkagome lattice9 and nonmagnetic Zn21 ions separate the lattice planes.A depiction of the crystal structure is shown in Supplementary Fig. 1.Whereas herbertsmithite typically contains a small percentage of excess

Cu21 ions (,5% of the total) substituting for Zn21 ions in the inter-layer sites, the kagome planes contain only Cu21 ions10. Measurementson powder samples11–13 indicate strong antiferromagnetic super-exchange (J < 17 meV, where J is the exchange coupling that appearsin the nearest-neighbour Heisenberg Hamiltonian) and the absenceof long-range magnetic order or spin freezing down to tempera-tures of T 5 0.05 K. The bulk magnetic properties reveal a smallDzyaloshinskii–Moriya interaction and an easy-axis exchangeanisotropy14,15, both of order J/10. Despite these small imperfections,the nearest-neighbour Heisenberg model on a kagome lattice is still anexcellent approximation of the spin Hamiltonian for herbertsmithite.This is especially important, because recent calculations on recordlattice sizes indicate that the ground state of this model is in fact aquantum spin liquid16. Thus, experiments to probe the spin correla-tions in herbertsmithite are all the more urgent.

To this end, we recently succeeded in developing a technique for thegrowth of large, high-quality single crystals of herbertsmithite17, andsmall pieces have been used in studies involving local probes18,19,anomalous X-ray diffraction10, susceptibility15 and Raman scattering20.In this Letter, we report inelastic neutron scattering measurements ona large, deuterated, single-crystal sample of herbertsmithite. The neu-tron scattering cross-section is directly proportional to the dynamicstructure factor Stot(Q, v) (where Q and v stand for the momentumand energy transferred to the sample, respectively), which includesboth the nuclear and magnetic signals. The magnetic part, Smag(Q, v),is the Fourier transform (in time and space) of the spin–spin correlationfunction and can be obtained by subtracting the nuclear scattering asdescribed in the Supplementary Information. After calibration withrespect to a vanadium standard, the measured structure factors areexpressed in absolute units.

Contour plots of Stot(Q, v) are shown in Fig. 1a–c for T 5 1.6 K andthree different energy transfers Bv (B denotes Planck’s constantdivided by 2p). Figure 1a shows data for Bv 5 6 meV. Surprisingly,the scattered intensity is exceedingly diffuse, spanning a large fractionof the hexagonal Brillouin zone. A similar pattern of diffuse scatter-ing is observed for Bv 5 2 meV (Fig. 1b). The diffuse nature of thescattering at a temperature that is two orders of magnitude below theexchange energy scale, J, is in strong contrast to observations in non-frustrated quantum magnets. The S 5 1/2 square-lattice antiferromag-net La2CuO4 develops substantial antiferromagnetic correlations forT , J/2 (ref. 21), temperatures at which the low-energy scattering isstrongly peaked near the (p,p) point in reciprocal space. In herbert-smithite, the scattered intensity is not strongly peaked at any spe-cific point, and this remains true for all energies measured fromBv 5 0.25 to 11 meV. This behaviour is also markedly differentfrom that observed in the larger, S 5 5/2 kagome antiferromagnetKFe3(OH)6(SO4)2 which becomes magnetically ordered at low tem-peratures and has magnetic peaks at q 5 0 wavevectors above theordering temperature22.

1Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA. 2NIST Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg,Maryland 20899, USA. 3Center for Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA. 4Department of Chemistry, Massachusetts Instituteof Technology, Cambridge, Massachusetts 02139, USA. 5Department of Materials Science and Engineering, University of Maryland, College Park, Maryland 20742, USA. 6Institute for Quantum Matter andDepartment of Physics and Astronomy, The Johns Hopkins University, Baltimore, Maryland 21218, USA.

4 0 6 | N A T U R E | V O L 4 9 2 | 2 0 / 2 7 D E C E M B E R 2 0 1 2

Macmillan Publishers Limited. All rights reserved©2012

Wednesday, January 16, 13

Page 29: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Gapped quantum matter Spin liquids, quantum Hall states

Conformal quantum matter Quantum critical points in antiferromagnets, superconductors, and ultracold atoms; graphene

Compressible quantum matter Strange metals in high temperature superconductors, Bose metals

S. Sachdev, 100th anniversary Solvay conference,arXiv:1203.4565

“Complex entangled” states of quantum matter,

not adiabatically connected to independent particle states

Wednesday, January 16, 13

Page 30: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Gapped quantum matter Spin liquids, quantum Hall states

Conformal quantum matter Quantum critical points in antiferromagnets, superconductors, and ultracold atoms; graphene

Compressible quantum matter Strange metals in high temperature superconductors, Bose metals

S. Sachdev, 100th anniversary Solvay conference,arXiv:1203.4565

“Complex entangled” states of quantum matter,

not adiabatically connected to independent particle states

Wednesday, January 16, 13

Page 31: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

1. Higgs quasi-normal mode near the superfluid-insulator transition in 2 dimensions

2. Quantum criticality and conformal field theories

3. Holography and the quasi-normal modes of black-hole horizons

Outline

Wednesday, January 16, 13

Page 32: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

1. Higgs quasi-normal mode near the superfluid-insulator transition in 2 dimensions

2. Quantum criticality and conformal field theories

3. Holography and the quasi-normal modes of black-hole horizons

Outline

Wednesday, January 16, 13

Page 33: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

M. Greiner, O. Mandel, T. Esslinger, T. W. Hänsch, and I. Bloch, Nature 415, 39 (2002).

Ultracold 87Rbatoms - bosons

Superfluid-insulator transition

Wednesday, January 16, 13

Page 34: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Insulator (the vacuum) at large repulsion between bosons

Wednesday, January 16, 13

Page 35: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Particles ⇠ †

Excitations of the insulator:

Wednesday, January 16, 13

Page 36: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Holes ⇠

Excitations of the insulator:

Wednesday, January 16, 13

Page 37: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Density of particles = density of holes )“Relativistic” field theory for :

S =

Zd2rd⌧

h|@⌧ |2 + c2|rr |2 + (�� �c)| |2 + u

�| |2

�2i

Excitations of the insulator:

M.P. A. Fisher, P.B. Weichmann, G. Grinstein, and D.S. Fisher, Phys. Rev. B 40, 546 (1989).

Particles ⇠ †

Holes ⇠

Wednesday, January 16, 13

Page 38: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

h i 6= 0 h i = 0

Wednesday, January 16, 13

Page 39: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

h i 6= 0 h i = 0

system with a recently developed scheme based on single-atom-resolved detection24. It is the high sensitivity of this method thatallowed us to reduce the modulation amplitude by almost an orderof magnitude compared with earlier experiments20,21 and to stay wellwithin the linear response regime (Supplementary Information).

The results for selected lattice depths V0 are shown in Fig. 2b. Weobserve a gapped response with an asymmetric overall shape that willbe analysed in the following paragraphs. Notably, the maximumobserved temperature after modulation is well below the ‘melting’temperature for a Mott insulator in the atomic limit25, Tmelt < 0.2U/kB

(kB, Boltzmann’s constant), demonstrating that our experiments probethe quantum gas in the degenerate regime. To obtain numerical valuesfor the onset of spectral response, we fitted each spectrum with an errorfunction centred at a frequency n0 (Fig. 2b, black lines). With japproaching jc, the shift of the gap to lower frequencies is alreadyvisible in the raw data (Fig. 2b) and becomes even more apparent forthe fitted gap n0 as a function of j/jc (Fig. 2a, filled circles). The n0 valuesare in quantitative agreement with a prediction for the Higgs gap nSF atcommensurate filling (solid line):

hnSF=U~ 3!!!2p

{4" #

1zj=jc! "$ %1=2

j=jc{1! "1=2

Here h denotes Planck’s constant. This value is based on an analysis ofvariations around a mean-field state7,16 (throughout the manuscript,we have rescaled jc in the theoretical calculations to match the valuejc<0:06 obtained from quantum Monte Carlo simulations26).

The sharpness of the spectral onset can be quantified by the width ofthe fitted error function, which is shown as vertical dashed lines inFig. 2a. Approaching the critical point, the spectral onset becomessharper, and the width normalized to the centre frequency n0 remainsconstant (Supplementary Fig. 3). The constancy of this ratio indicatesthat the width of the spectral onset scales with the distance to thecritical point in the same way as the gap frequency.

We observe similar gapped responses in the Mott insulating regime(Supplementary Information and Fig. 5a), with the gap closing con-tinuously when approaching the critical point (Fig. 2a, open circles).We interpret this as a result of combined particle and hole excitationswith a frequency given by the Mott excitation gap that closes at thetransition point16. The fitted gaps are consistent with the Mott gap

hnMI=U~ 1z 12!!!2p

{17" #

j=jc$ %1=2

1{j=jc! "1=2

where nMI is the Mott gap as predicted by mean-field theory16 (Fig. 2a,dashed line).

The observed softening of the onset of spectral response in thesuperfluid regime has led to an identification of the experimentalsignal with a response from collective excitations of Higgs type. Togain further insight into the full in-trap response, we calculated theeigenspectrum of the system in a Gutzwiller approach16,22 (Methodsand Supplementary Information). The result is a series of discreteeigenfrequencies (Fig. 3a), and the corresponding eigenmodes showin-trap superfluid density distributions, which are reminiscent of thevibrational modes of a drum (Fig. 3b). The frequency of the lowest-lying amplitude-like eigenmode n0,G closely follows the long-wave-length prediction for homogeneous commensurate filling nSF over awide range of couplings j/jc until the response rounds off in the vicinityof the critical point due to the finite size of the system (Fig. 3c). Fittingthe low-frequency edge of the experimental data can be interpreted asextracting the frequency of this mode, which explains the goodquantitative agreement with the prediction for the homogeneous com-mensurate filling in Fig. 2a. Modes at different frequencies from thelowest-lying amplitude-like mode broaden the spectrum only abovethe onset of spectral response.

An eigenmode analysis, however, does not yield any informationabout the finite spectral width of the modes, which stems from theinteraction between amplitude and phase excitations. We will considerthe question of the spectral width by analysing the low-, intermediate-and high-frequency parts of the response separately. We begin byexamining the low-frequency part of the response, which is expectedto be governed by a process coupling a virtually excited amplitudemode to a pair of phase modes with opposite momenta. As a result,the response of a strongly interacting, two-dimensional superfluid is

a1

2

V

Re( )Im( )

Higgs modeNambu–

Goldstonemode

j/jc 1

0 100 200 300 4000

5

10

15

20

Time (ms)

Lat

tice

dept

h (E

r)

Lattice loading Modulation Hold time Ramp to atomic limit

Temperaturemeasurement

V0

Ttot = 200 ms

A = 0.03V0

Tmod = 20WW

b

3

j/jc * 1

j/jc , 1

ΨΨ

Figure 1 | Illustration of the Higgs mode and experimental sequence.a, Classical energy density V as a function of the order parameter Y. Within theordered (superfluid) phase, Nambu–Goldstone and Higgs modes arise fromphase and amplitude modulations (blue and red arrows in panel 1). As thecoupling j 5 J/U (see main text) approaches the critical value jc, the energydensity transforms into a function with a minimum at Y 5 0 (panels 2 and 3).Simultaneously, the curvature in the radial direction decreases, leading to acharacteristic reduction of the excitation frequency for the Higgs mode. In thedisordered (Mott insulating) phase, two gapped modes exist, respectivelycorresponding to particle and hole excitations in our case (red and blue arrow inpanel 3). b, The Higgs mode can be excited with a periodic modulation of thecoupling j, which amounts to a ‘shaking’ of the classical energy densitypotential. In the experimental sequence, this is realized by a modulation of theoptical lattice potential (see main text for details). t 5 1/nmod; Er, lattice recoilenergy.

LETTER RESEARCH

2 6 J U L Y 2 0 1 2 | V O L 4 8 7 | N A T U R E | 4 5 5

Macmillan Publishers Limited. All rights reserved©2012

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

system with a recently developed scheme based on single-atom-resolved detection24. It is the high sensitivity of this method thatallowed us to reduce the modulation amplitude by almost an orderof magnitude compared with earlier experiments20,21 and to stay wellwithin the linear response regime (Supplementary Information).

The results for selected lattice depths V0 are shown in Fig. 2b. Weobserve a gapped response with an asymmetric overall shape that willbe analysed in the following paragraphs. Notably, the maximumobserved temperature after modulation is well below the ‘melting’temperature for a Mott insulator in the atomic limit25, Tmelt < 0.2U/kB

(kB, Boltzmann’s constant), demonstrating that our experiments probethe quantum gas in the degenerate regime. To obtain numerical valuesfor the onset of spectral response, we fitted each spectrum with an errorfunction centred at a frequency n0 (Fig. 2b, black lines). With japproaching jc, the shift of the gap to lower frequencies is alreadyvisible in the raw data (Fig. 2b) and becomes even more apparent forthe fitted gap n0 as a function of j/jc (Fig. 2a, filled circles). The n0 valuesare in quantitative agreement with a prediction for the Higgs gap nSF atcommensurate filling (solid line):

hnSF=U~ 3!!!2p

{4" #

1zj=jc! "$ %1=2

j=jc{1! "1=2

Here h denotes Planck’s constant. This value is based on an analysis ofvariations around a mean-field state7,16 (throughout the manuscript,we have rescaled jc in the theoretical calculations to match the valuejc<0:06 obtained from quantum Monte Carlo simulations26).

The sharpness of the spectral onset can be quantified by the width ofthe fitted error function, which is shown as vertical dashed lines inFig. 2a. Approaching the critical point, the spectral onset becomessharper, and the width normalized to the centre frequency n0 remainsconstant (Supplementary Fig. 3). The constancy of this ratio indicatesthat the width of the spectral onset scales with the distance to thecritical point in the same way as the gap frequency.

We observe similar gapped responses in the Mott insulating regime(Supplementary Information and Fig. 5a), with the gap closing con-tinuously when approaching the critical point (Fig. 2a, open circles).We interpret this as a result of combined particle and hole excitationswith a frequency given by the Mott excitation gap that closes at thetransition point16. The fitted gaps are consistent with the Mott gap

hnMI=U~ 1z 12!!!2p

{17" #

j=jc$ %1=2

1{j=jc! "1=2

where nMI is the Mott gap as predicted by mean-field theory16 (Fig. 2a,dashed line).

The observed softening of the onset of spectral response in thesuperfluid regime has led to an identification of the experimentalsignal with a response from collective excitations of Higgs type. Togain further insight into the full in-trap response, we calculated theeigenspectrum of the system in a Gutzwiller approach16,22 (Methodsand Supplementary Information). The result is a series of discreteeigenfrequencies (Fig. 3a), and the corresponding eigenmodes showin-trap superfluid density distributions, which are reminiscent of thevibrational modes of a drum (Fig. 3b). The frequency of the lowest-lying amplitude-like eigenmode n0,G closely follows the long-wave-length prediction for homogeneous commensurate filling nSF over awide range of couplings j/jc until the response rounds off in the vicinityof the critical point due to the finite size of the system (Fig. 3c). Fittingthe low-frequency edge of the experimental data can be interpreted asextracting the frequency of this mode, which explains the goodquantitative agreement with the prediction for the homogeneous com-mensurate filling in Fig. 2a. Modes at different frequencies from thelowest-lying amplitude-like mode broaden the spectrum only abovethe onset of spectral response.

An eigenmode analysis, however, does not yield any informationabout the finite spectral width of the modes, which stems from theinteraction between amplitude and phase excitations. We will considerthe question of the spectral width by analysing the low-, intermediate-and high-frequency parts of the response separately. We begin byexamining the low-frequency part of the response, which is expectedto be governed by a process coupling a virtually excited amplitudemode to a pair of phase modes with opposite momenta. As a result,the response of a strongly interacting, two-dimensional superfluid is

a1

2

V

Re( )Im( )

Higgs modeNambu–

Goldstonemode

j/jc 1

0 100 200 300 4000

5

10

15

20

Time (ms)

Lat

tice

dept

h (E

r)

Lattice loading Modulation Hold time Ramp to atomic limit

Temperaturemeasurement

V0

Ttot = 200 ms

A = 0.03V0

Tmod = 20WW

b

3

j/jc * 1

j/jc , 1

ΨΨ

Figure 1 | Illustration of the Higgs mode and experimental sequence.a, Classical energy density V as a function of the order parameter Y. Within theordered (superfluid) phase, Nambu–Goldstone and Higgs modes arise fromphase and amplitude modulations (blue and red arrows in panel 1). As thecoupling j 5 J/U (see main text) approaches the critical value jc, the energydensity transforms into a function with a minimum at Y 5 0 (panels 2 and 3).Simultaneously, the curvature in the radial direction decreases, leading to acharacteristic reduction of the excitation frequency for the Higgs mode. In thedisordered (Mott insulating) phase, two gapped modes exist, respectivelycorresponding to particle and hole excitations in our case (red and blue arrow inpanel 3). b, The Higgs mode can be excited with a periodic modulation of thecoupling j, which amounts to a ‘shaking’ of the classical energy densitypotential. In the experimental sequence, this is realized by a modulation of theoptical lattice potential (see main text for details). t 5 1/nmod; Er, lattice recoilenergy.

LETTER RESEARCH

2 6 J U L Y 2 0 1 2 | V O L 4 8 7 | N A T U R E | 4 5 5

Macmillan Publishers Limited. All rights reserved©2012

system with a recently developed scheme based on single-atom-resolved detection24. It is the high sensitivity of this method thatallowed us to reduce the modulation amplitude by almost an orderof magnitude compared with earlier experiments20,21 and to stay wellwithin the linear response regime (Supplementary Information).

The results for selected lattice depths V0 are shown in Fig. 2b. Weobserve a gapped response with an asymmetric overall shape that willbe analysed in the following paragraphs. Notably, the maximumobserved temperature after modulation is well below the ‘melting’temperature for a Mott insulator in the atomic limit25, Tmelt < 0.2U/kB

(kB, Boltzmann’s constant), demonstrating that our experiments probethe quantum gas in the degenerate regime. To obtain numerical valuesfor the onset of spectral response, we fitted each spectrum with an errorfunction centred at a frequency n0 (Fig. 2b, black lines). With japproaching jc, the shift of the gap to lower frequencies is alreadyvisible in the raw data (Fig. 2b) and becomes even more apparent forthe fitted gap n0 as a function of j/jc (Fig. 2a, filled circles). The n0 valuesare in quantitative agreement with a prediction for the Higgs gap nSF atcommensurate filling (solid line):

hnSF=U~ 3!!!2p

{4" #

1zj=jc! "$ %1=2

j=jc{1! "1=2

Here h denotes Planck’s constant. This value is based on an analysis ofvariations around a mean-field state7,16 (throughout the manuscript,we have rescaled jc in the theoretical calculations to match the valuejc<0:06 obtained from quantum Monte Carlo simulations26).

The sharpness of the spectral onset can be quantified by the width ofthe fitted error function, which is shown as vertical dashed lines inFig. 2a. Approaching the critical point, the spectral onset becomessharper, and the width normalized to the centre frequency n0 remainsconstant (Supplementary Fig. 3). The constancy of this ratio indicatesthat the width of the spectral onset scales with the distance to thecritical point in the same way as the gap frequency.

We observe similar gapped responses in the Mott insulating regime(Supplementary Information and Fig. 5a), with the gap closing con-tinuously when approaching the critical point (Fig. 2a, open circles).We interpret this as a result of combined particle and hole excitationswith a frequency given by the Mott excitation gap that closes at thetransition point16. The fitted gaps are consistent with the Mott gap

hnMI=U~ 1z 12!!!2p

{17" #

j=jc$ %1=2

1{j=jc! "1=2

where nMI is the Mott gap as predicted by mean-field theory16 (Fig. 2a,dashed line).

The observed softening of the onset of spectral response in thesuperfluid regime has led to an identification of the experimentalsignal with a response from collective excitations of Higgs type. Togain further insight into the full in-trap response, we calculated theeigenspectrum of the system in a Gutzwiller approach16,22 (Methodsand Supplementary Information). The result is a series of discreteeigenfrequencies (Fig. 3a), and the corresponding eigenmodes showin-trap superfluid density distributions, which are reminiscent of thevibrational modes of a drum (Fig. 3b). The frequency of the lowest-lying amplitude-like eigenmode n0,G closely follows the long-wave-length prediction for homogeneous commensurate filling nSF over awide range of couplings j/jc until the response rounds off in the vicinityof the critical point due to the finite size of the system (Fig. 3c). Fittingthe low-frequency edge of the experimental data can be interpreted asextracting the frequency of this mode, which explains the goodquantitative agreement with the prediction for the homogeneous com-mensurate filling in Fig. 2a. Modes at different frequencies from thelowest-lying amplitude-like mode broaden the spectrum only abovethe onset of spectral response.

An eigenmode analysis, however, does not yield any informationabout the finite spectral width of the modes, which stems from theinteraction between amplitude and phase excitations. We will considerthe question of the spectral width by analysing the low-, intermediate-and high-frequency parts of the response separately. We begin byexamining the low-frequency part of the response, which is expectedto be governed by a process coupling a virtually excited amplitudemode to a pair of phase modes with opposite momenta. As a result,the response of a strongly interacting, two-dimensional superfluid is

a1

2

V

Re( )Im( )

Higgs modeNambu–

Goldstonemode

j/jc 1

0 100 200 300 4000

5

10

15

20

Time (ms)

Lat

tice

dept

h (E

r)

Lattice loading Modulation Hold time Ramp to atomic limit

Temperaturemeasurement

V0

Ttot = 200 ms

A = 0.03V0

Tmod = 20WW

b

3

j/jc * 1

j/jc , 1

ΨΨ

Figure 1 | Illustration of the Higgs mode and experimental sequence.a, Classical energy density V as a function of the order parameter Y. Within theordered (superfluid) phase, Nambu–Goldstone and Higgs modes arise fromphase and amplitude modulations (blue and red arrows in panel 1). As thecoupling j 5 J/U (see main text) approaches the critical value jc, the energydensity transforms into a function with a minimum at Y 5 0 (panels 2 and 3).Simultaneously, the curvature in the radial direction decreases, leading to acharacteristic reduction of the excitation frequency for the Higgs mode. In thedisordered (Mott insulating) phase, two gapped modes exist, respectivelycorresponding to particle and hole excitations in our case (red and blue arrow inpanel 3). b, The Higgs mode can be excited with a periodic modulation of thecoupling j, which amounts to a ‘shaking’ of the classical energy densitypotential. In the experimental sequence, this is realized by a modulation of theoptical lattice potential (see main text for details). t 5 1/nmod; Er, lattice recoilenergy.

LETTER RESEARCH

2 6 J U L Y 2 0 1 2 | V O L 4 8 7 | N A T U R E | 4 5 5

Macmillan Publishers Limited. All rights reserved©2012

system with a recently developed scheme based on single-atom-resolved detection24. It is the high sensitivity of this method thatallowed us to reduce the modulation amplitude by almost an orderof magnitude compared with earlier experiments20,21 and to stay wellwithin the linear response regime (Supplementary Information).

The results for selected lattice depths V0 are shown in Fig. 2b. Weobserve a gapped response with an asymmetric overall shape that willbe analysed in the following paragraphs. Notably, the maximumobserved temperature after modulation is well below the ‘melting’temperature for a Mott insulator in the atomic limit25, Tmelt < 0.2U/kB

(kB, Boltzmann’s constant), demonstrating that our experiments probethe quantum gas in the degenerate regime. To obtain numerical valuesfor the onset of spectral response, we fitted each spectrum with an errorfunction centred at a frequency n0 (Fig. 2b, black lines). With japproaching jc, the shift of the gap to lower frequencies is alreadyvisible in the raw data (Fig. 2b) and becomes even more apparent forthe fitted gap n0 as a function of j/jc (Fig. 2a, filled circles). The n0 valuesare in quantitative agreement with a prediction for the Higgs gap nSF atcommensurate filling (solid line):

hnSF=U~ 3!!!2p

{4" #

1zj=jc! "$ %1=2

j=jc{1! "1=2

Here h denotes Planck’s constant. This value is based on an analysis ofvariations around a mean-field state7,16 (throughout the manuscript,we have rescaled jc in the theoretical calculations to match the valuejc<0:06 obtained from quantum Monte Carlo simulations26).

The sharpness of the spectral onset can be quantified by the width ofthe fitted error function, which is shown as vertical dashed lines inFig. 2a. Approaching the critical point, the spectral onset becomessharper, and the width normalized to the centre frequency n0 remainsconstant (Supplementary Fig. 3). The constancy of this ratio indicatesthat the width of the spectral onset scales with the distance to thecritical point in the same way as the gap frequency.

We observe similar gapped responses in the Mott insulating regime(Supplementary Information and Fig. 5a), with the gap closing con-tinuously when approaching the critical point (Fig. 2a, open circles).We interpret this as a result of combined particle and hole excitationswith a frequency given by the Mott excitation gap that closes at thetransition point16. The fitted gaps are consistent with the Mott gap

hnMI=U~ 1z 12!!!2p

{17" #

j=jc$ %1=2

1{j=jc! "1=2

where nMI is the Mott gap as predicted by mean-field theory16 (Fig. 2a,dashed line).

The observed softening of the onset of spectral response in thesuperfluid regime has led to an identification of the experimentalsignal with a response from collective excitations of Higgs type. Togain further insight into the full in-trap response, we calculated theeigenspectrum of the system in a Gutzwiller approach16,22 (Methodsand Supplementary Information). The result is a series of discreteeigenfrequencies (Fig. 3a), and the corresponding eigenmodes showin-trap superfluid density distributions, which are reminiscent of thevibrational modes of a drum (Fig. 3b). The frequency of the lowest-lying amplitude-like eigenmode n0,G closely follows the long-wave-length prediction for homogeneous commensurate filling nSF over awide range of couplings j/jc until the response rounds off in the vicinityof the critical point due to the finite size of the system (Fig. 3c). Fittingthe low-frequency edge of the experimental data can be interpreted asextracting the frequency of this mode, which explains the goodquantitative agreement with the prediction for the homogeneous com-mensurate filling in Fig. 2a. Modes at different frequencies from thelowest-lying amplitude-like mode broaden the spectrum only abovethe onset of spectral response.

An eigenmode analysis, however, does not yield any informationabout the finite spectral width of the modes, which stems from theinteraction between amplitude and phase excitations. We will considerthe question of the spectral width by analysing the low-, intermediate-and high-frequency parts of the response separately. We begin byexamining the low-frequency part of the response, which is expectedto be governed by a process coupling a virtually excited amplitudemode to a pair of phase modes with opposite momenta. As a result,the response of a strongly interacting, two-dimensional superfluid is

a1

2

V

Re( )Im( )

Higgs modeNambu–

Goldstonemode

j/jc 1

0 100 200 300 4000

5

10

15

20

Time (ms)

Lat

tice

dept

h (E

r)

Lattice loading Modulation Hold time Ramp to atomic limit

Temperaturemeasurement

V0

Ttot = 200 ms

A = 0.03V0

Tmod = 20WW

b

3

j/jc * 1

j/jc , 1

ΨΨ

Figure 1 | Illustration of the Higgs mode and experimental sequence.a, Classical energy density V as a function of the order parameter Y. Within theordered (superfluid) phase, Nambu–Goldstone and Higgs modes arise fromphase and amplitude modulations (blue and red arrows in panel 1). As thecoupling j 5 J/U (see main text) approaches the critical value jc, the energydensity transforms into a function with a minimum at Y 5 0 (panels 2 and 3).Simultaneously, the curvature in the radial direction decreases, leading to acharacteristic reduction of the excitation frequency for the Higgs mode. In thedisordered (Mott insulating) phase, two gapped modes exist, respectivelycorresponding to particle and hole excitations in our case (red and blue arrow inpanel 3). b, The Higgs mode can be excited with a periodic modulation of thecoupling j, which amounts to a ‘shaking’ of the classical energy densitypotential. In the experimental sequence, this is realized by a modulation of theoptical lattice potential (see main text for details). t 5 1/nmod; Er, lattice recoilenergy.

LETTER RESEARCH

2 6 J U L Y 2 0 1 2 | V O L 4 8 7 | N A T U R E | 4 5 5

Macmillan Publishers Limited. All rights reserved©2012

Particles and holes correspond

to the 2 normal modes in the

oscillation of about = 0.

Wednesday, January 16, 13

Page 40: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

h i 6= 0 h i = 0

Wednesday, January 16, 13

Page 41: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

h i 6= 0 h i = 0

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

system with a recently developed scheme based on single-atom-resolved detection24. It is the high sensitivity of this method thatallowed us to reduce the modulation amplitude by almost an orderof magnitude compared with earlier experiments20,21 and to stay wellwithin the linear response regime (Supplementary Information).

The results for selected lattice depths V0 are shown in Fig. 2b. Weobserve a gapped response with an asymmetric overall shape that willbe analysed in the following paragraphs. Notably, the maximumobserved temperature after modulation is well below the ‘melting’temperature for a Mott insulator in the atomic limit25, Tmelt < 0.2U/kB

(kB, Boltzmann’s constant), demonstrating that our experiments probethe quantum gas in the degenerate regime. To obtain numerical valuesfor the onset of spectral response, we fitted each spectrum with an errorfunction centred at a frequency n0 (Fig. 2b, black lines). With japproaching jc, the shift of the gap to lower frequencies is alreadyvisible in the raw data (Fig. 2b) and becomes even more apparent forthe fitted gap n0 as a function of j/jc (Fig. 2a, filled circles). The n0 valuesare in quantitative agreement with a prediction for the Higgs gap nSF atcommensurate filling (solid line):

hnSF=U~ 3!!!2p

{4" #

1zj=jc! "$ %1=2

j=jc{1! "1=2

Here h denotes Planck’s constant. This value is based on an analysis ofvariations around a mean-field state7,16 (throughout the manuscript,we have rescaled jc in the theoretical calculations to match the valuejc<0:06 obtained from quantum Monte Carlo simulations26).

The sharpness of the spectral onset can be quantified by the width ofthe fitted error function, which is shown as vertical dashed lines inFig. 2a. Approaching the critical point, the spectral onset becomessharper, and the width normalized to the centre frequency n0 remainsconstant (Supplementary Fig. 3). The constancy of this ratio indicatesthat the width of the spectral onset scales with the distance to thecritical point in the same way as the gap frequency.

We observe similar gapped responses in the Mott insulating regime(Supplementary Information and Fig. 5a), with the gap closing con-tinuously when approaching the critical point (Fig. 2a, open circles).We interpret this as a result of combined particle and hole excitationswith a frequency given by the Mott excitation gap that closes at thetransition point16. The fitted gaps are consistent with the Mott gap

hnMI=U~ 1z 12!!!2p

{17" #

j=jc$ %1=2

1{j=jc! "1=2

where nMI is the Mott gap as predicted by mean-field theory16 (Fig. 2a,dashed line).

The observed softening of the onset of spectral response in thesuperfluid regime has led to an identification of the experimentalsignal with a response from collective excitations of Higgs type. Togain further insight into the full in-trap response, we calculated theeigenspectrum of the system in a Gutzwiller approach16,22 (Methodsand Supplementary Information). The result is a series of discreteeigenfrequencies (Fig. 3a), and the corresponding eigenmodes showin-trap superfluid density distributions, which are reminiscent of thevibrational modes of a drum (Fig. 3b). The frequency of the lowest-lying amplitude-like eigenmode n0,G closely follows the long-wave-length prediction for homogeneous commensurate filling nSF over awide range of couplings j/jc until the response rounds off in the vicinityof the critical point due to the finite size of the system (Fig. 3c). Fittingthe low-frequency edge of the experimental data can be interpreted asextracting the frequency of this mode, which explains the goodquantitative agreement with the prediction for the homogeneous com-mensurate filling in Fig. 2a. Modes at different frequencies from thelowest-lying amplitude-like mode broaden the spectrum only abovethe onset of spectral response.

An eigenmode analysis, however, does not yield any informationabout the finite spectral width of the modes, which stems from theinteraction between amplitude and phase excitations. We will considerthe question of the spectral width by analysing the low-, intermediate-and high-frequency parts of the response separately. We begin byexamining the low-frequency part of the response, which is expectedto be governed by a process coupling a virtually excited amplitudemode to a pair of phase modes with opposite momenta. As a result,the response of a strongly interacting, two-dimensional superfluid is

a1

2

V

Re( )Im( )

Higgs modeNambu–

Goldstonemode

j/jc 1

0 100 200 300 4000

5

10

15

20

Time (ms)

Lat

tice

dept

h (E

r)

Lattice loading Modulation Hold time Ramp to atomic limit

Temperaturemeasurement

V0

Ttot = 200 ms

A = 0.03V0

Tmod = 20WW

b

3

j/jc * 1

j/jc , 1

ΨΨ

Figure 1 | Illustration of the Higgs mode and experimental sequence.a, Classical energy density V as a function of the order parameter Y. Within theordered (superfluid) phase, Nambu–Goldstone and Higgs modes arise fromphase and amplitude modulations (blue and red arrows in panel 1). As thecoupling j 5 J/U (see main text) approaches the critical value jc, the energydensity transforms into a function with a minimum at Y 5 0 (panels 2 and 3).Simultaneously, the curvature in the radial direction decreases, leading to acharacteristic reduction of the excitation frequency for the Higgs mode. In thedisordered (Mott insulating) phase, two gapped modes exist, respectivelycorresponding to particle and hole excitations in our case (red and blue arrow inpanel 3). b, The Higgs mode can be excited with a periodic modulation of thecoupling j, which amounts to a ‘shaking’ of the classical energy densitypotential. In the experimental sequence, this is realized by a modulation of theoptical lattice potential (see main text for details). t 5 1/nmod; Er, lattice recoilenergy.

LETTER RESEARCH

2 6 J U L Y 2 0 1 2 | V O L 4 8 7 | N A T U R E | 4 5 5

Macmillan Publishers Limited. All rights reserved©2012

Nambu-Goldstone mode is the

oscillation in the phase

at a constant non-zero | |.

Wednesday, January 16, 13

Page 42: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

h i 6= 0 h i = 0

Wednesday, January 16, 13

Page 43: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

A conformal field theory

in 2+1 spacetime dimensions:

a CFT3

h i 6= 0 h i = 0

Wednesday, January 16, 13

Page 44: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Quantum state with

complex, many-body,

“long-range” quantum entanglement

h i 6= 0 h i = 0

Wednesday, January 16, 13

Page 45: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

No well-defined normal modes,

or particle-like excitations

h i 6= 0 h i = 0

Wednesday, January 16, 13

Page 46: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

h i 6= 0 h i = 0

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

system with a recently developed scheme based on single-atom-resolved detection24. It is the high sensitivity of this method thatallowed us to reduce the modulation amplitude by almost an orderof magnitude compared with earlier experiments20,21 and to stay wellwithin the linear response regime (Supplementary Information).

The results for selected lattice depths V0 are shown in Fig. 2b. Weobserve a gapped response with an asymmetric overall shape that willbe analysed in the following paragraphs. Notably, the maximumobserved temperature after modulation is well below the ‘melting’temperature for a Mott insulator in the atomic limit25, Tmelt < 0.2U/kB

(kB, Boltzmann’s constant), demonstrating that our experiments probethe quantum gas in the degenerate regime. To obtain numerical valuesfor the onset of spectral response, we fitted each spectrum with an errorfunction centred at a frequency n0 (Fig. 2b, black lines). With japproaching jc, the shift of the gap to lower frequencies is alreadyvisible in the raw data (Fig. 2b) and becomes even more apparent forthe fitted gap n0 as a function of j/jc (Fig. 2a, filled circles). The n0 valuesare in quantitative agreement with a prediction for the Higgs gap nSF atcommensurate filling (solid line):

hnSF=U~ 3!!!2p

{4" #

1zj=jc! "$ %1=2

j=jc{1! "1=2

Here h denotes Planck’s constant. This value is based on an analysis ofvariations around a mean-field state7,16 (throughout the manuscript,we have rescaled jc in the theoretical calculations to match the valuejc<0:06 obtained from quantum Monte Carlo simulations26).

The sharpness of the spectral onset can be quantified by the width ofthe fitted error function, which is shown as vertical dashed lines inFig. 2a. Approaching the critical point, the spectral onset becomessharper, and the width normalized to the centre frequency n0 remainsconstant (Supplementary Fig. 3). The constancy of this ratio indicatesthat the width of the spectral onset scales with the distance to thecritical point in the same way as the gap frequency.

We observe similar gapped responses in the Mott insulating regime(Supplementary Information and Fig. 5a), with the gap closing con-tinuously when approaching the critical point (Fig. 2a, open circles).We interpret this as a result of combined particle and hole excitationswith a frequency given by the Mott excitation gap that closes at thetransition point16. The fitted gaps are consistent with the Mott gap

hnMI=U~ 1z 12!!!2p

{17" #

j=jc$ %1=2

1{j=jc! "1=2

where nMI is the Mott gap as predicted by mean-field theory16 (Fig. 2a,dashed line).

The observed softening of the onset of spectral response in thesuperfluid regime has led to an identification of the experimentalsignal with a response from collective excitations of Higgs type. Togain further insight into the full in-trap response, we calculated theeigenspectrum of the system in a Gutzwiller approach16,22 (Methodsand Supplementary Information). The result is a series of discreteeigenfrequencies (Fig. 3a), and the corresponding eigenmodes showin-trap superfluid density distributions, which are reminiscent of thevibrational modes of a drum (Fig. 3b). The frequency of the lowest-lying amplitude-like eigenmode n0,G closely follows the long-wave-length prediction for homogeneous commensurate filling nSF over awide range of couplings j/jc until the response rounds off in the vicinityof the critical point due to the finite size of the system (Fig. 3c). Fittingthe low-frequency edge of the experimental data can be interpreted asextracting the frequency of this mode, which explains the goodquantitative agreement with the prediction for the homogeneous com-mensurate filling in Fig. 2a. Modes at different frequencies from thelowest-lying amplitude-like mode broaden the spectrum only abovethe onset of spectral response.

An eigenmode analysis, however, does not yield any informationabout the finite spectral width of the modes, which stems from theinteraction between amplitude and phase excitations. We will considerthe question of the spectral width by analysing the low-, intermediate-and high-frequency parts of the response separately. We begin byexamining the low-frequency part of the response, which is expectedto be governed by a process coupling a virtually excited amplitudemode to a pair of phase modes with opposite momenta. As a result,the response of a strongly interacting, two-dimensional superfluid is

a1

2

V

Re( )Im( )

Higgs modeNambu–

Goldstonemode

j/jc 1

0 100 200 300 4000

5

10

15

20

Time (ms)

Lat

tice

dept

h (E

r)

Lattice loading Modulation Hold time Ramp to atomic limit

Temperaturemeasurement

V0

Ttot = 200 ms

A = 0.03V0

Tmod = 20WW

b

3

j/jc * 1

j/jc , 1

ΨΨ

Figure 1 | Illustration of the Higgs mode and experimental sequence.a, Classical energy density V as a function of the order parameter Y. Within theordered (superfluid) phase, Nambu–Goldstone and Higgs modes arise fromphase and amplitude modulations (blue and red arrows in panel 1). As thecoupling j 5 J/U (see main text) approaches the critical value jc, the energydensity transforms into a function with a minimum at Y 5 0 (panels 2 and 3).Simultaneously, the curvature in the radial direction decreases, leading to acharacteristic reduction of the excitation frequency for the Higgs mode. In thedisordered (Mott insulating) phase, two gapped modes exist, respectivelycorresponding to particle and hole excitations in our case (red and blue arrow inpanel 3). b, The Higgs mode can be excited with a periodic modulation of thecoupling j, which amounts to a ‘shaking’ of the classical energy densitypotential. In the experimental sequence, this is realized by a modulation of theoptical lattice potential (see main text for details). t 5 1/nmod; Er, lattice recoilenergy.

LETTER RESEARCH

2 6 J U L Y 2 0 1 2 | V O L 4 8 7 | N A T U R E | 4 5 5

Macmillan Publishers Limited. All rights reserved©2012

Nambu-Goldstone mode is the

oscillation in the phase

at a constant non-zero | |.

Wednesday, January 16, 13

Page 47: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

h i 6= 0 h i = 0

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

system with a recently developed scheme based on single-atom-resolved detection24. It is the high sensitivity of this method thatallowed us to reduce the modulation amplitude by almost an orderof magnitude compared with earlier experiments20,21 and to stay wellwithin the linear response regime (Supplementary Information).

The results for selected lattice depths V0 are shown in Fig. 2b. Weobserve a gapped response with an asymmetric overall shape that willbe analysed in the following paragraphs. Notably, the maximumobserved temperature after modulation is well below the ‘melting’temperature for a Mott insulator in the atomic limit25, Tmelt < 0.2U/kB

(kB, Boltzmann’s constant), demonstrating that our experiments probethe quantum gas in the degenerate regime. To obtain numerical valuesfor the onset of spectral response, we fitted each spectrum with an errorfunction centred at a frequency n0 (Fig. 2b, black lines). With japproaching jc, the shift of the gap to lower frequencies is alreadyvisible in the raw data (Fig. 2b) and becomes even more apparent forthe fitted gap n0 as a function of j/jc (Fig. 2a, filled circles). The n0 valuesare in quantitative agreement with a prediction for the Higgs gap nSF atcommensurate filling (solid line):

hnSF=U~ 3!!!2p

{4" #

1zj=jc! "$ %1=2

j=jc{1! "1=2

Here h denotes Planck’s constant. This value is based on an analysis ofvariations around a mean-field state7,16 (throughout the manuscript,we have rescaled jc in the theoretical calculations to match the valuejc<0:06 obtained from quantum Monte Carlo simulations26).

The sharpness of the spectral onset can be quantified by the width ofthe fitted error function, which is shown as vertical dashed lines inFig. 2a. Approaching the critical point, the spectral onset becomessharper, and the width normalized to the centre frequency n0 remainsconstant (Supplementary Fig. 3). The constancy of this ratio indicatesthat the width of the spectral onset scales with the distance to thecritical point in the same way as the gap frequency.

We observe similar gapped responses in the Mott insulating regime(Supplementary Information and Fig. 5a), with the gap closing con-tinuously when approaching the critical point (Fig. 2a, open circles).We interpret this as a result of combined particle and hole excitationswith a frequency given by the Mott excitation gap that closes at thetransition point16. The fitted gaps are consistent with the Mott gap

hnMI=U~ 1z 12!!!2p

{17" #

j=jc$ %1=2

1{j=jc! "1=2

where nMI is the Mott gap as predicted by mean-field theory16 (Fig. 2a,dashed line).

The observed softening of the onset of spectral response in thesuperfluid regime has led to an identification of the experimentalsignal with a response from collective excitations of Higgs type. Togain further insight into the full in-trap response, we calculated theeigenspectrum of the system in a Gutzwiller approach16,22 (Methodsand Supplementary Information). The result is a series of discreteeigenfrequencies (Fig. 3a), and the corresponding eigenmodes showin-trap superfluid density distributions, which are reminiscent of thevibrational modes of a drum (Fig. 3b). The frequency of the lowest-lying amplitude-like eigenmode n0,G closely follows the long-wave-length prediction for homogeneous commensurate filling nSF over awide range of couplings j/jc until the response rounds off in the vicinityof the critical point due to the finite size of the system (Fig. 3c). Fittingthe low-frequency edge of the experimental data can be interpreted asextracting the frequency of this mode, which explains the goodquantitative agreement with the prediction for the homogeneous com-mensurate filling in Fig. 2a. Modes at different frequencies from thelowest-lying amplitude-like mode broaden the spectrum only abovethe onset of spectral response.

An eigenmode analysis, however, does not yield any informationabout the finite spectral width of the modes, which stems from theinteraction between amplitude and phase excitations. We will considerthe question of the spectral width by analysing the low-, intermediate-and high-frequency parts of the response separately. We begin byexamining the low-frequency part of the response, which is expectedto be governed by a process coupling a virtually excited amplitudemode to a pair of phase modes with opposite momenta. As a result,the response of a strongly interacting, two-dimensional superfluid is

a1

2

V

Re( )Im( )

Higgs modeNambu–

Goldstonemode

j/jc 1

0 100 200 300 4000

5

10

15

20

Time (ms)

Lat

tice

dept

h (E

r)

Lattice loading Modulation Hold time Ramp to atomic limit

Temperaturemeasurement

V0

Ttot = 200 ms

A = 0.03V0

Tmod = 20WW

b

3

j/jc * 1

j/jc , 1

ΨΨ

Figure 1 | Illustration of the Higgs mode and experimental sequence.a, Classical energy density V as a function of the order parameter Y. Within theordered (superfluid) phase, Nambu–Goldstone and Higgs modes arise fromphase and amplitude modulations (blue and red arrows in panel 1). As thecoupling j 5 J/U (see main text) approaches the critical value jc, the energydensity transforms into a function with a minimum at Y 5 0 (panels 2 and 3).Simultaneously, the curvature in the radial direction decreases, leading to acharacteristic reduction of the excitation frequency for the Higgs mode. In thedisordered (Mott insulating) phase, two gapped modes exist, respectivelycorresponding to particle and hole excitations in our case (red and blue arrow inpanel 3). b, The Higgs mode can be excited with a periodic modulation of thecoupling j, which amounts to a ‘shaking’ of the classical energy densitypotential. In the experimental sequence, this is realized by a modulation of theoptical lattice potential (see main text for details). t 5 1/nmod; Er, lattice recoilenergy.

LETTER RESEARCH

2 6 J U L Y 2 0 1 2 | V O L 4 8 7 | N A T U R E | 4 5 5

Macmillan Publishers Limited. All rights reserved©2012

Higgs mode is theoscillation in theamplitude | |. This decaysrapidly by emitting pairsof Nambu-Goldstone modes.

Wednesday, January 16, 13

Page 48: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

h i 6= 0 h i = 0

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

system with a recently developed scheme based on single-atom-resolved detection24. It is the high sensitivity of this method thatallowed us to reduce the modulation amplitude by almost an orderof magnitude compared with earlier experiments20,21 and to stay wellwithin the linear response regime (Supplementary Information).

The results for selected lattice depths V0 are shown in Fig. 2b. Weobserve a gapped response with an asymmetric overall shape that willbe analysed in the following paragraphs. Notably, the maximumobserved temperature after modulation is well below the ‘melting’temperature for a Mott insulator in the atomic limit25, Tmelt < 0.2U/kB

(kB, Boltzmann’s constant), demonstrating that our experiments probethe quantum gas in the degenerate regime. To obtain numerical valuesfor the onset of spectral response, we fitted each spectrum with an errorfunction centred at a frequency n0 (Fig. 2b, black lines). With japproaching jc, the shift of the gap to lower frequencies is alreadyvisible in the raw data (Fig. 2b) and becomes even more apparent forthe fitted gap n0 as a function of j/jc (Fig. 2a, filled circles). The n0 valuesare in quantitative agreement with a prediction for the Higgs gap nSF atcommensurate filling (solid line):

hnSF=U~ 3!!!2p

{4" #

1zj=jc! "$ %1=2

j=jc{1! "1=2

Here h denotes Planck’s constant. This value is based on an analysis ofvariations around a mean-field state7,16 (throughout the manuscript,we have rescaled jc in the theoretical calculations to match the valuejc<0:06 obtained from quantum Monte Carlo simulations26).

The sharpness of the spectral onset can be quantified by the width ofthe fitted error function, which is shown as vertical dashed lines inFig. 2a. Approaching the critical point, the spectral onset becomessharper, and the width normalized to the centre frequency n0 remainsconstant (Supplementary Fig. 3). The constancy of this ratio indicatesthat the width of the spectral onset scales with the distance to thecritical point in the same way as the gap frequency.

We observe similar gapped responses in the Mott insulating regime(Supplementary Information and Fig. 5a), with the gap closing con-tinuously when approaching the critical point (Fig. 2a, open circles).We interpret this as a result of combined particle and hole excitationswith a frequency given by the Mott excitation gap that closes at thetransition point16. The fitted gaps are consistent with the Mott gap

hnMI=U~ 1z 12!!!2p

{17" #

j=jc$ %1=2

1{j=jc! "1=2

where nMI is the Mott gap as predicted by mean-field theory16 (Fig. 2a,dashed line).

The observed softening of the onset of spectral response in thesuperfluid regime has led to an identification of the experimentalsignal with a response from collective excitations of Higgs type. Togain further insight into the full in-trap response, we calculated theeigenspectrum of the system in a Gutzwiller approach16,22 (Methodsand Supplementary Information). The result is a series of discreteeigenfrequencies (Fig. 3a), and the corresponding eigenmodes showin-trap superfluid density distributions, which are reminiscent of thevibrational modes of a drum (Fig. 3b). The frequency of the lowest-lying amplitude-like eigenmode n0,G closely follows the long-wave-length prediction for homogeneous commensurate filling nSF over awide range of couplings j/jc until the response rounds off in the vicinityof the critical point due to the finite size of the system (Fig. 3c). Fittingthe low-frequency edge of the experimental data can be interpreted asextracting the frequency of this mode, which explains the goodquantitative agreement with the prediction for the homogeneous com-mensurate filling in Fig. 2a. Modes at different frequencies from thelowest-lying amplitude-like mode broaden the spectrum only abovethe onset of spectral response.

An eigenmode analysis, however, does not yield any informationabout the finite spectral width of the modes, which stems from theinteraction between amplitude and phase excitations. We will considerthe question of the spectral width by analysing the low-, intermediate-and high-frequency parts of the response separately. We begin byexamining the low-frequency part of the response, which is expectedto be governed by a process coupling a virtually excited amplitudemode to a pair of phase modes with opposite momenta. As a result,the response of a strongly interacting, two-dimensional superfluid is

a1

2

V

Re( )Im( )

Higgs modeNambu–

Goldstonemode

j/jc 1

0 100 200 300 4000

5

10

15

20

Time (ms)

Lat

tice

dept

h (E

r)

Lattice loading Modulation Hold time Ramp to atomic limit

Temperaturemeasurement

V0

Ttot = 200 ms

A = 0.03V0

Tmod = 20WW

b

3

j/jc * 1

j/jc , 1

ΨΨ

Figure 1 | Illustration of the Higgs mode and experimental sequence.a, Classical energy density V as a function of the order parameter Y. Within theordered (superfluid) phase, Nambu–Goldstone and Higgs modes arise fromphase and amplitude modulations (blue and red arrows in panel 1). As thecoupling j 5 J/U (see main text) approaches the critical value jc, the energydensity transforms into a function with a minimum at Y 5 0 (panels 2 and 3).Simultaneously, the curvature in the radial direction decreases, leading to acharacteristic reduction of the excitation frequency for the Higgs mode. In thedisordered (Mott insulating) phase, two gapped modes exist, respectivelycorresponding to particle and hole excitations in our case (red and blue arrow inpanel 3). b, The Higgs mode can be excited with a periodic modulation of thecoupling j, which amounts to a ‘shaking’ of the classical energy densitypotential. In the experimental sequence, this is realized by a modulation of theoptical lattice potential (see main text for details). t 5 1/nmod; Er, lattice recoilenergy.

LETTER RESEARCH

2 6 J U L Y 2 0 1 2 | V O L 4 8 7 | N A T U R E | 4 5 5

Macmillan Publishers Limited. All rights reserved©2012

Despite rapid decay,there is a well-defined

Higgs “quasi-normal mode”.This is associated with a pole

in the lower-halfof the complex frequency plane.

A. V. Chubukov, S. Sachdev, and J. Ye, PRB 49,11919 (1994). S. Sachdev, PRB 59, 14054 (1999).

W. Zwerger, PRL 92, 027203 (2004).

D. Podolsky, A. Auerbach, and D. P. Arovas, PRB 84, 174522 (2011).

D. Podolsky and S. Sachdev, PRB 86, 054508 (2012). L. Pollet and N. Prokof’ev, PRL 109, 010401 (2012).

Wednesday, January 16, 13

Page 49: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

system with a recently developed scheme based on single-atom-resolved detection24. It is the high sensitivity of this method thatallowed us to reduce the modulation amplitude by almost an orderof magnitude compared with earlier experiments20,21 and to stay wellwithin the linear response regime (Supplementary Information).

The results for selected lattice depths V0 are shown in Fig. 2b. Weobserve a gapped response with an asymmetric overall shape that willbe analysed in the following paragraphs. Notably, the maximumobserved temperature after modulation is well below the ‘melting’temperature for a Mott insulator in the atomic limit25, Tmelt < 0.2U/kB

(kB, Boltzmann’s constant), demonstrating that our experiments probethe quantum gas in the degenerate regime. To obtain numerical valuesfor the onset of spectral response, we fitted each spectrum with an errorfunction centred at a frequency n0 (Fig. 2b, black lines). With japproaching jc, the shift of the gap to lower frequencies is alreadyvisible in the raw data (Fig. 2b) and becomes even more apparent forthe fitted gap n0 as a function of j/jc (Fig. 2a, filled circles). The n0 valuesare in quantitative agreement with a prediction for the Higgs gap nSF atcommensurate filling (solid line):

hnSF=U~ 3!!!2p

{4" #

1zj=jc! "$ %1=2

j=jc{1! "1=2

Here h denotes Planck’s constant. This value is based on an analysis ofvariations around a mean-field state7,16 (throughout the manuscript,we have rescaled jc in the theoretical calculations to match the valuejc<0:06 obtained from quantum Monte Carlo simulations26).

The sharpness of the spectral onset can be quantified by the width ofthe fitted error function, which is shown as vertical dashed lines inFig. 2a. Approaching the critical point, the spectral onset becomessharper, and the width normalized to the centre frequency n0 remainsconstant (Supplementary Fig. 3). The constancy of this ratio indicatesthat the width of the spectral onset scales with the distance to thecritical point in the same way as the gap frequency.

We observe similar gapped responses in the Mott insulating regime(Supplementary Information and Fig. 5a), with the gap closing con-tinuously when approaching the critical point (Fig. 2a, open circles).We interpret this as a result of combined particle and hole excitationswith a frequency given by the Mott excitation gap that closes at thetransition point16. The fitted gaps are consistent with the Mott gap

hnMI=U~ 1z 12!!!2p

{17" #

j=jc$ %1=2

1{j=jc! "1=2

where nMI is the Mott gap as predicted by mean-field theory16 (Fig. 2a,dashed line).

The observed softening of the onset of spectral response in thesuperfluid regime has led to an identification of the experimentalsignal with a response from collective excitations of Higgs type. Togain further insight into the full in-trap response, we calculated theeigenspectrum of the system in a Gutzwiller approach16,22 (Methodsand Supplementary Information). The result is a series of discreteeigenfrequencies (Fig. 3a), and the corresponding eigenmodes showin-trap superfluid density distributions, which are reminiscent of thevibrational modes of a drum (Fig. 3b). The frequency of the lowest-lying amplitude-like eigenmode n0,G closely follows the long-wave-length prediction for homogeneous commensurate filling nSF over awide range of couplings j/jc until the response rounds off in the vicinityof the critical point due to the finite size of the system (Fig. 3c). Fittingthe low-frequency edge of the experimental data can be interpreted asextracting the frequency of this mode, which explains the goodquantitative agreement with the prediction for the homogeneous com-mensurate filling in Fig. 2a. Modes at different frequencies from thelowest-lying amplitude-like mode broaden the spectrum only abovethe onset of spectral response.

An eigenmode analysis, however, does not yield any informationabout the finite spectral width of the modes, which stems from theinteraction between amplitude and phase excitations. We will considerthe question of the spectral width by analysing the low-, intermediate-and high-frequency parts of the response separately. We begin byexamining the low-frequency part of the response, which is expectedto be governed by a process coupling a virtually excited amplitudemode to a pair of phase modes with opposite momenta. As a result,the response of a strongly interacting, two-dimensional superfluid is

a1

2

V

Re( )Im( )

Higgs modeNambu–

Goldstonemode

j/jc 1

0 100 200 300 4000

5

10

15

20

Time (ms)

Lat

tice

dept

h (E

r)

Lattice loading Modulation Hold time Ramp to atomic limit

Temperaturemeasurement

V0

Ttot = 200 ms

A = 0.03V0

Tmod = 20WW

b

3

j/jc * 1

j/jc , 1

ΨΨ

Figure 1 | Illustration of the Higgs mode and experimental sequence.a, Classical energy density V as a function of the order parameter Y. Within theordered (superfluid) phase, Nambu–Goldstone and Higgs modes arise fromphase and amplitude modulations (blue and red arrows in panel 1). As thecoupling j 5 J/U (see main text) approaches the critical value jc, the energydensity transforms into a function with a minimum at Y 5 0 (panels 2 and 3).Simultaneously, the curvature in the radial direction decreases, leading to acharacteristic reduction of the excitation frequency for the Higgs mode. In thedisordered (Mott insulating) phase, two gapped modes exist, respectivelycorresponding to particle and hole excitations in our case (red and blue arrow inpanel 3). b, The Higgs mode can be excited with a periodic modulation of thecoupling j, which amounts to a ‘shaking’ of the classical energy densitypotential. In the experimental sequence, this is realized by a modulation of theoptical lattice potential (see main text for details). t 5 1/nmod; Er, lattice recoilenergy.

LETTER RESEARCH

2 6 J U L Y 2 0 1 2 | V O L 4 8 7 | N A T U R E | 4 5 5

Macmillan Publishers Limited. All rights reserved©2012

The Higgs quasi-normal mode is at the frequency

!pole

�= �i

4

⇡+

1

N

16�4 +

p2 log

�3� 2

p2��

⇡2

+ 2.46531203396 i

!+O

✓1

N2

where � is the particle gap at the complementary point in the “paramagnetic” statewith the same value of |�� �c|, and N = 2 is the number of vector components of .The universal answer is a consequence of the strong interactions in the CFT3

!

D. Podolsky and S. Sachdev, Phys. Rev. B 86, 054508 (2012).

Wednesday, January 16, 13

Page 50: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

expected to diverge at low frequencies, if the probe in use coupleslongitudinally to the order parameter2,4,5,9 (for example to the real partof Y, if the equilibrium value of Y was chosen along the real axis), as isthe case for neutron scattering. If, instead, the coupling is rotationallyinvariant (for example through coupling to jYj2), as expected forlattice modulation, such a divergence could be avoided and the

response is expected to scale as n3 at low frequencies3,6,9,17.Combining this result with the scaling dimensions of the responsefunction for a rotationally symmetric perturbation coupling to jYj2,we expect the low-frequency response to be proportional to(1 2 j/jc)

22n3 (ref. 9 and Methods). The experimentally observed sig-nal is consistent with this scaling at the ‘base’ of the absorption feature(Fig. 4). This indicates that the low-frequency part is dominated byonly a few in-trap eigenmodes, which approximately show the genericscaling of the homogeneous system for a response function describingcoupling to jYj2.

In the intermediate-frequency regime, it remains a challenge toconstruct a first-principles analytical treatment of the in-trap systemincluding all relevant decay and coupling processes. Lacking such atheory, we constructed a heuristic model combining the discrete spec-trum from the Gutzwiller approach (Fig. 3a) with the line shape for ahomogeneous system based on an O(N) field theory in two dimen-sions, calculated in the large-N limit3,6 (Methods). An implicit assump-tion of this approach is a continuum of phase modes, which is

Sup

er!u

id d

ensi

ty

0.1

0.3

0.5

0.7

a

b1

0 0.5 1 1.5 2

0

0.02

0.04

d

0 0.5 0.6 0.7 0.8 0.9 1 1.1

Q0,G

0

0.5

1

Line

str

engt

h (a

.u.)

hQmod/U

hQ/U

j/jc

c

1 1.5 2 2.50

0.4

0.8

1.2V0 = 9.5Erj/jc = 1.4

k BΔT

/U, Δ

E/U

Q0,G

2

2

1

3

3

4

4

hQmod/U

Figure 3 | Theory of in-trap response. a, A diagonalization of the trappedsystem in a Gutzwiller approximation shows a discrete spectrum of amplitude-like eigenmodes. Shown on the vertical axis is the strength of the response to amodulation of j. Eigenmodes of phase type are not shown (Methods) and n0,G

denotes the gap as calculated in the Gutzwiller approximation. a.u., arbitraryunits. b, In-trap superfluid density distribution for the four amplitude modeswith the lowest frequencies, as labelled in a. In contrast to the superfluiddensity, the total density of the system stays almost constant (not shown).c, Discrete amplitude mode spectrum for various couplings j/jc. Each red circlecorresponds to a single eigenmode, with the intensity of the colour beingproportional to the line strength. The gap frequency of the lowest-lying modefollows the prediction for commensurate filling (solid line; same as in Fig. 2a)until a rounding off takes place close to the critical point due to the finite size ofthe system. d, Comparison of the experimental response at V0 5 9.5Er (bluecircles and connecting blue line; error bars, s.e.m.) with a 2 3 2 cluster mean-field simulation (grey line and shaded area) and a heuristic model (dashed line;for details see text and Methods). The simulation was done for V0 5 9.5Er (greyline) and for V0 5 (1 6 0.02) 3 9.5Er (shaded grey area), to account for theexperimental uncertainty in the lattice depth, and predicts the energyabsorption per particle DE.

0 0.2 0.4 0.6 0.8 1

0

0.01

0.02

0.03

Qmod/U

(1 –

j/j c)

2 kBΔT

/U

0 0.2 0.4 0.6 0.8 1

0

0.01

0.02

0.03

Qmod/U

k BΔT

/U

Figure 4 | Scaling of the low-frequency response. The low-frequencyresponse in the superfluid regime shows a scaling compatible with theprediction (1 2 j/jc)

22n3 (Methods). Shown is the temperature responserescaled with (1 2 j/jc)

2 for V0 5 10Er (grey), 9.5Er (black), 9Er (green), 8.5Er

(blue) and 8Er (red) as a function of the modulation frequency. The black line isa fit of the form anb with a fitted exponent b 5 2.9(5). The inset shows the samedata points without rescaling, for comparison. Error bars, s.e.m.

hQ0/U

j/jc

0 0.5 1 1.5 2 2.50

0.2

0.4

0.6

0.8

1

1.20 0.03 0.06 0.09 0.12 0.15

j

Super!uidMott Insulator

a b

3

1

2

V0 = 8Erj/jc = 2.2

k BT/U

1

2

3

V0 = 9Erj/jc = 1.6

V0 = 10Erj/jc = 1.2

Q0

0.11

0.13

0.15

0.17

0.12

0.14

0.16

0.18

0 400 8000.12

0.14

0.16

0.18

Qmod (Hz)

Q0

Q0

Figure 2 | Softening of the Higgs mode. a, The fitted gap values hn0/U(circles) show a characteristic softening close to the critical point in quantitativeagreement with analytic predictions for the Higgs and the Mott gap (solid lineand dashed line, respectively; see text). Horizontal and vertical error barsdenote the experimental uncertainty of the lattice depths and the fit error for thecentre frequency of the error function, respectively (Methods). Vertical dashedlines denote the widths of the fitted error function and characterize thesharpness of the spectral onset. The blue shading highlights the superfluid

region. b, Temperature response to lattice modulation (circles and connectingblue line) and fit with an error function (solid black line) for the three differentpoints labelled in a. As the coupling j approaches the critical value jc, the changein the gap values to lower frequencies is clearly visible (from panel 1 to panel 3).Vertical dashed lines mark the frequency U/h corresponding to the on-siteinteraction. Each data point results from an average of the temperatures over,50 experimental runs. Error bars, s.e.m.

RESEARCH LETTER

4 5 6 | N A T U R E | V O L 4 8 7 | 2 6 J U L Y 2 0 1 2

Macmillan Publishers Limited. All rights reserved©2012

system with a recently developed scheme based on single-atom-resolved detection24. It is the high sensitivity of this method thatallowed us to reduce the modulation amplitude by almost an orderof magnitude compared with earlier experiments20,21 and to stay wellwithin the linear response regime (Supplementary Information).

The results for selected lattice depths V0 are shown in Fig. 2b. Weobserve a gapped response with an asymmetric overall shape that willbe analysed in the following paragraphs. Notably, the maximumobserved temperature after modulation is well below the ‘melting’temperature for a Mott insulator in the atomic limit25, Tmelt < 0.2U/kB

(kB, Boltzmann’s constant), demonstrating that our experiments probethe quantum gas in the degenerate regime. To obtain numerical valuesfor the onset of spectral response, we fitted each spectrum with an errorfunction centred at a frequency n0 (Fig. 2b, black lines). With japproaching jc, the shift of the gap to lower frequencies is alreadyvisible in the raw data (Fig. 2b) and becomes even more apparent forthe fitted gap n0 as a function of j/jc (Fig. 2a, filled circles). The n0 valuesare in quantitative agreement with a prediction for the Higgs gap nSF atcommensurate filling (solid line):

hnSF=U~ 3!!!2p

{4" #

1zj=jc! "$ %1=2

j=jc{1! "1=2

Here h denotes Planck’s constant. This value is based on an analysis ofvariations around a mean-field state7,16 (throughout the manuscript,we have rescaled jc in the theoretical calculations to match the valuejc<0:06 obtained from quantum Monte Carlo simulations26).

The sharpness of the spectral onset can be quantified by the width ofthe fitted error function, which is shown as vertical dashed lines inFig. 2a. Approaching the critical point, the spectral onset becomessharper, and the width normalized to the centre frequency n0 remainsconstant (Supplementary Fig. 3). The constancy of this ratio indicatesthat the width of the spectral onset scales with the distance to thecritical point in the same way as the gap frequency.

We observe similar gapped responses in the Mott insulating regime(Supplementary Information and Fig. 5a), with the gap closing con-tinuously when approaching the critical point (Fig. 2a, open circles).We interpret this as a result of combined particle and hole excitationswith a frequency given by the Mott excitation gap that closes at thetransition point16. The fitted gaps are consistent with the Mott gap

hnMI=U~ 1z 12!!!2p

{17" #

j=jc$ %1=2

1{j=jc! "1=2

where nMI is the Mott gap as predicted by mean-field theory16 (Fig. 2a,dashed line).

The observed softening of the onset of spectral response in thesuperfluid regime has led to an identification of the experimentalsignal with a response from collective excitations of Higgs type. Togain further insight into the full in-trap response, we calculated theeigenspectrum of the system in a Gutzwiller approach16,22 (Methodsand Supplementary Information). The result is a series of discreteeigenfrequencies (Fig. 3a), and the corresponding eigenmodes showin-trap superfluid density distributions, which are reminiscent of thevibrational modes of a drum (Fig. 3b). The frequency of the lowest-lying amplitude-like eigenmode n0,G closely follows the long-wave-length prediction for homogeneous commensurate filling nSF over awide range of couplings j/jc until the response rounds off in the vicinityof the critical point due to the finite size of the system (Fig. 3c). Fittingthe low-frequency edge of the experimental data can be interpreted asextracting the frequency of this mode, which explains the goodquantitative agreement with the prediction for the homogeneous com-mensurate filling in Fig. 2a. Modes at different frequencies from thelowest-lying amplitude-like mode broaden the spectrum only abovethe onset of spectral response.

An eigenmode analysis, however, does not yield any informationabout the finite spectral width of the modes, which stems from theinteraction between amplitude and phase excitations. We will considerthe question of the spectral width by analysing the low-, intermediate-and high-frequency parts of the response separately. We begin byexamining the low-frequency part of the response, which is expectedto be governed by a process coupling a virtually excited amplitudemode to a pair of phase modes with opposite momenta. As a result,the response of a strongly interacting, two-dimensional superfluid is

a1

2

V

Re( )Im( )

Higgs modeNambu–

Goldstonemode

j/jc 1

0 100 200 300 4000

5

10

15

20

Time (ms)

Lat

tice

dept

h (E

r)

Lattice loading Modulation Hold time Ramp to atomic limit

Temperaturemeasurement

V0

Ttot = 200 ms

A = 0.03V0

Tmod = 20WW

b

3

j/jc * 1

j/jc , 1

ΨΨ

Figure 1 | Illustration of the Higgs mode and experimental sequence.a, Classical energy density V as a function of the order parameter Y. Within theordered (superfluid) phase, Nambu–Goldstone and Higgs modes arise fromphase and amplitude modulations (blue and red arrows in panel 1). As thecoupling j 5 J/U (see main text) approaches the critical value jc, the energydensity transforms into a function with a minimum at Y 5 0 (panels 2 and 3).Simultaneously, the curvature in the radial direction decreases, leading to acharacteristic reduction of the excitation frequency for the Higgs mode. In thedisordered (Mott insulating) phase, two gapped modes exist, respectivelycorresponding to particle and hole excitations in our case (red and blue arrow inpanel 3). b, The Higgs mode can be excited with a periodic modulation of thecoupling j, which amounts to a ‘shaking’ of the classical energy densitypotential. In the experimental sequence, this is realized by a modulation of theoptical lattice potential (see main text for details). t 5 1/nmod; Er, lattice recoilenergy.

LETTER RESEARCH

2 6 J U L Y 2 0 1 2 | V O L 4 8 7 | N A T U R E | 4 5 5

Macmillan Publishers Limited. All rights reserved©2012

Observation of Higgs quasi-normal mode across the superfluid-insulator transition of ultracold atoms in a 2-dimensional optical lattice:Response to modulation of lattice depth scales as expected from the LHP pole

Manuel Endres, Takeshi Fukuhara, David Pekker, Marc Cheneau, Peter Schaub, Christian Gross, Eugene Demler, Stefan Kuhr, and Immanuel Bloch, Nature 487, 454 (2012).

Wednesday, January 16, 13

Page 51: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

expected to diverge at low frequencies, if the probe in use coupleslongitudinally to the order parameter2,4,5,9 (for example to the real partof Y, if the equilibrium value of Y was chosen along the real axis), as isthe case for neutron scattering. If, instead, the coupling is rotationallyinvariant (for example through coupling to jYj2), as expected forlattice modulation, such a divergence could be avoided and the

response is expected to scale as n3 at low frequencies3,6,9,17.Combining this result with the scaling dimensions of the responsefunction for a rotationally symmetric perturbation coupling to jYj2,we expect the low-frequency response to be proportional to(1 2 j/jc)

22n3 (ref. 9 and Methods). The experimentally observed sig-nal is consistent with this scaling at the ‘base’ of the absorption feature(Fig. 4). This indicates that the low-frequency part is dominated byonly a few in-trap eigenmodes, which approximately show the genericscaling of the homogeneous system for a response function describingcoupling to jYj2.

In the intermediate-frequency regime, it remains a challenge toconstruct a first-principles analytical treatment of the in-trap systemincluding all relevant decay and coupling processes. Lacking such atheory, we constructed a heuristic model combining the discrete spec-trum from the Gutzwiller approach (Fig. 3a) with the line shape for ahomogeneous system based on an O(N) field theory in two dimen-sions, calculated in the large-N limit3,6 (Methods). An implicit assump-tion of this approach is a continuum of phase modes, which is

Sup

er!u

id d

ensi

ty

0.1

0.3

0.5

0.7

a

b1

0 0.5 1 1.5 2

0

0.02

0.04

d

0 0.5 0.6 0.7 0.8 0.9 1 1.1

Q0,G

0

0.5

1

Line

str

engt

h (a

.u.)

hQmod/U

hQ/U

j/jc

c

1 1.5 2 2.50

0.4

0.8

1.2V0 = 9.5Erj/jc = 1.4

k BΔT

/U, Δ

E/U

Q0,G

2

2

1

3

3

4

4

hQmod/U

Figure 3 | Theory of in-trap response. a, A diagonalization of the trappedsystem in a Gutzwiller approximation shows a discrete spectrum of amplitude-like eigenmodes. Shown on the vertical axis is the strength of the response to amodulation of j. Eigenmodes of phase type are not shown (Methods) and n0,G

denotes the gap as calculated in the Gutzwiller approximation. a.u., arbitraryunits. b, In-trap superfluid density distribution for the four amplitude modeswith the lowest frequencies, as labelled in a. In contrast to the superfluiddensity, the total density of the system stays almost constant (not shown).c, Discrete amplitude mode spectrum for various couplings j/jc. Each red circlecorresponds to a single eigenmode, with the intensity of the colour beingproportional to the line strength. The gap frequency of the lowest-lying modefollows the prediction for commensurate filling (solid line; same as in Fig. 2a)until a rounding off takes place close to the critical point due to the finite size ofthe system. d, Comparison of the experimental response at V0 5 9.5Er (bluecircles and connecting blue line; error bars, s.e.m.) with a 2 3 2 cluster mean-field simulation (grey line and shaded area) and a heuristic model (dashed line;for details see text and Methods). The simulation was done for V0 5 9.5Er (greyline) and for V0 5 (1 6 0.02) 3 9.5Er (shaded grey area), to account for theexperimental uncertainty in the lattice depth, and predicts the energyabsorption per particle DE.

0 0.2 0.4 0.6 0.8 1

0

0.01

0.02

0.03

Qmod/U

(1 –

j/j c)

2 kBΔT

/U

0 0.2 0.4 0.6 0.8 1

0

0.01

0.02

0.03

Qmod/U

k BΔT

/U

Figure 4 | Scaling of the low-frequency response. The low-frequencyresponse in the superfluid regime shows a scaling compatible with theprediction (1 2 j/jc)

22n3 (Methods). Shown is the temperature responserescaled with (1 2 j/jc)

2 for V0 5 10Er (grey), 9.5Er (black), 9Er (green), 8.5Er

(blue) and 8Er (red) as a function of the modulation frequency. The black line isa fit of the form anb with a fitted exponent b 5 2.9(5). The inset shows the samedata points without rescaling, for comparison. Error bars, s.e.m.

hQ0/U

j/jc

0 0.5 1 1.5 2 2.50

0.2

0.4

0.6

0.8

1

1.20 0.03 0.06 0.09 0.12 0.15

j

Super!uidMott Insulator

a b

3

1

2

V0 = 8Erj/jc = 2.2

k BT/U

1

2

3

V0 = 9Erj/jc = 1.6

V0 = 10Erj/jc = 1.2

Q0

0.11

0.13

0.15

0.17

0.12

0.14

0.16

0.18

0 400 8000.12

0.14

0.16

0.18

Qmod (Hz)

Q0

Q0

Figure 2 | Softening of the Higgs mode. a, The fitted gap values hn0/U(circles) show a characteristic softening close to the critical point in quantitativeagreement with analytic predictions for the Higgs and the Mott gap (solid lineand dashed line, respectively; see text). Horizontal and vertical error barsdenote the experimental uncertainty of the lattice depths and the fit error for thecentre frequency of the error function, respectively (Methods). Vertical dashedlines denote the widths of the fitted error function and characterize thesharpness of the spectral onset. The blue shading highlights the superfluid

region. b, Temperature response to lattice modulation (circles and connectingblue line) and fit with an error function (solid black line) for the three differentpoints labelled in a. As the coupling j approaches the critical value jc, the changein the gap values to lower frequencies is clearly visible (from panel 1 to panel 3).Vertical dashed lines mark the frequency U/h corresponding to the on-siteinteraction. Each data point results from an average of the temperatures over,50 experimental runs. Error bars, s.e.m.

RESEARCH LETTER

4 5 6 | N A T U R E | V O L 4 8 7 | 2 6 J U L Y 2 0 1 2

Macmillan Publishers Limited. All rights reserved©2012

Manuel Endres, Takeshi Fukuhara, David Pekker, Marc Cheneau, Peter Schaub, Christian Gross, Eugene Demler, Stefan Kuhr, and Immanuel Bloch, Nature 487, 454 (2012).

Observation of Higgs quasi-normal mode across the superfluid-insulator transition of ultracold atoms in a 2-dimensional optical lattice:Response to modulation of lattice depth scales as expected from the LHP pole

D. Podolsky and S. Sachdev, Phy. Rev. B 86, 054508 (2012).

Wednesday, January 16, 13

Page 52: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

1. Higgs quasi-normal mode near the superfluid-insulator transition in 2 dimensions

2. Quantum criticality and conformal field theories

3. Holography and the quasi-normal modes of black-hole horizons

Outline

Wednesday, January 16, 13

Page 53: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

1. Higgs quasi-normal mode near the superfluid-insulator transition in 2 dimensions

2. Quantum criticality and conformal field theories

3. Holography and the quasi-normal modes of black-hole horizons

Outline

Wednesday, January 16, 13

Page 54: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

S =

Zd2rd⌧

⇥|@⌧ |2 + c2|rr |2 + V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

A conformal field theory

in 2+1 spacetime dimensions:

a CFT3

h i 6= 0 h i = 0

Wednesday, January 16, 13

Page 55: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

Wednesday, January 16, 13

Page 56: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

Classical vortices and Goldstone oscillations

Wednesday, January 16, 13

Page 57: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

Classical Boltzmann gas of particles and holes

Wednesday, January 16, 13

Page 58: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

CFT3 at T>0

Wednesday, January 16, 13

Page 59: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Quantum critical dynamics

S. Sachdev, Quantum Phase Transitions, Cambridge (1999).

Quantum “nearly perfect fluid”with shortest possible local equilibration time, ⌧eq

⌧eq = C ~kBT

where C is a universal constant.

Response functions are characterized by poles in LHPwith ! ⇠ kBT/~

(analogs of Higgs quasi-normal mode.)

Wednesday, January 16, 13

Page 60: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

M.P.A. Fisher, G. Grinstein, and S.M. Girvin, Phys. Rev. Lett. 64, 587 (1990) K. Damle and S. Sachdev, Phys. Rev. B 56, 8714 (1997).

� =

Q2

h⇥ [Universal constant O(1) ]

(Q is the “charge” of one boson)

Transport co-oe�cients not determinedby collision rate, but by

universal constants of nature

Conductivity

Quantum critical dynamics

Wednesday, January 16, 13

Page 61: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Transport co-oe�cients not determinedby collision rate, but by

universal constants of nature

Momentum transport�

s⇥ viscosity

entropy density

=�

kB� [Universal constant O(1) ]P. Kovtun, D. T. Son, and A. Starinets, Phys. Rev. Lett. 94, 11601 (2005)

Quantum critical dynamics

Wednesday, January 16, 13

Page 62: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Describe charge transport using Boltzmann theory of in-

teracting bosons:

dv

dt+

v

⇥c= F.

This gives a frequency (⇤) dependent conductivity

�(⇤) =�0

1� i⇤ ⇥c

where ⇥c ⇠ ~/(kBT ) is the time between boson collisions.

Also, we have �(⇤ ! 1) = �1, associated with the den-

sity of states for particle-hole creation (the “optical con-

ductivity”) in the CFT3.

Quantum critical dynamics

Wednesday, January 16, 13

Page 63: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Describe charge transport using Boltzmann theory of in-

teracting bosons:

dv

dt+

v

⇥c= F.

This gives a frequency (⇤) dependent conductivity

�(⇤) =�0

1� i⇤ ⇥c

where ⇥c ⇠ ~/(kBT ) is the time between boson collisions.

Also, we have �(⇤ ! 1) = �1, associated with the den-

sity of states for particle-hole creation (the “optical con-

ductivity”) in the CFT3.

Quantum critical dynamics

Wednesday, January 16, 13

Page 64: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Boltzmann theory of bosons

�0

�1

!

1/�c

Re[�(!)]

Wednesday, January 16, 13

Page 65: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT⇥�⇤ �= 0 ⇥�⇤ = 0

So far, we have described the quantum critical point using

the boson particle and hole excitations of the insulator.

��c

Wednesday, January 16, 13

Page 66: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

However, we could equally well describe the conductivityusing the excitations of the superfluid, which are vortices.

These are quantum particles (in 2+1 dimensions) whichdescribed by a (mirror/e.m.) “dual” CFT3 with an emer-gent U(1) gauge field. Their T > 0 dynamics can also bedescribed by a Boltzmann equation:

Conductivity = Resistivity of vortices

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT⇥�⇤ �= 0 ⇥�⇤ = 0

��c

Wednesday, January 16, 13

Page 67: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT⇥�⇤ �= 0 ⇥�⇤ = 0

However, we could equally well describe the conductivityusing the excitations of the superfluid, which are vortices.

These are quantum particles (in 2+1 dimensions) whichdescribed by a (mirror/e.m.) “dual” CFT3 with an emer-gent U(1) gauge field. Their T > 0 dynamics can also bedescribed by a Boltzmann equation:

Conductivity = Resistivity of vortices

M.P.A. Fisher, Physical Review Letters 65, 923 (1990)

��c

Wednesday, January 16, 13

Page 68: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Boltzmann theory of bosons

�0

�1

!

1/�c

Re[�(!)]

Wednesday, January 16, 13

Page 69: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Boltzmann theory of vortices

�11/�cv

1/�0v

Re[�(!)]

!

Wednesday, January 16, 13

Page 70: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Boltzmann theory of bosons

�0

�1

!

1/�c

Re[�(!)]

Wednesday, January 16, 13

Page 71: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Vector large N expansion for CFT3

K. Damle and S. Sachdev, Phys. Rev. B 56, 8714 (1997).

~!kBT

1

; ⌃ ! a universal function� =Q2

h�

✓~⇥kBT

O(N)

O(1/N)

Re[�(!)]

Wednesday, January 16, 13

Page 72: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Vector large N expansion for CFT3

K. Damle and S. Sachdev, Phys. Rev. B 56, 8714 (1997).

~!kBT

1

; ⌃ ! a universal function� =Q2

h�

✓~⇥kBT

O(N)

O(1/N)

Re[�(!)]

Small ! and vector large Nlimits do not commute

at T > 0.

Wednesday, January 16, 13

Page 73: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Vector large N expansion for CFT3

K. Damle and S. Sachdev, Phys. Rev. B 56, 8714 (1997).

~!kBT

1

; ⌃ ! a universal function� =Q2

h�

✓~⇥kBT

O(N)

O(1/N)

Re[�(!)]

Wednesday, January 16, 13

Page 74: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Vector large N expansion for CFT3

K. Damle and S. Sachdev, Phys. Rev. B 56, 8714 (1997).

~!kBT

1

; ⌃ ! a universal function� =Q2

h�

✓~⇥kBT

O(N)

O(1/N)

Re[�(!)]

Needed: an accurate theory

of quantum critical dynamicsof “long-range” entangled states

at small N

Wednesday, January 16, 13

Page 75: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

1. Higgs quasi-normal mode near the superfluid-insulator transition in 2 dimensions

2. Quantum criticality and conformal field theories

3. Holography and the quasi-normal modes of black-hole horizons

Outline

Wednesday, January 16, 13

Page 76: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

1. Higgs quasi-normal mode near the superfluid-insulator transition in 2 dimensions

2. Quantum criticality and conformal field theories

3. Holography and the quasi-normal modes of black-hole horizons

Outline

Wednesday, January 16, 13

Page 77: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Field theories in d + 1 spacetime dimensions arecharacterized by couplings g which obey the renor-malization group equation

udg

du= �(g)

where u is the energy scale. The RG equation islocal in energy scale, i.e. the RHS does not dependupon u.

Wednesday, January 16, 13

Page 78: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

u

J. McGreevy, arXiv0909.0518

xi

Wednesday, January 16, 13

Page 79: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Key idea: ) Implement r as an extra dimen-

sion, and map to a local theory in d + 2 spacetime

dimensions.

r xi

Wednesday, January 16, 13

Page 80: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

r

Holography

xi

For a relativistic CFT in d spatial dimensions, the

metric in the holographic space is fixed by de-

manding the scale transformation (i = 1 . . . d)

xi ! ⇣xi , t ! ⇣t , ds ! ds

Wednesday, January 16, 13

Page 81: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

r

Holography

xi

This gives the unique metric

ds

2=

1

r

2

��dt

2+ dr

2+ dx

2i

This is the metric of anti-de Sitter space AdSd+2.

Wednesday, January 16, 13

Page 82: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

zr

AdS4R

2,1

Minkowski

CFT3

AdS/CFT correspondence

xi

Wednesday, January 16, 13

Page 83: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

zr

AdS4R

2,1

Minkowski

CFT3

AdS/CFT correspondence

A

xi

Wednesday, January 16, 13

Page 84: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

zr

AdS4R

2,1

Minkowski

CFT3

AdS/CFT correspondence

A

S. Ryu and T. Takayanagi, Phys. Rev. Lett. 96, 18160 (2006).

Associate entanglement entropy with an observer in the enclosed spacetime region, who cannot observe “outside” : i.e. the region is surrounded by an imaginary horizon.

xi

Wednesday, January 16, 13

Page 85: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

zr

AdS4R

2,1

Minkowski

CFT3

AdS/CFT correspondence

A

S. Ryu and T. Takayanagi, Phys. Rev. Lett. 96, 18160 (2006).

The entropy of this region is bounded by its surface area (Bekenstein-Hawking-’t Hooft-Susskind)

xi

Wednesday, January 16, 13

Page 86: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

zr

AdS4R

2,1

Minkowski

CFT3

AdS/CFT correspondence

AMinimal

surface area measures

entanglemententropy

S. Ryu and T. Takayanagi, Phys. Rev. Lett. 96, 18160 (2006).

xi

Wednesday, January 16, 13

Page 87: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

zr

AdS4R

2,1

Minkowski

CFT3

AdS/CFT correspondence

A

S. Ryu and T. Takayanagi, Phys. Rev. Lett. 96, 18160 (2006).

• Computation of minimal surface area yields

SE = aP � �,where � is a shape-dependent universal number.

xi

Wednesday, January 16, 13

Page 88: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

• Entanglement entropy obeys SE = aP � �, where

� is a shape-dependent universal number associated

with the CFT3.

Entanglement entropy from field theory of CFT3

B

M. A. Metlitski, C. A. Fuertes, and S. Sachdev, Physical Review B 80, 115122 (2009).H. Casini, M. Huerta, and R. Myers, JHEP 1105:036, (2011)

I. Klebanov, S. Pufu, and B. Safdi, arXiv:1105.4598

A P

Wednesday, January 16, 13

Page 89: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

zr

AdS4R

2,1

Minkowski

CFT3

AdS/CFT correspondence

xi

Wednesday, January 16, 13

Page 90: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

zrThis emergent spacetime is a solution of Einstein gravity

with a negative cosmological constant

AdS4R

2,1

Minkowski

CFT3

AdS/CFT correspondence

SE =

Zd

4x

p�g

1

22

✓R+

6

L

2

◆�

xi

Wednesday, January 16, 13

Page 91: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

AdS/CFT correspondence at non-zero temperatures

AdS4-Schwarzschild black-brane

There is a family of solutions of Einstein

gravity which describe non-zero

temperatures

r

S =

Zd

4x

p�g

1

22

✓R+

6

L

2

◆�

Wednesday, January 16, 13

Page 92: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

AdS/CFT correspondence at non-zero temperatures

AdS4-Schwarzschild black-brane

ds

2 =

✓L

r

◆2 dr

2

f(r)� f(r)dt2 + dx

2 + dy

2

with f(r) = 1� (r/R)3

r

Black-brane at temperature of

2+1 dimensional quantum critical

systemWednesday, January 16, 13

Page 93: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

AdS/CFT correspondence at non-zero temperatures

AdS4-Schwarzschild black-brane

ds

2 =

✓L

r

◆2 dr

2

f(r)� f(r)dt2 + dx

2 + dy

2

with f(r) = 1� (r/R)3

rA 2+1

dimensional

system at its

quantum

critical point:

kBT =

3~4⇡R

.

Black-brane at temperature of

2+1 dimensional quantum critical

systemWednesday, January 16, 13

Page 94: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Black-brane at temperature of

2+1 dimensional quantum critical

system

Beckenstein-Hawking entropy of black brane

= entropy of CFT3

AdS4-Schwarzschild black-brane

AdS/CFT correspondence at non-zero temperatures

rA 2+1

dimensional

system at its

quantum

critical point:

kBT =

3~4⇡R

.

Wednesday, January 16, 13

Page 95: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Black-brane at temperature of

2+1 dimensional quantum critical

system

Friction of quantum criticality = waves

falling into black brane

AdS4-Schwarzschild black-brane

AdS/CFT correspondence at non-zero temperatures

A 2+1

dimensional

system at its

quantum

critical point:

kBT =

3~4⇡R

.

D. T. SonWednesday, January 16, 13

Page 96: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

AdS4 theory of quantum criticalityMost general e↵ective holographic theory for linear charge

transport with 4 spatial derivatives:

Sbulk =

1

g

2M

Zd

4x

pg

1

4

FabFab

+ �L

2CabcdF

abF

cd

+

Zd

4x

pg

� 1

2

2

✓R+

6

L

2

◆�,

where Cabcd is the Weyl tensor, and Fab is the bulk

gauge flux dual to the conserved current of the CFT.

Three dimensionless parameters, gM , L

2/

2, and �, are

related to CFT3 correlators: �1, the central charge,

and to 3-point correlators of current and stress energy

tensors. Boundary and bulk methods show that |�| 1/12, and the bound is saturated by free fields.

R. C. Myers, S. Sachdev, and A. Singh, Physical Review D 83, 066017 (2011)

D. Chowdhury, S. Raju, S. Sachdev, A. Singh, and P. Strack, arXiv:1210.5247

Wednesday, January 16, 13

Page 97: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

�/T

Conductivity is

independent of ⇥/Tfor � = 0.

1

g2M

C. P. Herzog, P. K. Kovtun, S. Sachdev, and D. T. Son,Phys. Rev. D 75, 085020 (2007).

AdS4 theory of quantum criticality

Wednesday, January 16, 13

Page 98: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

�/T

C. P. Herzog, P. K. Kovtun, S. Sachdev, and D. T. Son,Phys. Rev. D 75, 085020 (2007).

Consequence of self-duality of Maxwell theory in 3+1 dimensions

Conductivity is

independent of ⇥/Tfor � = 0.

1

g2M

AdS4 theory of quantum criticality

Wednesday, January 16, 13

Page 99: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

R. C. Myers, S. Sachdev, and A. Singh, Physical Review D 83, 066017 (2011)

h

Q2�

0.0 0.5 1.0 1.5 2.00.0

0.5

1.0

1.5

4�T

� = 0

� =1

12

� = � 1

12

AdS4 theory of quantum criticality

�(!)

�(1)

Wednesday, January 16, 13

Page 100: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

h

Q2�

0.0 0.5 1.0 1.5 2.00.0

0.5

1.0

1.5

4�T

� = 0

� =1

12

� = � 1

12 • The � > 0 result has similarities to

the quantum-Boltzmann result for

transport of particle-like excitations

R. C. Myers, S. Sachdev, and A. Singh, Physical Review D 83, 066017 (2011)

�(!)

�(1)

AdS4 theory of quantum criticality

Wednesday, January 16, 13

Page 101: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

h

Q2�

0.0 0.5 1.0 1.5 2.00.0

0.5

1.0

1.5

4�T

� = 0

� =1

12

� = � 1

12• The � < 0 result can be interpreted

as the transport of vortex-like

excitations

R. C. Myers, S. Sachdev, and A. Singh, Physical Review D 83, 066017 (2011)

�(!)

�(1)

AdS4 theory of quantum criticality

Wednesday, January 16, 13

Page 102: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

h

Q2�

0.0 0.5 1.0 1.5 2.00.0

0.5

1.0

1.5

4�T

� = 0

� =1

12

� = � 1

12

R. C. Myers, S. Sachdev, and A. Singh, Physical Review D 83, 066017 (2011)

• The � = 0 case is the exact result for the large N limit

of SU(N) gauge theory with N = 8 supersymmetry (the

ABJM model). The ⇥-independence is a consequence of

self-duality under particle-vortex duality (S-duality).

�(!)

�(1)

AdS4 theory of quantum criticality

Wednesday, January 16, 13

Page 103: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

h

Q2�

0.0 0.5 1.0 1.5 2.00.0

0.5

1.0

1.5

4�T

� = 0

� =1

12

� = � 1

12

R. C. Myers, S. Sachdev, and A. Singh, Physical Review D 83, 066017 (2011)

• Stability constraints on the e�ectivetheory (|�| < 1/12) allow only a lim-ited ⇥-dependence in the conductivity

�(!)

�(1)

AdS4 theory of quantum criticality

Wednesday, January 16, 13

Page 104: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

(a)<{�(w; � = 1/12)} (b)<{�(w; � = 1/12)}

(c)<{�(w; � = �1/12)} (d)<{�(w; � = �1/12)}

FIG. 4. Conductivity � and its S-dual � = 1/� in the LHP, w00 = =w 0, for |�| = 1/12. The zerosof �(w; �) are the poles of �(w; �). We further note the qualitative correspondence between the poles of�(w; �) and the zeros of �(w;��).

low-frequency behavior is dictated by a Drude pole, located closest to the origin. The numerical

solution also shows the presence of satellite poles, the two dominant ones being shown. These

are symmetrically distributed about the =w axis as required by time-reversal, and are essential

to capture the behavior of � beyond the small frequency limit. In contrast, the conductivity at

� = �1/12 in Fig. 4(c) shows a minimum at w = 0 on the real axis, see also Fig. 7(b) for a plot

restricted to real frequencies. The corresponding pole structure shows no poles on the imaginary

axis, in particular no Drude pole. The conductivity at � = �1/12 is said to be vortex-like because

it can be put in correspondence with the conductivity of the CFT S-dual to the one with � = 1/12,

as we now explain.

B. S-duality and conductivity zeros

Great insight into the behavior of the conductivity can be gained by means of S-duality, a

generalization of the familiar particle-vortex duality of the O(2) model. S-duality on the boundary

14

AdS4 theory of quantum criticality

W. Witzack-Krempa and S. Sachdev, Physical Review D 86, 235115 (2012)

Poles in LHPof conductivityat ! ⇠ kBT/~;

analog ofHiggs quasinormal mode–

quasinormal modesof black brane

Wednesday, January 16, 13

Page 105: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

(a)<{�(w; � = 1/12)} (b)<{�(w; � = 1/12)}

(c)<{�(w; � = �1/12)} (d)<{�(w; � = �1/12)}

FIG. 4. Conductivity � and its S-dual � = 1/� in the LHP, w00 = =w 0, for |�| = 1/12. The zerosof �(w; �) are the poles of �(w; �). We further note the qualitative correspondence between the poles of�(w; �) and the zeros of �(w;��).

low-frequency behavior is dictated by a Drude pole, located closest to the origin. The numerical

solution also shows the presence of satellite poles, the two dominant ones being shown. These

are symmetrically distributed about the =w axis as required by time-reversal, and are essential

to capture the behavior of � beyond the small frequency limit. In contrast, the conductivity at

� = �1/12 in Fig. 4(c) shows a minimum at w = 0 on the real axis, see also Fig. 7(b) for a plot

restricted to real frequencies. The corresponding pole structure shows no poles on the imaginary

axis, in particular no Drude pole. The conductivity at � = �1/12 is said to be vortex-like because

it can be put in correspondence with the conductivity of the CFT S-dual to the one with � = 1/12,

as we now explain.

B. S-duality and conductivity zeros

Great insight into the behavior of the conductivity can be gained by means of S-duality, a

generalization of the familiar particle-vortex duality of the O(2) model. S-duality on the boundary

14

AdS4 theory of quantum criticality

W. Witzack-Krempa and S. Sachdev, Physical Review D 86, 235115 (2012)

Poles in LHPof resistivity —

quasinormal modesof S-dual theory

Wednesday, January 16, 13

Page 106: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

6

The holographic solutions for the conductivity satisfy two

sum rules, valid for all CFT3s. (W. Witzack-Krempa and

S. Sachdev, Phys. Rev. B 86, 235115 (2012))

Z 1

0d!Re [�(!)� �(1)] = 0

Z 1

0d!Re

1

�(!)� 1

�(1)

�= 0

The second rule follows from the existence of a EM-dual

CFT3.

Boltzmann theory chooses a “particle” basis: this satis-

fies only one sum rule but not the other.

Holographic theory satisfies both sum rules.

AdS4 theory of quantum criticality

Wednesday, January 16, 13

Page 107: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Traditional CMT

Identify quasiparticles and their dispersions

Compute scattering matrix elements of quasiparticles (or of collective modes)

These parameters are input into a quantum Boltzmann equation

Deduce dissipative and dynamic properties at non-zero temperatures

Wednesday, January 16, 13

Page 108: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Traditional CMT

Identify quasiparticles and their dispersions

Compute scattering matrix elements of quasiparticles (or of collective modes)

These parameters are input into a quantum Boltzmann equation

Deduce dissipative and dynamic properties at non-zero temperatures

Wednesday, January 16, 13

Page 109: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Traditional CMT

Identify quasiparticles and their dispersions

Compute scattering matrix elements of quasiparticles (or of collective modes)

These parameters are input into a quantum Boltzmann equation

Deduce dissipative and dynamic properties at non-zero temperatures

Wednesday, January 16, 13

Page 110: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Traditional CMT

Identify quasiparticles and their dispersions

Compute scattering matrix elements of quasiparticles (or of collective modes)

These parameters are input into a quantum Boltzmann equation

Deduce dissipative and dynamic properties at non-zero temperatures

Wednesday, January 16, 13

Page 111: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Start with strongly interacting CFT without particle- or wave-like excitations

Compute OPE co-efficients of operators of the CFT

Relate OPE co-efficients to couplings of an effective graviational theory on AdS

Solve Einsten-Maxwell equations. Dynamics of quasi-normal modes of black branes.

Traditional CMT Holography and black-branes

Identify quasiparticles and their dispersions

Compute scattering matrix elements of quasiparticles (or of collective modes)

These parameters are input into a quantum Boltzmann equation

Deduce dissipative and dynamic properties at non-zero temperatures

Wednesday, January 16, 13

Page 112: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Start with strongly interacting CFT without particle- or wave-like excitations

Compute OPE co-efficients of operators of the CFT

Relate OPE co-efficients to couplings of an effective graviational theory on AdS

Solve Einsten-Maxwell equations. Dynamics of quasi-normal modes of black branes.

Traditional CMT

Identify quasiparticles and their dispersions

Compute scattering matrix elements of quasiparticles (or of collective modes)

These parameters are input into a quantum Boltzmann equation

Deduce dissipative and dynamic properties at non-zero temperatures

Holography and black-branes

Wednesday, January 16, 13

Page 113: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Start with strongly interacting CFT without particle- or wave-like excitations

Compute OPE co-efficients of operators of the CFT

Relate OPE co-efficients to couplings of an effective graviational theory on AdS

Solve Einsten-Maxwell equations. Dynamics of quasi-normal modes of black branes.

Traditional CMT

Identify quasiparticles and their dispersions

Compute scattering matrix elements of quasiparticles (or of collective modes)

These parameters are input into a quantum Boltzmann equation

Deduce dissipative and dynamic properties at non-zero temperatures

Holography and black-branes

Wednesday, January 16, 13

Page 114: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Traditional CMT

Identify quasiparticles and their dispersions

Compute scattering matrix elements of quasiparticles (or of collective modes)

These parameters are input into a quantum Boltzmann equation

Deduce dissipative and dynamic properties at non-zero temperatures

Start with strongly interacting CFT without particle- or wave-like excitations

Compute OPE co-efficients of operators of the CFT

Relate OPE co-efficients to couplings of an effective graviational theory on AdS

Solve Einsten-Maxwell equations. Dynamics of quasi-normal modes of black branes.

Holography and black-branes

Wednesday, January 16, 13

Page 115: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Gapped quantum matter Spin liquids, quantum Hall states

Conformal quantum matter Quantum critical points in antiferromagnets, superconductors, and ultracold atoms; graphene

Compressible quantum matter Strange metals in high temperature superconductors, Bose metals

S. Sachdev, 100th anniversary Solvay conference,arXiv:1203.4565

“Complex entangled” states of quantum matter,

not adiabatically connected to independent particle states

Wednesday, January 16, 13

Page 116: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Gapped quantum matter Spin liquids, quantum Hall states

Conformal quantum matter Quantum critical points in antiferromagnets, superconductors, and ultracold atoms; graphene

Compressible quantum matter Strange metals in high temperature superconductors, Bose metals

S. Sachdev, 100th anniversary Solvay conference,arXiv:1203.4565

“Complex entangled” states of quantum matter,

not adiabatically connected to independent particle states

Wednesday, January 16, 13

Page 117: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

BaFe2(As1�x

Px

)2

K. Hashimoto, K. Cho, T. Shibauchi, S. Kasahara, Y. Mizukami, R. Katsumata, Y. Tsuruhara, T. Terashima, H. Ikeda, M. A. Tanatar, H. Kitano, N. Salovich, R. W. Giannetta, P. Walmsley, A. Carrington, R. Prozorov, and Y. Matsuda, Science 336, 1554 (2012).

Resistivity⇠ ⇢0 +ATn

FL

Wednesday, January 16, 13

Page 118: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Logarithmic violation of “area law”: SE / (kFP ) ln(kFP )

for a circular Fermi surface with Fermi momentum kF ,where P is the perimeter of region A.

The coe�cient is independent of the shape of A.

B

A

Entanglement entropy of Fermi liquids (FL) and non-Fermi liquids (non-FL)

P

D. Gioev and I. Klich, Physical Review Letters 96, 100503 (2006)B. Swingle, Physical Review Letters 105, 050502 (2010)

Y. Zhang, T. Grover, and A. Vishwanath, Physical Review Letters 107, 067202 (2011)Wednesday, January 16, 13

Page 119: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Logarithmic violation of “area law”: SE / (kFP ) ln(kFP )

for a circular Fermi surface with Fermi momentum kF ,where P is the perimeter of region A.

The coe�cient is independent of the shape of A.

B

D. Gioev and I. Klich, Physical Review Letters 96, 100503 (2006)B. Swingle, Physical Review Letters 105, 050502 (2010)

Y. Zhang, T. Grover, and A. Vishwanath, Physical Review Letters 107, 067202 (2011)

A P

Entanglement entropy of Fermi liquids (FL) and non-Fermi liquids (non-FL)

Wednesday, January 16, 13

Page 120: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Logarithmic violation of “area law”: SE / (kFP ) ln(kFP )

for a circular Fermi surface with Fermi momentum kF ,where P is the perimeter of region A.

The coe�cient is independent of the shape of A.

B

A P

D. Gioev and I. Klich, Physical Review Letters 96, 100503 (2006)B. Swingle, Physical Review Letters 105, 050502 (2010)

Y. Zhang, T. Grover, and A. Vishwanath, Physical Review Letters 107, 067202 (2011)

Entanglement entropy of Fermi liquids (FL) and non-Fermi liquids (non-FL)

Wednesday, January 16, 13

Page 121: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

r

Holography

xi

Wednesday, January 16, 13

Page 122: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

r

Holography

xi

Consider a metric which transforms under rescaling as

xi ! ⇣ xi, t ! ⇣

zt, ds ! ⇣

✓/dds.

Recall: conformal matter has ✓ = 0, z = 1, and the metric is

anti-de Sitter

Wednesday, January 16, 13

Page 123: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

r

Holography

xi

Consider a metric which transforms under rescaling as

xi ! ⇣ xi, t ! ⇣

zt, ds ! ⇣

✓/dds.

Recall: conformal matter has ✓ = 0, z = 1, and the metric is

anti-de Sitter

The value ✓ = d�1 reproduces all the essential characteristicsof the entropy and entanglement entropy of a non-FL.

L. Huijse, S. Sachdev, B. Swingle, Physical Review B 85, 035121 (2012)N. Ogawa, T. Takayanagi, and T. Ugajin, JHEP 1201, 125 (2012).

Wednesday, January 16, 13

Page 124: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

r

Holography

xi

L. Huijse, S. Sachdev, B. Swingle, Physical Review B 85, 035121 (2012)N. Ogawa, T. Takayanagi, and T. Ugajin, JHEP 1201, 125 (2012).

Consider a metric which transforms under rescaling as

xi ! ⇣ xi, t ! ⇣

zt, ds ! ⇣

✓/dds.

Recall: conformal matter has ✓ = 0, z = 1, and the metric is

anti-de Sitter

The null-energy condition of gravity yields z � 1+ ✓/d. In d = 2, this

leads to z � 3/2. Field theory on non-FL yields z = 3/2 to 3 loops!

P. A. Lee, Phys. Rev. Lett. 63, 680 (1989)B. Blok and H. Monien, Phys. Rev. B 47, 3454 (1993)

M. A. Metlitski and S. Sachdev, Phys. Rev. B 82, 075127 (2010)

Wednesday, January 16, 13

Page 125: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Conclusions

Realizations of many-particle entanglement:

Z2 spin liquids and conformal quantum critical points

Wednesday, January 16, 13

Page 126: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

New insights and solvable models for diffusion and transport of strongly interacting systems near quantum critical points

The description is far removed from, and complementary to, that of the quantum Boltzmann equation which builds on the quasiparticle/vortex picture.

Good prospects for experimental tests of frequency-dependent, non-linear, and non-equilibrium transport

Conclusions

Conformal quantum matter

Wednesday, January 16, 13

Page 127: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Conclusions

More complex examples in metallic states are experimentally

ubiquitous, but pose difficult strong-coupling problems to conventional methods of field

theory

Wednesday, January 16, 13

Page 128: Entanglement, holography, and the quantum phases of matterqpt.physics.harvard.edu/talks/bonn13.pdf · Holography and the quasi-normal modes of black-hole horizons Outline Wednesday,

Conclusions

String theory and gravity in emergent dimensions

offer a remarkable new approach to describing states with many-particle quantum entanglement.

Much recent progress offers hope of a holographic description of “strange metals”

Wednesday, January 16, 13