quantum criticality in condensed matter

99
Quantum criticality in condensed matter HARVARD Inaugural Symposium of the Wolgang Pauli Center Hamburg, April 17, 2013 Subir Sachdev Scientific American 308, 44 (January 2013) Monday, April 29, 13

Upload: others

Post on 10-May-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Quantum criticality in condensed matter

Quantum criticality in

condensed matter

HARVARD

Inaugural Symposium of theWolgang Pauli Center

Hamburg, April 17, 2013

Subir Sachdev

Scientific American 308, 44 (January 2013)

Monday, April 29, 13

Page 2: Quantum criticality in condensed matter

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

connected to independent electron states

E

MetalMetal

carryinga current

InsulatorSuperconductor

k

E

MetalMetal

carryinga current

InsulatorSuperconductor

k

Metals

Monday, April 29, 13

Page 3: Quantum criticality in condensed matter

E

MetalMetal

carryinga current

InsulatorSuperconductor

k

E

MetalMetal

carryinga current

InsulatorSuperconductor

k

Boltzmann-Landau theory of dynamics of metals:

Long-lived quasiparticles (and quasiholes) have weak interactions which can be described by a Boltzmann equation

Metals

Monday, April 29, 13

Page 4: Quantum criticality in condensed matter

Modern phases of quantum matterNot adiabatically connected

to independent electron states:many-particle

quantum entanglement,and no quasiparticles

Monday, April 29, 13

Page 5: Quantum criticality in condensed matter

1. Superfluid-insulator transition of ultracold atoms in optical lattices:

Conformal field theories and gauge-gravity duality

2. Metals with antiferromagnetism, and high temperature superconductivity The pnictides and the cuprates

Outline

Monday, April 29, 13

Page 6: Quantum criticality in condensed matter

1. Superfluid-insulator transition of ultracold atoms in optical lattices:

Conformal field theories and gauge-gravity duality

2. Metals with antiferromagnetism, and high temperature superconductivity The pnictides and the cuprates

Outline

Monday, April 29, 13

Page 7: Quantum criticality in condensed matter

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

Ultracold 87Rbatoms - bosons

Superfluid-insulator transition

Monday, April 29, 13

Page 8: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

h i 6= 0 h i = 0

Monday, April 29, 13

Page 9: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

h i 6= 0 h i = 0

! a complex field representing the

Bose-Einstein condensate of the superfluid

Monday, April 29, 13

Page 10: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

h i 6= 0 h i = 0

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Monday, April 29, 13

Page 11: Quantum criticality in condensed matter

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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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.

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Monday, April 29, 13

Page 12: Quantum criticality in condensed matter

Insulator (the vacuum) at large repulsion between bosons

Monday, April 29, 13

Page 13: Quantum criticality in condensed matter

Particles ⇠ †

Excitations of the insulator:

Monday, April 29, 13

Page 14: Quantum criticality in condensed matter

Holes ⇠

Excitations of the insulator:

Monday, April 29, 13

Page 15: Quantum criticality in condensed matter

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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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.

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Monday, April 29, 13

Page 16: Quantum criticality in condensed matter

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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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 | |.

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Monday, April 29, 13

Page 17: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

A conformal field theory

in 2+1 spacetime dimensions:

a CFT3

h i 6= 0 h i = 0

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Monday, April 29, 13

Page 18: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

Quantum state with

complex, many-body,

“long-range” quantum entanglement

h i 6= 0 h i = 0

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Monday, April 29, 13

Page 19: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

No well-defined normal modes,

or particle-like excitations

h i 6= 0 h i = 0

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Monday, April 29, 13

Page 20: Quantum criticality in condensed matter

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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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.

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Monday, April 29, 13

Page 21: Quantum criticality in condensed matter

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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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.

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

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

Monday, April 29, 13

Page 22: Quantum criticality in condensed matter

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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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).

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Monday, April 29, 13

Page 23: Quantum criticality in condensed matter

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~ 3ffiffiffi2p

{4" #

1zj=jcð Þ$ %1=2

j=jc{1ð Þ1=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 12ffiffiffi2p

{17" #

j=jc$ %1=2

1{j=jcð Þ1=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).

Monday, April 29, 13

Page 24: Quantum criticality in condensed matter

3

finite-temperature kernel, N(⌧,!) = e�!⌧ + e�!(1/T�⌧):

�(⌧) =

Z +1

0N(⌧,!)S(!) . (4)

We employ the same protocol of collecting and analyzingdata as in Ref. [13]. More specifically, in the MC simu-lation we collect statistics for the correlation function atMatsubara frequencies !n = 2⇡Tn with integer n

�(i!n) = hK(⌧)K(0)ii!n + hKi (5)

which is related to �(⌧) by a Fourier transform. In thepath integral representation, �(i!n) has a direct unbi-ased estimator, |

Pk e

i!n⌧k |2, where the sum runs overall hopping transitions in a given configuration. Once�(⌧) is recovered from �(i!n), the analytical continuationmethods described in Ref. [13] are applied to extract thespectral function S(!). A discussion on the reproducibil-ity of the analytically continued results for this type ofproblems can also be found in Ref. [13].

We consider system sizes significantly larger than thecorrelation length by a factor of at least four to ensurethat our results are e↵ectively in the thermodynamiclimit. Furthermore, for the SF and MI phases, we setthe temperature T = 1/� to be much smaller than thecharacteristic Higgs energy, so that no details in the rel-evant energy part of spectral function are missed.

We consider two paths in the SF phase to approach theQCP: by increasing the interaction U ! Uc at unity fill-ing factor n = 1 (trajectory i perpendicular to the phaseboundary in Fig 1), and by increasing µ ! µc while keep-ing U = Uc constant (trajectory ii tangential to the phaseboundary in Fig 1). We start with trajectory i by consid-ering three parameter sets for (|g|, L,�): (0.2424, 20, 10),(0.0924, 40, 20, and (0.0462, 80, 40). The prime data areshown in the inset of Fig. 3. After rescaling results ac-cording to Eq. (1) we observe data collapse shown in themain panel of Fig.3. This defines the universal spectralfunction in the superfluid phase, �SF.

When approaching the QCP along trajectory ii,with (|gµ|, L,�) = (0.40, 25, 15), (0.30, 30, 15), and(0.20, 40, 20) we observe a similar data collapse and arriveat the same universal function �SF, see Fig. 4. The fi-nal match is possible only when the characteristic energyscale �(gµ) = C�(g(gµ)) involves a factor of C = 1.2.

The universal spectral function �SF has three distinctfeatures: a) A pronounced peak at !H/� ⇡ 3.3, whichis associated with the Higgs resonance. Since the peak’swidth �/� ⇡ 1 is comparable to its energy, the Higgsmode is strongly damped. It can be identified as awell-defined particle only in a moving reference frame;b) A minimum and another broad maximum between!/� 2 [5, 25] which may originate from multi-Higgs ex-citations [13]; c) The onset of the quantum critical quasi-plateau, in agreement with the scaling hypothesis (1),starting at !/� ⇡ 25. These features are captured by an

0

0.5

1

1.5

2

2.5

3

0 5 10 15 20 25 30 35 40

[Δ(-g

)/J]2/

ν−3 S SF

(ω)/J

ω/Δ(-g)

|g|=0.2424|g|=0.0924|g|=0.0462

0.5

1

1.5

2

2 4 6 8

S SF(ω

)/J

ω/J

FIG. 3. (Color online) Collapse of spectral functions for dif-ferent values of U along trajectory i in the SF phase, labeledby the detuning g = (U � Uc)/J . Inset: original data forSSF(!).

0

0.5

1

1.5

2

2.5

0 5 10 15 20 25 30 35 40

[(Δ(g

µ)/J

]2/ν−3 S SF

(ω)/J

ω/Δ(gµ)

|gµ|=0.40|gµ|=0.30|gµ|=0.20

0.5

1

1.5

2 4 6 8

S SF(ω

)/J

ω/J

FIG. 4. (Color online) Collapse of spectral functions for dif-ferent µ along trajectory ii in the SF phase, labeled by thedetuning gµ = (µ� µc)/J . Inset: original data for SSF(!).

approximate analytic expression with normalized �2 ⇠ 1,

�SF(x) =0.65x3

35 + x2/⌫

1 +

7 sin(0.55x)

1 + 0.02x3

�(6)

We only claim that a plateau is consistent with our imag-inary time data and emerges from the analytic continua-tion procedure which penalizes gradients in the spectralfunction; i.e., other analytic continuation methods mayproduce a di↵erent (oscillating) behavior in the same fre-quency range that is also consistent with the imaginarytime data.We now switch to the MI phase, where we approach the

QCP along trajectory iii in Fig 1. The scaling hypothesisfor the spectral function has a similar structure to the onein Eq. (1),

SMI(!) / �3�2/⌫�MI(!

�) . (7)

The low-energy behavior of �MI starts with the thresh-old singularity at the particle-hole gap value, �MI(x) ⇡

4

of the susceptibility was predicted [12, 13, 17] to be

�00� ⇠ (!/�)3 , ! ⌧ � ⌧ 1. (9)

The !3 rise is due to the decay of a Higgs mode into apair of Goldstone modes. On the other hand, Fig. 4 doesnot display a clear !3 low frequency tail. An alterna-tive method to look for this tail exists, without the needto analytically continue the numerical data to real time.Equation (9) transforms into the large imaginary timeasymptotics �s (⌧) ⇠ 1/⌧4.

For N = 3 we confirm the asymptotic behavior of 1/⌧4

within the numerical limitations. Interestingly, forN = 2we do not find a conclusive asymptotic fall-o↵ as 1/⌧4,Instead, the data fits better to an exponential decay, asin the disordered phase (see Eq. (7)), although the powerlaw might have a small amplitude below our statisticalerrors. In both cases, we can safely conclude that thespectral weight of the Higgs peak dominates over the lowfrequency !3 tail, enhancing its visibility. The large ⌧analysis is discussed elsewhere [23].

Discussion and Summary– Our results are directly ap-plicable to all experimental probes that couple to a func-tion of the order parameter magnitude. For example, thelattice potential amplitude in the trapped bosons sys-tem [16, 18], or pump-probe spectroscopy in Charge Den-sity Wave systems [6–8]. Such a probe can be expandednear criticality in terms of the order parameter fields andtheir derivatives,

⇥(x, ⌧) = ↵|~�|2 + �|@µ~�|2 + �(|~�|2)2 + . . . (10)

So long as ↵ 6= 0, the first term is more relevant than therest. Hence, the scalar susceptibility defined in Eq. (2)dominates the experimental response at low frequenciesand wave vectors.

In summary, we have calculated the scalar susceptibil-ity for relativistic O(2) and O(3) models in 2+1 dimen-

FIG. 4: Rescaled spectral function vs. !/� for N = 2, 3,at µ = 0.5. At low values of !/�, these curves collapse tothe universal scaling function �00

�(!/�), in accordance withEq. 5.

sions near criticality. We have demonstrated that theHiggs mode appears as a universal spectral feature sur-viving all the way to the quantum critical point. Sincethis is a strongly coupling fixed point, the existence ofa well defined mode that is not protected by symmetryis an interesting, not obvious, result. We presented newuniversal quantities to be compared with experimentalresults.

Acknowledgements.– We are very grateful to NikolayProkof’ev and Lode Pollet for helpful comments and wealso thank Dan Arovas, Netanel Lindner, Subir Sachdevand Manuel Endres for helpful discussions. AA and DPacknowledge support from the Israel Science Founda-tion, the European Union under grant IRG-276923, theU.S.-Israel Binational Science Foundation, and thank theAspen Center for Physics, supported by the NSF-PHY-1066293, for its hospitality. SG acknowledges the Clorefoundation for a fellowship.

[1] C. Varma, Journal of Low Temperature Physics 126, 901(2002).

[2] S. D. Huber, E. Altman, H. P. Buechler, and G. Blatter,Physical Review B 75, 085106 (2007).

[3] R. Sooryakumar and M. Klein, Physical Review Letters45, 660 (1980).

[4] P. Littlewood and C. Varma, Physical Review B 26, 4883(1982).

[5] C. Ruegg et al., Physical Review Letters 100, 205701(2008).

[6] Y. Ren, Z. Xu, and G. Lupke, The Journal of ChemicalPhysics 120, 4755 (2004).

[7] J. P. Pouget, B. Hennion, C. Escribe-Filippini, and M.Sato, Phys. Rev. B 43, 8421 (1991).

[8] R. Yusupov et al., Nature Physics (2010).[9] I. A✏eck and G. F. Wellman, Physical Review B 46,

8934 (1992).[10] K. G. Wilson and M. E. Fisher, Phys. Rev. Lett. 28, 240

(1972).[11] N. Lindner and A. Auerbach, Physical Review B 81,

054512 (2010).[12] D. Podolsky, A. Auerbach, and D. P. Arovas, Physical

Review B 84, 174522 (2011).[13] S. Sachdev, Physical Review B 59, 14054 (1999).[14] W. Zwerger, Physical Review Letters 92, 027203 (2004).[15] N. Dupuis, Phys. Rev. A 80, 043627 (2009).[16] M. Endres et al., Nature 487, 454 (2012).[17] D. Podolsky and S. Sachdev, Phys. Rev. B 86, 054508

(2012).[18] L. Pollet and N. Prokof’ev, Physical Review Letters 109,

10401 (2012).[19] S. Sachdev, Quantum Phase Transitions, second edition

ed. (Cambridge University Press, New York, 2011).[20] M. P. A. Fisher, P. B. Weichman, G. Grinstein, and D. S.

Fisher, Physical Review B 40, 546 (1989).[21] F. D. M. Haldane, Physical Review Letters 50, 1153

(1983).[22] S. Chakravarty, B. I. Halperin, and D. R. Nelson, Physi-

cal Review B 39, 2344 (1989).

4

of the susceptibility was predicted [12, 13, 17] to be

�00� ⇠ (!/�)3 , ! ⌧ � ⌧ 1. (9)

The !3 rise is due to the decay of a Higgs mode into apair of Goldstone modes. On the other hand, Fig. 4 doesnot display a clear !3 low frequency tail. An alterna-tive method to look for this tail exists, without the needto analytically continue the numerical data to real time.Equation (9) transforms into the large imaginary timeasymptotics �s (⌧) ⇠ 1/⌧4.

For N = 3 we confirm the asymptotic behavior of 1/⌧4

within the numerical limitations. Interestingly, forN = 2we do not find a conclusive asymptotic fall-o↵ as 1/⌧4,Instead, the data fits better to an exponential decay, asin the disordered phase (see Eq. (7)), although the powerlaw might have a small amplitude below our statisticalerrors. In both cases, we can safely conclude that thespectral weight of the Higgs peak dominates over the lowfrequency !3 tail, enhancing its visibility. The large ⌧analysis is discussed elsewhere [23].

Discussion and Summary– Our results are directly ap-plicable to all experimental probes that couple to a func-tion of the order parameter magnitude. For example, thelattice potential amplitude in the trapped bosons sys-tem [16, 18], or pump-probe spectroscopy in Charge Den-sity Wave systems [6–8]. Such a probe can be expandednear criticality in terms of the order parameter fields andtheir derivatives,

⇥(x, ⌧) = ↵|~�|2 + �|@µ~�|2 + �(|~�|2)2 + . . . (10)

So long as ↵ 6= 0, the first term is more relevant than therest. Hence, the scalar susceptibility defined in Eq. (2)dominates the experimental response at low frequenciesand wave vectors.

In summary, we have calculated the scalar susceptibil-ity for relativistic O(2) and O(3) models in 2+1 dimen-

FIG. 4: Rescaled spectral function vs. !/� for N = 2, 3,at µ = 0.5. At low values of !/�, these curves collapse tothe universal scaling function �00

�(!/�), in accordance withEq. 5.

sions near criticality. We have demonstrated that theHiggs mode appears as a universal spectral feature sur-viving all the way to the quantum critical point. Sincethis is a strongly coupling fixed point, the existence ofa well defined mode that is not protected by symmetryis an interesting, not obvious, result. We presented newuniversal quantities to be compared with experimentalresults.

Acknowledgements.– We are very grateful to NikolayProkof’ev and Lode Pollet for helpful comments and wealso thank Dan Arovas, Netanel Lindner, Subir Sachdevand Manuel Endres for helpful discussions. AA and DPacknowledge support from the Israel Science Founda-tion, the European Union under grant IRG-276923, theU.S.-Israel Binational Science Foundation, and thank theAspen Center for Physics, supported by the NSF-PHY-1066293, for its hospitality. SG acknowledges the Clorefoundation for a fellowship.

[1] C. Varma, Journal of Low Temperature Physics 126, 901(2002).

[2] S. D. Huber, E. Altman, H. P. Buechler, and G. Blatter,Physical Review B 75, 085106 (2007).

[3] R. Sooryakumar and M. Klein, Physical Review Letters45, 660 (1980).

[4] P. Littlewood and C. Varma, Physical Review B 26, 4883(1982).

[5] C. Ruegg et al., Physical Review Letters 100, 205701(2008).

[6] Y. Ren, Z. Xu, and G. Lupke, The Journal of ChemicalPhysics 120, 4755 (2004).

[7] J. P. Pouget, B. Hennion, C. Escribe-Filippini, and M.Sato, Phys. Rev. B 43, 8421 (1991).

[8] R. Yusupov et al., Nature Physics (2010).[9] I. A✏eck and G. F. Wellman, Physical Review B 46,

8934 (1992).[10] K. G. Wilson and M. E. Fisher, Phys. Rev. Lett. 28, 240

(1972).[11] N. Lindner and A. Auerbach, Physical Review B 81,

054512 (2010).[12] D. Podolsky, A. Auerbach, and D. P. Arovas, Physical

Review B 84, 174522 (2011).[13] S. Sachdev, Physical Review B 59, 14054 (1999).[14] W. Zwerger, Physical Review Letters 92, 027203 (2004).[15] N. Dupuis, Phys. Rev. A 80, 043627 (2009).[16] M. Endres et al., Nature 487, 454 (2012).[17] D. Podolsky and S. Sachdev, Phys. Rev. B 86, 054508

(2012).[18] L. Pollet and N. Prokof’ev, Physical Review Letters 109,

10401 (2012).[19] S. Sachdev, Quantum Phase Transitions, second edition

ed. (Cambridge University Press, New York, 2011).[20] M. P. A. Fisher, P. B. Weichman, G. Grinstein, and D. S.

Fisher, Physical Review B 40, 546 (1989).[21] F. D. M. Haldane, Physical Review Letters 50, 1153

(1983).[22] S. Chakravarty, B. I. Halperin, and D. R. Nelson, Physi-

cal Review B 39, 2344 (1989).

Snir Gazit, Daniel Podolsky,

and Assa Auerbach,

arXiv:1212.3759

Kun Chen, Longxiang Liu,

Youjin Deng, Lode Pollet,

and Nikolay Prokof’ev,

arXiv:1301.3139

Observation of Higgs quasi-normal mode in quantum Monte Carlo

Scaling of spectral response

functions predicted in

D. Podolsky and S. Sachdev,

Phys. Rev. B 86, 054508 (2012).

Monday, April 29, 13

Page 25: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

A conformal field theory

in 2+1 spacetime dimensions:

a CFT3

h i 6= 0 h i = 0

S =

Zd2rdt

⇥|@t |2 � c2|rr |2 � V ( )

V ( ) = (�� �c)| |2 + u�| |2

�2

Monday, April 29, 13

Page 26: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

Monday, April 29, 13

Page 27: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

“Boltzmann” theory of Nambu-

Goldstone and vortices

Boltzmann theory of

particles/holes

Monday, April 29, 13

Page 28: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

CFT3 at T>0

Monday, April 29, 13

Page 29: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

CFT3 at T>0

Boltzmann theory of particles/holes/vortices

does not apply

Monday, April 29, 13

Page 30: Quantum criticality in condensed matter

g

T

gc

0

InsulatorSuperfluid

Quantumcritical

TKT

��c

CFT3 at T>0

Needed: Accurate theory of

quantum critical dynamics

Monday, April 29, 13

Page 31: Quantum criticality in condensed matter

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/~.

These poles (quasi-normal modes) appear naturally inthe holographic theory.

(Analogs of Higgs quasi-normal mode.)

Monday, April 29, 13

Page 32: Quantum criticality in condensed matter

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)

Quantum critical dynamics

Transport co-oe�cients not determined

by collision rate of quasiparticles, but by

fundamental constants of nature

Conductivity

Monday, April 29, 13

Page 33: Quantum criticality in condensed matter

J. McGreevy, arXiv0909.0518

xir

Renormalization group: ) Follow coupling

constants of quantum many body theory as a func-

tion of length scale r

Monday, April 29, 13

Page 34: Quantum criticality in condensed matter

Key idea: ) Implement r as an extra dimen-

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

dimensions.

r xi

Renormalization group: ) Follow coupling

constants of quantum many body theory as a func-

tion of length scale r

J. McGreevy, arXiv0909.0518

Monday, April 29, 13

Page 35: Quantum criticality in condensed matter

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

Monday, April 29, 13

Page 36: Quantum criticality in condensed matter

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.

Monday, April 29, 13

Page 37: Quantum criticality in condensed matter

zr

AdS4R

2,1

Minkowski

CFT3

AdS/CFT correspondence

SE =

Zd

4x

p�g

1

22

✓R+

6

L

2

◆�

xi

This emergent spacetime is a solution of Einstein gravity with a negative cosmological constant

Monday, April 29, 13

Page 38: Quantum criticality in condensed matter

There is a family of

solutions of Einstein

gravity which describe non-zero

temperatures

r

Gauge-gravity duality at non-zero temperatures

SE =

Zd

4x

p�g

1

22

✓R+

6

L

2

◆�

Monday, April 29, 13

Page 39: Quantum criticality in condensed matter

Gauge-gravity duality at non-zero temperatures

A 2+1 dimensional

system at T > 0with

couplings at its quantum critical point

r

SE =

Zd

4x

p�g

1

22

✓R+

6

L

2

◆�

Monday, April 29, 13

Page 40: Quantum criticality in condensed matter

A “horizon”, similar to the surface of a black hole !

Gauge-gravity duality at non-zero temperatures

r

SE =

Zd

4x

p�g

1

22

✓R+

6

L

2

◆�

A 2+1 dimensional

system at T > 0with

couplings at its quantum critical point

Monday, April 29, 13

Page 41: Quantum criticality in condensed matter

The temperature and entropy of the

horizon equal those of the quantum

critical point

Gauge-gravity duality at non-zero temperatures

r A 2+1 dimensional

system at T > 0with

couplings at its quantum critical point

Monday, April 29, 13

Page 42: Quantum criticality in condensed matter

Quasi-normal modes of quantum criticality = waves

falling into black hole

Gauge-gravity duality at non-zero temperatures

The temperature and entropy of the

horizon equal those of the quantum

critical point

r A 2+1 dimensional

system at T > 0with

couplings at its quantum critical point

Monday, April 29, 13

Page 43: Quantum criticality in condensed matter

Characteristic damping timeof quasi-normal modes:

(kB/~)⇥ Hawking temperature

Gauge-gravity duality at non-zero temperatures

The temperature and entropy of the

horizon equal those of the quantum

critical point

r A 2+1 dimensional

system at T > 0with

couplings at its quantum critical point

Monday, April 29, 13

Page 44: Quantum criticality in condensed matter

AdS4 theory of quantum criticalityMost general e↵ective holographic theory for lin-ear charge transport with 4 spatial derivatives:

Sbulk =1

g

2M

Zd

4x

pg

1

4FabF

ab + �L

2CabcdF

abF

cd

+

Zd

4x

pg

� 1

22

✓R+

6

L

2

◆�,

This action is characterized by 3 dimensionless pa-rameters, which can be linked to data of the CFT(OPE coe�cients): 2-point correlators of the con-served current Jµ and the stress energy tensor Tµ⌫ ,and a 3-point T , J , J correlator.

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

D. Chowdhury, S. Raju, S. Sachdev, A. Singh, and P. Strack, Phys. Rev. B 87, 085138 (2013)

Monday, April 29, 13

Page 45: Quantum criticality in condensed matter

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)

Monday, April 29, 13

Page 46: Quantum criticality in condensed matter

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)

�(!)

�(1)

AdS4 theory of quantum criticality

• Stability constraints on the e↵ective theory (|�| < 1/12)allow only a limited !-dependence in the conductiv-

ity. This contrasts with the Boltzmann theory in which

�(!)/�1 becomes very large in the regime of its validity.

Monday, April 29, 13

Page 47: Quantum criticality in condensed matter

AdS4 theory of quantum criticality

Universal Scaling of the Conductivity at the Superfluid-Insulator Phase Transition

Jurij Smakov and Erik SørensenDepartment of Physics and Astronomy, McMaster University, Hamilton, Ontario L8S 4M1, Canada

(Received 30 May 2005; published 27 October 2005)

The scaling of the conductivity at the superfluid-insulator quantum phase transition in two dimensionsis studied by numerical simulations of the Bose-Hubbard model. In contrast to previous studies, we focuson properties of this model in the experimentally relevant thermodynamic limit at finite temperature T. Wefind clear evidence for deviations from !k scaling of the conductivity towards !k=T scaling at lowMatsubara frequencies !k. By careful analytic continuation using Pade approximants we show that thisbehavior carries over to the real frequency axis where the conductivity scales with !=T at smallfrequencies and low temperatures. We estimate the universal dc conductivity to be !! " 0:45#5$Q2=h,distinct from previous estimates in the T " 0, !=T % 1 limit.

DOI: 10.1103/PhysRevLett.95.180603 PACS numbers: 05.60.Gg, 02.70.Ss, 05.70.Jk

The nontrivial properties of materials in the vicinity ofquantum phase transitions [1] (QPTs) are an object ofintense theoretical [1–3] and experimental studies. Theeffect of quantum fluctuations driving the QPTs is espe-cially pronounced in low-dimensional systems, such ashigh-temperature superconductors and two-dimensional(2D) electron gases, exhibiting the quantum Hall effect.Particularly valuable are theoretical predictions of thebehavior of the dynamical response functions, such as theoptical conductivity and the dynamic structure factor, sincethey allow for direct comparison of the theoretical resultswith experimental data. It was pointed out by Damle andSachdev [2] that at the quantum-critical coupling thescaled dynamic conductivity T#2&d$=z!#!; T$ at low fre-quencies and temperatures is a function of the singlevariable @!=kBT:

!#!=T; T ! 0$ " #kBT=@c$#d&2$=z!Q!#@!=kBT$: (1)

Here !Q " Q2=h is the conductivity ‘‘quantum’’ (Q " 2efor the models we consider), !#x ' @!=kBT$ is a universaldimensionless scaling function, c a nonuniversal constant,and z the dynamical critical exponent. For d " 2 the ex-ponent vanishes, leading to a purely universal conductivity[4], depending only on frequency !, measured against acharacteristic time @" set by finite temperature T as@!=kBT. Once @!=kBT % 1, for fixed T, the system nolonger ‘‘feels’’ the effect of finite temperature and it isnatural to expect that at such high ! a crossover to atemperature-independent regime will take place [3], sothat !#!; T$ ( !#!$ with !#!$ decaying at high frequen-cies as 1=!2 [2]. Deviations from scaling of ! with !therefore signal that temperature effects have become im-portant. Note that the predicted universal behavior occursfor fixed !=T as T ! 0. The physical mechanisms oftransport are predicted [2] to be quite distinct in the differ-ent regimes determined by the value of the scaling variablex: hydrodynamic, collision dominated for x) 1, and col-lisionless, phase coherent for x% 1 with ! " !#1$

largely independent of x in d " 2 and ! independent ofT [2,5].

Intriguingly, early numerical studies [6–9] of QPTs inmodel systems have failed to observe scaling with@!=kBT. The results of the experiments seeking to verifythe scaling hypothesis are ambiguous as well. Some ofthem, performed at the 2D quantum Hall transitions [10]and 3D metal-insulator transitions [11], appear to supportit. Others either note the absence of scaling [12] or suggesta different scaling form [13]. While the discrepancy be-tween theory and experiment may be attributed to theunsuitable choice of the measurement regime [2], typicallyleading to @!=kBT % 1, there is no good reason why thepredicted scaling would not be observable in numericalsimulations if careful extrapolations first to L! 1 andthen T ! 0 for fixed !=T are performed.

Our primary goal is to resolve this controversy by per-forming precise numerical simulations of the frequency-dependent conductivity at finite temperatures in the vicin-ity of the 2D QPT, exploiting recent algorithmic advancesto access larger system sizes and a wider temperaturerange. After the extrapolation of the results to the thermo-dynamic and T " 0 limits and careful analytic continu-ation, we are able to demonstrate how the predicteduniversal behavior of the conductivity may indeed berevealed.

We consider the 2D Bose-Hubbard (BH) model with theHamiltonian H BH "H 0 *H 1, where the first termdescribes the noninteracting soft core bosons hopping viathe nearest-neighbor links of a 2D square lattice, and thesecond one includes the Hubbard-like on-site interactions:

H 0 " &tX

r;##byr br*# * byr*#br$ &$

Xrnr; (2)

H 1 "U2

Xrnr#nr & 1$: (3)

Here # " x; y, nr " byr br is the particle number operatoron site r, and byr ; br are the boson creation and annihilation

PRL 95, 180603 (2005) P H Y S I C A L R E V I E W L E T T E R S week ending28 OCTOBER 2005

0031-9007=05=95(18)=180603(4)$23.00 180603-1 2005 The American Physical Society

The SSE data also display a high narrow peak at very lowfrequencies, whose position and shape are unstable withrespect to the choice of the initial image and MaxEntparameters. This is clearly an artifact of the method; how-ever, its presence is indicative of the tendency to accumu-late the weight at very low frequencies, in qualitativeagreement with H V result. The subsequent falloff in theconductivity at high frequencies is physically consistent,but its functional form depends on the Pade approximantused. For !=!c * 1=2, we expect the analytic continu-ation of the data for H V to become sensitive to the order ofthe approximant used and we therefore indicate the resultsin this regime by dotted lines only. We note that results atall temperatures yield the same dc conductivity !? !0:45"5#!Q, theoretically predicted [4] to be universal.Because of the very different scaling procedure this resultdiffers from previous numerical result !? ! 0:285"20#!Qon the same model [6] in the T ! 0 limit. It also differssignificantly from a theoretical estimate [2], !$ !1:037!Q, valid to leading order in " ! 3% d. Remark-ably, our result for the dc conductivity is very close to theone obtained in Ref. [8] for the phase transition in thedisordered Bose-Hubbard model. Experimental results in-dicate a value close to unity [26]; however, it was previ-ously observed [7] that long-range Coulomb interactions,impossible to include in the present study, tend to increase! considerably. The same data are shown versus !=T inFig. 4(b). Notably, when using this parametrization !ccancels out and all our data follow the same functionalform. The scaling with !=T at low frequencies is nowimmediately apparent, with a surprisingly wide low !=Tpeak. The width of this peak is consistent with the data inFig. 3(d). Furthermore, on the same !=T scale the con-tinuous time SSE data for H BH and the results for H Vqualitatively agree.

In summary, we have demonstrated that by doing a verycareful data analysis it is possible to observe the theoreti-cally predicted universal !=T scaling at the 2D superfluid-

insulator transition. We have also estimated the universaldc conductivity at this transition and found that it differssignificantly from existing numerical and theoreticalestimates.

We thank S. Sachdev, S. Girvin, and A. P. Young forvaluable comments and critical remarks. Financial sup-port from SHARCNET, NSERC, and CFI is gratefullyacknowledged. All calculations were done at theSHARCNET facility at McMaster University.

[1] J. Hertz, Phys. Rev. B 14, 1165 (1976); A. J. Millis, Phys.Rev. B 48, 7183 (1993); S. Sachdev, Quantum PhaseTransitions (Cambridge University Press, Cambridge,England, 2001); M. Vojta, Rep. Prog. Phys. 66, 2069(2003).

[2] K. Damle and S. Sachdev, Phys. Rev. B 56, 8714 (1997).[3] S. L. Sondhi et al., Rev. Mod. Phys. 69, 315 (1997).[4] M. P. A. Fisher, G. Grinstein, and S. M. Girvin, Phys. Rev.

Lett. 64, 587 (1990).[5] S. Sachdev, Phys. Rev. B 55, 142 (1997).[6] M.-C. Cha et al., Phys. Rev. B 44, 6883 (1991).[7] E. S. Sørensen et al., Phys. Rev. Lett. 69, 828 (1992);

M. Wallin et al., Phys. Rev. B 49, 12 115 (1994).[8] G. G. Batrouni et al., Phys. Rev. B 48, 9628 (1993).[9] R. T. Scalettar, N. Trivedi, and C. Huscroft, Phys. Rev. B

59, 4364 (1999).[10] L. W. Engel et al., Phys. Rev. Lett. 71, 2638 (1993).[11] H.-K. Lee et al., Phys. Rev. Lett. 80, 4261 (1998); Science

287, 633 (2000).[12] N. Q. Balaban, U. Meirav, and I. Bar-Joseph, Phys. Rev.

Lett. 81, 4967 (1998).[13] Y. S. Lee et al., Phys. Rev. B 66, 041104(R) (2002).[14] M. P. A. Fisher et al., Phys. Rev. B 40, 546 (1989).[15] A. W. Sandvik, J. Phys. A 25, 3667 (1992).[16] O. F. Syljuasen and A. W. Sandvik, Phys. Rev. E 66,

046701 (2002); O. F. Syljuasen, ibid. 67, 046701 (2003).[17] M. P. A. Fisher and D. H. Lee, Phys. Rev. B 39, 2756

(1989).[18] F. Alet and E. S. Sørensen, Phys. Rev. E 67, 015701(R)

(2003); 68, 026702 (2003).[19] W. Krauth and N. Trivedi, Europhys. Lett. 14, 627 (1991).[20] J. K. Freericks and H. Monien, Phys. Rev. B 53, 2691

(1996); N. Elstner and H. Monien, ibid. 59, 12 184 (1999).[21] G. D. Mahan, Many-Particle Physics (Plenum, New York,

1990), 2nd ed..[22] D. J. Scalapino, S. R. White, and S. Zhang, Phys. Rev.

Lett. 68, 2830 (1992); Phys. Rev. B 47, 7995 (1993).[23] R. Fazio and D. Zappala, Phys. Rev. B 53, R8883 (1996).[24] M. Jarrell and J. E. Gubernatis, Phys. Rep. 269, 133

(1996).[25] W. H. Press et al., Numerical Recipes in C (Cambridge

University Press, Cambridge, England, 1992).[26] See Y. Liu et al., Physica (Amsterdam) 83D, 163 (1995),

and references therein.[27] Finite Size Scaling and Numerical Simulations of Statis-

tical Systems, edited by V. Privman (World Scientific,Singapore, 1990).

0

0.2

0.4

0.6

0.8

σ(’ω

)

0 0.5 1 1.5ω/ωc

Lτ=32Lτ=40Lτ=48Lτ=64SSE

0

0.2

0.4

0.6

0.8

σ(’ω

)T/

0 25 50 75ω/T

Lτ=32Lτ=40Lτ=48Lτ=64SSE

(a) (b)

FIG. 4. The real part of the conductivity !0 at the criticalcoupling in units of !Q. The data marked L#, plotted vs!=!c, were obtained using H V , combined with the analyticcontinuation of $"!=!c# as explained in the text. Results for!=!c * 1=2 are denoted by dotted lines. The data marked SSE,plotted vs !=10, were obtained by direct SSE simulations ofH BH with L ! 20, % ! 10 and subsequent maximum entropyanalysis (a). Results as a function of !=T (b).

PRL 95, 180603 (2005) P H Y S I C A L R E V I E W L E T T E R S week ending28 OCTOBER 2005

180603-4

!!i!k"=!2"!Q" #h$kxi$!xx!i!k"

!k% #!i!k"

!k: (7)

Here h$kxi is the kinetic energy per link and #!i!k" is thefrequency-dependent stiffness. To measure !xx!i!k" wenote that !xx!$" may be expressed in terms of the correla-tion functions !%&

xx !r; $" # hK%x !r; $"K&

x !0; 0"i of operatorsK&x !r; $" # tbyr&x!$"br!$" and K$x !r; $" # tbyr !$"br&x!$",which may be estimated efficiently in SSE [15].Remarkably, it is possible to analytically perform theFourier transform with respect to $ yielding

!%&xx !r; !k" #

!1

'

Xn$2

m#0

"amn!!k"N!&;%;m""; (8)

where N!&;%;m" is the number of times the operatorsK%!r" and K&!0" appear in the SSE operator sequenceseparated by m operator positions, and n is the expansionorder. The coefficients "amn!!k" are given by the degeneratehypergeometric (Kummer) function: "amn!!k"#1 F1!m&1; n;$i'!k". This expression and (8) allow us to directlyevaluate !xx!r; !k" as a function of Matsubara frequencies,eliminating any errors associated with the discretization ofthe imaginary time interval. Analogously, in the link-current representation #!i!k" can be calculated [7], andthe conductivity can be obtained from Eq. (7).

In Fig. 3 we show results for !!i!k" versus !k obtainedusing the geometrical worm algorithm on H V at Kc[Fig. 3(a)] and by SSE simulations at tc;(c of H BH[Fig. 3(c)]. In both cases the results have been extrapolatedto the thermodynamic limit L! 1 at fixed '. As evidentfrom Fig. 3(a), the results deviate from scaling with !k atsmall !k and more significantly so at higher temperatures(small L$). These deviations are also visible in the con-tinuous time SSE data in Fig. 3(c), demonstrating that theycannot be attributed to time discretization errors. Similardeviations have been noted previously [6,7] but were notanalyzed at fixed '. Since the deviations persist in the L!1 limit at fixed ', they may only be interpreted as finite Teffects. Expecting a crossover to!k=T scaling at small!k,we plot our results versus !k=T in Fig. 3(d). For L$ ' 32,!!!1=T" is already independent of T (L$). In fact, asshown in Fig. 3(d), for !1...5, !!!k=T; T" can unambigu-ously be extrapolated to a finite !!!k=T; T ! 0" ( #!x"limit. This fact is a clear indication that !k=T scalingindeed occurs as T ! 0. Tentatively, for increasing!k=T, !!!k=T; T ! 0" appears to reach a constant valueof roughly 0:33!Q (#!1" in excellent agreement withtheoretical estimates [2,23]. We note that deviations from!k scaling appear to be largely absent in simulations ofH BH with disorder [7,8]. However, at this QCP the dy-namical critical exponent is different (z # 2). As is evidentfrom the size of the error bars in Fig. 3, simulations of H Vare much more efficient than the SSE simulations directlyon H BH. In the following analytic continuation we there-fore use the SSE data mostly as a consistency check.

Our results on the imaginary frequency axis are limitedby the lowest Matsubara frequency, !1 # 2"[email protected], the information about the behavior of !0!!" %Re!!!" at low! is embedded in values of the CCCF at allMatsubara frequencies, allowing us to determine it. Inorder to study the !=T scaling predicted for the hydro-dynamic collision-dominated regime [2] @!=kBT ) 1, wehave attempted analytic continuations of #!i!k" to obtain!0!!" at real frequencies. SSE results for H BH wereanalytically continued using the Bryan maximum entropy(ME) method [24] with flat initial image. For the resultsobtained for the link-current model H V we use a methodthat should be most sensitive to low frequencies !=!c < 1or #!x) 1". We fit the extrapolated low frequency part(first 10–15 Matsubara frequencies) of #!i!k" to a 6th-order polynomial. The resulting 6 coefficients are thenused to obtain a !3; 3" Pade approximant using standardtechniques [25]. This approximant is then used for theanalytic continuation of # by i!k ! !& i). Resultingreal frequency conductivities !0!!" are displayed inFig. 4(a) versus !=!c. The typical SSE data are plottedversus !=10 and are only shown for L # 20, ' # 10.

The results for H V show a broadened peak as !! 0,due to inelastic scattering, followed by a second peaknicely consistent in height and width with the SSE data.

0.2

0.3

0.4

σ(iω

k)

0 0.1 0.2 0.3 0.4 0.5ωk/(2πωc)

Lτ=16

Lτ=24

Lτ=32

Lτ=40

Lτ=48

Lτ=64

0 25 50 75 100ωk/T

Lτ=16

Lτ=24

Lτ=32

Lτ=40

Lτ=48

T=0 Ext.

0.1

0.2

0.3

0.4σ(

iωk)

0 10 20 30ωk

β=2β=4β=6β=8β=10

0.006

0.007

0.008

0.009

0.010

ρ(ω

1)

100 200 300L

Lτ=64

Fit

(a)(b)

(c) (d)

FIG. 3 (color online). The conductivity !!!k" in units of !Qvs Matsubara frequency !k=!c as obtained from H V (a). Allresults have been extrapolated to the thermodynamic limit L!1 using the scaling form f!L" # a& b exp!$L=*"=

####Lp

[27] bycalculating #!!k" at fixed L$ using 9 lattice sizes from L #L$ . . . 4L$ as shown in (b). !!!k" in units of !Q vs Matsubarafrequency !k as obtained from SSE calculations of H BH, withsome typical error bars shown. All results have been extrapolatedto the thermodynamic limit by calculating !xx!!k" for fixed 'using 5 lattice sizes L # 12; . . . ; 30 (c). Scaling plot of theconductivity data from (a) vs !k=T. ! denotes extrapola-tions to T ! 0 (L$ ! 1) at fixed !k=T using: f!L$" # c&d exp!$L$=*$"=

######L$p

[27] (d).

PRL 95, 180603 (2005) P H Y S I C A L R E V I E W L E T T E R S week ending28 OCTOBER 2005

180603-3

Monday, April 29, 13

Page 48: Quantum criticality in condensed matter

AdS4 theory of quantum criticality

Universal Scaling of the Conductivity at the Superfluid-Insulator Phase Transition

Jurij Smakov and Erik SørensenDepartment of Physics and Astronomy, McMaster University, Hamilton, Ontario L8S 4M1, Canada

(Received 30 May 2005; published 27 October 2005)

The scaling of the conductivity at the superfluid-insulator quantum phase transition in two dimensionsis studied by numerical simulations of the Bose-Hubbard model. In contrast to previous studies, we focuson properties of this model in the experimentally relevant thermodynamic limit at finite temperature T. Wefind clear evidence for deviations from !k scaling of the conductivity towards !k=T scaling at lowMatsubara frequencies !k. By careful analytic continuation using Pade approximants we show that thisbehavior carries over to the real frequency axis where the conductivity scales with !=T at smallfrequencies and low temperatures. We estimate the universal dc conductivity to be !! " 0:45#5$Q2=h,distinct from previous estimates in the T " 0, !=T % 1 limit.

DOI: 10.1103/PhysRevLett.95.180603 PACS numbers: 05.60.Gg, 02.70.Ss, 05.70.Jk

The nontrivial properties of materials in the vicinity ofquantum phase transitions [1] (QPTs) are an object ofintense theoretical [1–3] and experimental studies. Theeffect of quantum fluctuations driving the QPTs is espe-cially pronounced in low-dimensional systems, such ashigh-temperature superconductors and two-dimensional(2D) electron gases, exhibiting the quantum Hall effect.Particularly valuable are theoretical predictions of thebehavior of the dynamical response functions, such as theoptical conductivity and the dynamic structure factor, sincethey allow for direct comparison of the theoretical resultswith experimental data. It was pointed out by Damle andSachdev [2] that at the quantum-critical coupling thescaled dynamic conductivity T#2&d$=z!#!; T$ at low fre-quencies and temperatures is a function of the singlevariable @!=kBT:

!#!=T; T ! 0$ " #kBT=@c$#d&2$=z!Q!#@!=kBT$: (1)

Here !Q " Q2=h is the conductivity ‘‘quantum’’ (Q " 2efor the models we consider), !#x ' @!=kBT$ is a universaldimensionless scaling function, c a nonuniversal constant,and z the dynamical critical exponent. For d " 2 the ex-ponent vanishes, leading to a purely universal conductivity[4], depending only on frequency !, measured against acharacteristic time @" set by finite temperature T as@!=kBT. Once @!=kBT % 1, for fixed T, the system nolonger ‘‘feels’’ the effect of finite temperature and it isnatural to expect that at such high ! a crossover to atemperature-independent regime will take place [3], sothat !#!; T$ ( !#!$ with !#!$ decaying at high frequen-cies as 1=!2 [2]. Deviations from scaling of ! with !therefore signal that temperature effects have become im-portant. Note that the predicted universal behavior occursfor fixed !=T as T ! 0. The physical mechanisms oftransport are predicted [2] to be quite distinct in the differ-ent regimes determined by the value of the scaling variablex: hydrodynamic, collision dominated for x) 1, and col-lisionless, phase coherent for x% 1 with ! " !#1$

largely independent of x in d " 2 and ! independent ofT [2,5].

Intriguingly, early numerical studies [6–9] of QPTs inmodel systems have failed to observe scaling with@!=kBT. The results of the experiments seeking to verifythe scaling hypothesis are ambiguous as well. Some ofthem, performed at the 2D quantum Hall transitions [10]and 3D metal-insulator transitions [11], appear to supportit. Others either note the absence of scaling [12] or suggesta different scaling form [13]. While the discrepancy be-tween theory and experiment may be attributed to theunsuitable choice of the measurement regime [2], typicallyleading to @!=kBT % 1, there is no good reason why thepredicted scaling would not be observable in numericalsimulations if careful extrapolations first to L! 1 andthen T ! 0 for fixed !=T are performed.

Our primary goal is to resolve this controversy by per-forming precise numerical simulations of the frequency-dependent conductivity at finite temperatures in the vicin-ity of the 2D QPT, exploiting recent algorithmic advancesto access larger system sizes and a wider temperaturerange. After the extrapolation of the results to the thermo-dynamic and T " 0 limits and careful analytic continu-ation, we are able to demonstrate how the predicteduniversal behavior of the conductivity may indeed berevealed.

We consider the 2D Bose-Hubbard (BH) model with theHamiltonian H BH "H 0 *H 1, where the first termdescribes the noninteracting soft core bosons hopping viathe nearest-neighbor links of a 2D square lattice, and thesecond one includes the Hubbard-like on-site interactions:

H 0 " &tX

r;##byr br*# * byr*#br$ &$

Xrnr; (2)

H 1 "U2

Xrnr#nr & 1$: (3)

Here # " x; y, nr " byr br is the particle number operatoron site r, and byr ; br are the boson creation and annihilation

PRL 95, 180603 (2005) P H Y S I C A L R E V I E W L E T T E R S week ending28 OCTOBER 2005

0031-9007=05=95(18)=180603(4)$23.00 180603-1 2005 The American Physical Society

QMC yields �(0)/�1 ⇡ 1.36

Holography yields �(0)/�1 = 1 + 4� with |�| 1/12.

Maximum possible holographic value �(0)/�1 = 1.33

W. Witzack-Krempa and S. Sachdev, arXiv:1302.0847Monday, April 29, 13

Page 49: Quantum criticality in condensed matter

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

Monday, April 29, 13

Page 50: Quantum criticality in condensed matter

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

Monday, April 29, 13

Page 51: Quantum criticality in condensed matter

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

Monday, April 29, 13

Page 52: Quantum criticality in condensed matter

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

Monday, April 29, 13

Page 53: Quantum criticality in condensed matter

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

Monday, April 29, 13

Page 54: Quantum criticality in condensed matter

1. Superfluid-insulator transition of ultracold atoms in optical lattices:

Conformal field theories and gauge-gravity duality

2. Metals with antiferromagnetism, and high temperature superconductivity The pnictides and the cuprates

Outline

Monday, April 29, 13

Page 55: Quantum criticality in condensed matter

1. Superfluid-insulator transition of ultracold atoms in optical lattices:

Conformal field theories and gauge-gravity duality

2. Metals with antiferromagnetism, and high temperature superconductivity The pnictides and the cuprates

Outline

Monday, April 29, 13

Page 56: Quantum criticality in condensed matter

Ishida, Nakai, and HosonoarXiv:0906.2045v1

Iron pnictides: a new class of high temperature superconductors

Monday, April 29, 13

Page 57: Quantum criticality in condensed matter

TSDW Tc

T0

2.0

0

α"

1.0 SDW

Superconductivity

BaFe2(As1-xPx)2

AF

Resistivity⇠ ⇢0 +AT↵

S. Kasahara, T. Shibauchi, K. Hashimoto, K. Ikada, S. Tonegawa, R. Okazaki, H. Shishido, H. Ikeda, H. Takeya, K. Hirata, T. Terashima, and Y. Matsuda,

Physical Review B 81, 184519 (2010)Monday, April 29, 13

Page 58: Quantum criticality in condensed matter

TSDW Tc

T0

2.0

0

α"

1.0 SDW

Superconductivity

BaFe2(As1-xPx)2

AF

Resistivity⇠ ⇢0 +AT↵

S. Kasahara, T. Shibauchi, K. Hashimoto, K. Ikada, S. Tonegawa, R. Okazaki, H. Shishido, H. Ikeda, H. Takeya, K. Hirata, T. Terashima, and Y. Matsuda,

Physical Review B 81, 184519 (2010)

Short-range entanglement in state with Neel (AF) order

Monday, April 29, 13

Page 59: Quantum criticality in condensed matter

TSDW Tc

T0

2.0

0

α"

1.0 SDW

Superconductivity

BaFe2(As1-xPx)2

AF

Resistivity⇠ ⇢0 +AT↵

S. Kasahara, T. Shibauchi, K. Hashimoto, K. Ikada, S. Tonegawa, R. Okazaki, H. Shishido, H. Ikeda, H. Takeya, K. Hirata, T. Terashima, and Y. Matsuda,

Physical Review B 81, 184519 (2010)

SuperconductorBose condensate of pairs of electrons

Short-range entanglementMonday, April 29, 13

Page 60: Quantum criticality in condensed matter

TSDW Tc

T0

2.0

0

α"

1.0 SDW

Superconductivity

BaFe2(As1-xPx)2

AF

Resistivity⇠ ⇢0 +AT↵

S. Kasahara, T. Shibauchi, K. Hashimoto, K. Ikada, S. Tonegawa, R. Okazaki, H. Shishido, H. Ikeda, H. Takeya, K. Hirata, T. Terashima, and Y. Matsuda,

Physical Review B 81, 184519 (2010)

Ordinary metal(Fermi liquid)

Monday, April 29, 13

Page 61: Quantum criticality in condensed matter

TSDW Tc

T0

2.0

0

α"

1.0 SDW

Superconductivity

BaFe2(As1-xPx)2

AF

Resistivity⇠ ⇢0 +AT↵

S. Kasahara, T. Shibauchi, K. Hashimoto, K. Ikada, S. Tonegawa, R. Okazaki, H. Shishido, H. Ikeda, H. Takeya, K. Hirata, T. Terashima, and Y. Matsuda,

Physical Review B 81, 184519 (2010)

StrangeMetal

no quasiparticles,Landau-Boltzmann theory

does not apply

Monday, April 29, 13

Page 62: Quantum criticality in condensed matter

YBa2Cu3O6+x

High temperature superconductors

Monday, April 29, 13

Page 63: Quantum criticality in condensed matter

� �

StrangeMetal

Monday, April 29, 13

Page 64: Quantum criticality in condensed matter

� �

StrangeMetal

Pseudogap

Monday, April 29, 13

Page 65: Quantum criticality in condensed matter

Fermi surface+antiferromagnetism

The electron spin polarization obeys�

⌃S(r, �)⇥

= ⌃⇥(r, �)eiK·r

where K is the ordering wavevector.

+

Metal with “large” Fermi surface

Monday, April 29, 13

Page 66: Quantum criticality in condensed matter

Metal with “large” Fermi surface

Fermi surface+antiferromagnetism

Monday, April 29, 13

Page 67: Quantum criticality in condensed matter

Fermi surfaces translated by K = (�,�).

Fermi surface+antiferromagnetism

Monday, April 29, 13

Page 68: Quantum criticality in condensed matter

“Hot” spots

Fermi surface+antiferromagnetism

Monday, April 29, 13

Page 69: Quantum criticality in condensed matter

Fermi surface+antiferromagnetism

Electron and hole pockets in

antiferromagnetic phase

with antiferromagnetic order parameter h~'i 6= 0

Monday, April 29, 13

Page 70: Quantum criticality in condensed matter

r

Metal with “large” Fermi surface

h~'i = 0

Metal with electron and hole pockets

Increasing SDW order

h~'i 6= 0

Fermi surface+antiferromagnetism

Monday, April 29, 13

Page 71: Quantum criticality in condensed matter

Unconventional pairing at and near hot spots

V. J. Emery, J. Phys. (Paris) Colloq. 44, C3-977 (1983)D.J. Scalapino, E. Loh, and J.E. Hirsch, Phys. Rev. B 34, 8190 (1986)K. Miyake, S. Schmitt-Rink, and C. M. Varma, Phys. Rev. B 34, 6554 (1986)S. Raghu, S.A. Kivelson, and D.J. Scalapino, Phys. Rev. B 81, 224505 (2010)

Monday, April 29, 13

Page 72: Quantum criticality in condensed matter

Unconventional pairing at and near hot spots

Dc†k↵c

†�k�

E= "

↵�

S

(cos kx

� cos ky

)

�S

��S

V. J. Emery, J. Phys. (Paris) Colloq. 44, C3-977 (1983)D.J. Scalapino, E. Loh, and J.E. Hirsch, Phys. Rev. B 34, 8190 (1986)K. Miyake, S. Schmitt-Rink, and C. M. Varma, Phys. Rev. B 34, 6554 (1986)S. Raghu, S.A. Kivelson, and D.J. Scalapino, Phys. Rev. B 81, 224505 (2010)

Monday, April 29, 13

Page 73: Quantum criticality in condensed matter

Sign-problem-free Quantum Monte Carlo for antiferromagnetism in metals

−2 −1 0 1 2 3−2

0

2

4

6

8

10 x 10−4

r

P ±(xmax

)L = 10

L = 14

L = 12

rc

P+

_

_P_

|

s/d pairing amplitudes P+/P�as a function of the tuning parameter r

E. Berg, M. Metlitski, and S. Sachdev, Science 338, 1606 (2012).

Monday, April 29, 13

Page 74: Quantum criticality in condensed matter

FluctuatingFermi

pocketsLargeFermi

surface

StrangeMetal

Spin density wave (SDW)

Underlying SDW ordering quantum critical pointin metal at x = xm

Increasing SDW order

T*QuantumCritical

Fermi surface+antiferromagnetism

Monday, April 29, 13

Page 75: Quantum criticality in condensed matter

LargeFermi

surface

StrangeMetal

Spin density wave (SDW)

d-wavesuperconductor

Small Fermipockets with

pairing fluctuationsFluctuating, paired Fermi

pockets

QCP for the onset of SDW order is actually within a superconductor

Fermi surface+antiferromagnetism

QuantumCritical

Monday, April 29, 13

Page 76: Quantum criticality in condensed matter

TSDW Tc

T0

2.0

0

α"

1.0 SDW

Superconductivity

BaFe2(As1-xPx)2

AF

Resistivity⇠ ⇢0 +AT↵

S. Kasahara, T. Shibauchi, K. Hashimoto, K. Ikada, S. Tonegawa, R. Okazaki, H. Shishido, H. Ikeda, H. Takeya, K. Hirata, T. Terashima, and Y. Matsuda,

Physical Review B 81, 184519 (2010)

StrangeMetal

no quasiparticles,Landau-Boltzmann theory

does not apply

Monday, April 29, 13

Page 77: Quantum criticality in condensed matter

� �

StrangeMetal

Pseudogap

What about the pseudogap ?

Monday, April 29, 13

Page 78: Quantum criticality in condensed matter

There is an approximate pseudospin symmetry in metals with antiferromagnetic spin correlations.

The pseudospin partner of d-wave superconductivity is an incommensurate d-wave bond order

These orders form a pseudospin doublet, which is responsible for the “pseudogap” phase.

M. A. Metlitski and S. Sachdev, Phys. Rev. B 85, 075127 (2010)T. Holder and W. Metzner, Phys. Rev. B 85, 165130 (2012)

C. Husemann and W. Metzner, Phys. Rev. B 86, 085113 (2012)K. B. Efetov, H. Meier, and C. Pepin, arXiv:1210.3276.

S. Sachdev and R. La Placa, arXiv:1303.2114

Monday, April 29, 13

Page 79: Quantum criticality in condensed matter

HJ =

X

i<j

Jij ~Si · ~Sj

with

~Si =

12c

†i↵~�↵�ci� is the antiferromagnetic exchange interaction.

Introduce the Nambu spinor

i" =

✓ci"c†i#

◆, i# =

✓ci#�c†i"

Then we can write

HJ =

1

8

X

i<j

Jij

†i↵~�↵� i�

⌘·⇣

†j�~��� i�

which is invariant under independent SU(2) pseudospin transforma-

tions on each site

i↵ ! Ui i↵

This pseudospin (gauge) symmetry is important in classifying spin

liquid ground states of HJ . It is fully broken by the electron hopping

tij but does have remnant consequences in doped spin liquid states.

Pseudospin symmetry of the exchange interaction

I. A✏eck, Z. Zou, T. Hsu, and P. W. Anderson, Phys. Rev. B 38, 745 (1988)E. Dagotto, E. Fradkin, and A. Moreo, Phys. Rev. B 38, 2926 (1988)P. A. Lee, N. Nagaosa, and X.-G. Wen, Rev. Mod. Phys. 78, 17 (2006)

Monday, April 29, 13

Page 80: Quantum criticality in condensed matter

HJ =

X

i<j

Jij ~Si · ~Sj

with

~Si =

12c

†i↵~�↵�ci� is the antiferromagnetic exchange interaction.

Introduce the Nambu spinor

i" =

✓ci"c†i#

◆, i# =

✓ci#�c†i"

Then we can write

HJ =

1

8

X

i<j

Jij

†i↵~�↵� i�

⌘·⇣

†j�~��� i�

which is invariant under independent SU(2) pseudospin transforma-

tions on each site

i↵ ! Ui i↵

This pseudospin (gauge) symmetry is important in classifying spin

liquid ground states of HJ . It is fully broken by the electron hopping

tij but does have remnant consequences in doped spin liquid states.

Pseudospin symmetry of the exchange interaction

I. A✏eck, Z. Zou, T. Hsu, and P. W. Anderson, Phys. Rev. B 38, 745 (1988)E. Dagotto, E. Fradkin, and A. Moreo, Phys. Rev. B 38, 2926 (1988)P. A. Lee, N. Nagaosa, and X.-G. Wen, Rev. Mod. Phys. 78, 17 (2006)

HtJ = �X

i,j

tijc†i↵cj↵ +

X

i<j

Jij ~Si · ~Sj

Monday, April 29, 13

Page 81: Quantum criticality in condensed matter

HJ =

X

i<j

Jij ~Si · ~Sj

with

~Si =

12c

†i↵~�↵�ci� is the antiferromagnetic exchange interaction.

Introduce the Nambu spinor

i" =

✓ci"c†i#

◆, i# =

✓ci#�c†i"

Then we can write

HJ =

1

8

X

i<j

Jij

†i↵~�↵� i�

⌘·⇣

†j�~��� i�

which is invariant under independent SU(2) pseudospin transforma-

tions on each site

i↵ ! Ui i↵

This pseudospin (gauge) symmetry is important in classifying spin

liquid ground states of HJ . It is fully broken by the electron hopping

tij but does have remnant consequences in doped spin liquid states.

Pseudospin symmetry of the exchange interaction

HtJ = �X

i,j

tijc†i↵cj↵ +

X

i<j

Jij ~Si · ~Sj

We will start with the Neel state, and find important

consequences of the pseudospin symmetry in metals

with antiferromagnetic correlations.

M. A. Metlitski and S. Sachdev, Phys. Rev. B 85, 075127 (2010)Monday, April 29, 13

Page 82: Quantum criticality in condensed matter

r

Metal with “large” Fermi surface

h~'i = 0

Metal with electron and hole pockets

Increasing SDW order

h~'i 6= 0

Fermi surface+antiferromagnetism

Monday, April 29, 13

Page 83: Quantum criticality in condensed matter

r

Metal with “large” Fermi surface

h~'i = 0

Metal with electron and hole pockets

Increasing SDW order

h~'i 6= 0

Fermi surface+antiferromagnetism

Focus on this

region

Monday, April 29, 13

Page 84: Quantum criticality in condensed matter

“Hot” spots

Fermi surface+antiferromagnetism

Monday, April 29, 13

Page 85: Quantum criticality in condensed matter

Low energy theory for critical point near hot spots

Fermi surface+antiferromagnetism

Monday, April 29, 13

Page 86: Quantum criticality in condensed matter

Low energy theory for critical point near hot spots

Fermi surface+antiferromagnetism

Monday, April 29, 13

Page 87: Quantum criticality in condensed matter

v1 v2

�2 fermionsoccupied

�1 fermionsoccupied

Theory has fermions 1,2 (with Fermi velocities v1,2)

and boson order parameter ~',interacting with coupling �

kx

ky

Ar. Abanov and A. V. Chubukov,

Phys. Rev. Lett. 84, 5608 (2000).

Monday, April 29, 13

Page 88: Quantum criticality in condensed matter

v1 v2

�2 fermionsoccupied

�1 fermionsoccupied

kx

ky

Ar. Abanov and A. V. Chubukov,

Phys. Rev. Lett. 84, 5608 (2000).

S =

Zd2rd⌧

†1↵ (@⌧ � iv1 ·rr) 1↵ + †

2↵ (@⌧ � iv2 ·rr) 2↵

+1

2(rr ~')

2 +s

2~'2 +

u

4~'4 � �~' ·

⇣ †1↵~�↵� 2� + †

2↵~�↵� 1�

⌘�

Monday, April 29, 13

Page 89: Quantum criticality in condensed matter

v1 v2

�2 fermionsoccupied

�1 fermionsoccupied

kx

ky

S =

Zd2rd⌧

†1↵ (@⌧ � iv1 ·rr) 1↵ + †

2↵ (@⌧ � iv2 ·rr) 2↵

+1

2(rr ~')

2 +s

2~'2 +

u

4~'4 � �~' ·

⇣ †1↵~�↵� 2� + †

2↵~�↵� 1�

⌘�

M. A. Metlitski and S. Sachdev,

Phys. Rev. B 85, 075127 (2010)

This low-energy theory is invariant under

independent SU(2) pseudospin rotations on each

pair of hot-spots: there is a global

SU(2)⇥SU(2)⇥SU(2)⇥SU(2) pseudospin symmetry.

Monday, April 29, 13

Page 90: Quantum criticality in condensed matter

Unconventional pairing at and near hot spots

V. J. Emery, J. Phys. (Paris) Colloq. 44, C3-977 (1983)D.J. Scalapino, E. Loh, and J.E. Hirsch, Phys. Rev. B 34, 8190 (1986)K. Miyake, S. Schmitt-Rink, and C. M. Varma, Phys. Rev. B 34, 6554 (1986)S. Raghu, S.A. Kivelson, and D.J. Scalapino, Phys. Rev. B 81, 224505 (2010)

Monday, April 29, 13

Page 91: Quantum criticality in condensed matter

Unconventional pairing at and near hot spots

Dc†k↵c

†�k�

E= "

↵�

S

(cos kx

� cos ky

)

�S

��S

V. J. Emery, J. Phys. (Paris) Colloq. 44, C3-977 (1983)D.J. Scalapino, E. Loh, and J.E. Hirsch, Phys. Rev. B 34, 8190 (1986)K. Miyake, S. Schmitt-Rink, and C. M. Varma, Phys. Rev. B 34, 6554 (1986)S. Raghu, S.A. Kivelson, and D.J. Scalapino, Phys. Rev. B 81, 224505 (2010)

Monday, April 29, 13

Page 92: Quantum criticality in condensed matter

After pseudospin rotation on

half the hot-spots

Unconventional particle-hole pairing at and near hot spots

Q is ‘2kF ’wavevector

M. A. Metlitski and S. Sachdev,

Phys. Rev. B 85, 075127 (2010)

��Q

�Q

Dc†k�Q/2,↵ck+Q/2,↵

E= �Q(cos k

x

� cos ky

)

Monday, April 29, 13

Page 93: Quantum criticality in condensed matter

Q

Q

�Q

��Q

Dc†k�Q/2,↵ck+Q/2,↵

E= �Q(cos k

x

� cos ky

)

M. A. Metlitski and S. Sachdev,

Phys. Rev. B 85, 075127 (2010)

Incommensurate d-wave bond order

Monday, April 29, 13

Page 94: Quantum criticality in condensed matter

⌦c†r↵cs↵

↵=

X

Q

X

k

eiQ·(r+s)/2e�ik·(r�s)Dc†k�Q/2,↵ck+Q/2,↵

E

where Q extends over Q = (±Q0,±Q0) with Q0 = 2⇡/(7.3) and

Dc†k�Q/2,↵ck+Q/2,↵

E= �Q(cos k

x

� cos ky

)

Note

⌦c†r↵cs↵

↵is non-zero only when r, s are nearest neighbors.

“Bond density” measures amplitude for electrons to be

in spin-singlet valence bond.

-1

+1

M. A. Metlitski and S. Sachdev,

Phys. Rev. B 85, 075127 (2010)

Incommensurate d-wave bond order

Monday, April 29, 13

Page 95: Quantum criticality in condensed matter

A Four Unit Cell PeriodicPattern of Quasi-Particle StatesSurrounding Vortex Cores in

Bi2Sr2CaCu2O8!"J. E. Hoffman,1 E. W. Hudson,1,2* K. M. Lang,1 V. Madhavan,1

H. Eisaki,3† S. Uchida,3 J. C. Davis1,2‡

Scanning tunneling microscopy is used to image the additional quasi-particlestates generated by quantized vortices in the high critical temperature super-conductor Bi2Sr2CaCu2O8!". They exhibit a copper-oxygen bond–oriented“checkerboard” pattern, with four unit cell (4a0) periodicity and a #30 ang-strom decay length. These electronic modulations may be related to the mag-netic field–induced, 8a0 periodic, spin density modulations with decay lengthof #70 angstroms recently discovered in La1.84Sr0.16CuO4. The proposed ex-planation is a spin density wave localized surrounding each vortex core. Generaltheoretical principles predict that, in the cuprates, a localized spin modulationof wavelength $ should be associated with a corresponding electronic modu-lation of wavelength $/2, in good agreement with our observations.

Theory indicates that the electronic structure ofthe cuprates is susceptible to transitions into avariety of ordered states (1–10). Experimentally,antiferromagnetism (AF) and high-temperaturesuperconductivity (HTSC) occupy well-knownregions of the phase diagram, but, outside theseregions, several unidentified ordered states exist.For example, at low hole densities and above thesuperconducting transition temperature, the un-identified “pseudogap” state exhibits gappedelectronic excitations (11). Other unidentifiedordered states, both insulating (12) and conduct-ing (13), exist in magnetic fields sufficient toquench superconductivity. Categorization of thecuprate electronic ordered states and clarifica-tion of their relationship to HTSC are among thekey challenges in condensed matter physicstoday.

Because the suppression of superconductiv-ity inside a vortex core can allow one of thealternative ordered states (1–10) to appear there,the electronic structure of HTSC vortices hasattracted wide attention. Initially, theoretical ef-forts focused on the quantized vortex in a Bard-een-Cooper-Schrieffer (BCS) superconductorwith dx2-y2 symmetry (14–18). These modelsincluded predictions that, because of the gap

nodes, the local density of electronic states(LDOS) inside the core is strongly peaked at theFermi level. This peak, which would appear intunneling studies as a zero bias conductancepeak (ZBCP), should display a four-fold sym-metric “star shape” oriented toward the gapnodes and decaying as a power law withdistance.

Scanning tunneling microscopy (STM) stud-ies of HTSC vortices have revealed a very dif-ferent electronic structure from that predicted bythe pure d-wave BCS models. Vortices inYBa2Cu3O7 (YBCO) lack ZBCPs but exhibitadditional quasi-particle states at %5.5 meV(19), whereas those in Bi2Sr2CaCu2O8!" (Bi-2212) also lack ZBCPs (20). More recently, theadditional quasi-particle states at Bi-2212 vorti-ces were discovered at energies near %7 meV(21). Thus, a common phenomenology for low-energy quasi-particles associated with vortices isbecoming apparent. Its features include (i) theabsence of ZBCPs, (ii) a radius for the actualvortex core (where the coherence peaks areabsent) of #10 Å (21), (iii) low-energy quasi-particle states at %5.5 meV (YBCO) and %7meV (Bi-2212), (iv) a radius of up to #75 Åwithin which these states are detected (21), and(v) the absence of a four-fold symmetric star-shaped LDOS.

Because d-wave BCS models do not explainthis phenomenology, new theoretical approach-es have been developed. Zhang (5) and Arovaset al. (22) first focused attention on magneticphenomena associated with HTSC vortices withproposals that a magnetic field induces antifer-romagnetic order localized by the core. Moregenerally, new theories describe vortex-inducedelectronic and magnetic phenomena when theanticipated effects of strong correlations and

strong antiferromagnetic spin fluctuations areincluded (22–26). Common elements of theirpredictions include the following: (i) the prox-imity of a phase transition into a magnetic or-dered state can be revealed when the supercon-ductivity is weakened by the influence of avortex (22–26), (ii) the resulting magnetic order,either spin (22, 23, 25) or orbital (24, 26), willcoexist with superconductivity in some regionnear the core, and (iii) this localized magneticorder will generate associated spatial modula-tions in the quasi-particle density of states (23–26). Given the relevance of such predictions tothe identification of alternative ordered states,determination of the magnetic and electronicstructure of the HTSC vortex is an urgentpriority.

Information on the magnetic structure ofHTSC vortices has recently become availablefrom neutron scattering and nuclear magneticresonance (NMR) studies. Near optimum dop-ing, some cuprates show strong inelastic neutronscattering (INS) peaks at the four k-space points(1/2 % ", 1/2) and (1/2, 1/2 % "), where " # 1/8and k-space distances are measured in units of2&/a0. This demonstrates the existence, in realspace, of fluctuating magnetization density withspatial periodicity of 8a0 oriented along theCu-O bond directions, in the superconductingphase. The first evidence for field-induced mag-netic order in the cuprates came from INS ex-periments on La1.84Sr0.16CuO4 by Lake et al.(27). When La1.84Sr0.16CuO4 is cooled into thesuperconducting state, the scattering intensity atthese characteristic k-space locations disappearsat energies below #7 meV, opening up a “spingap.” Application of a 7.5 T magnetic fieldbelow 10 K causes the scattering intensity toreappear with strength almost equal to that in thenormal state. These field-induced spin fluctua-tions have a spatial periodicity of 8a0 and wavevector pointing along the Cu-O bond direction.Their magnetic coherence length LM is at least20a0, although the vortex-core diameter is only#5a0. More recently, studies by Khaykovich etal. (28) on a related material, La2CuO4!y, foundfield-induced enhancement of elastic neutronscattering (ENS) intensity at these same incom-mensurate k-space locations, but with LM '100a0. Thus, field-induced static AF order with8a0 periodicity exists in this material. Finally,NMR studies by Mitrovic et al. (29) exploredthe spatial distribution of magnetic fluctuationsnear the core. NMR is used because 1/T1, theinverse spin-lattice relaxation time, is a measureof spin fluctuations, and the Larmor frequencyof the probe nucleus is a measure of their loca-tions relative to the vortex center. In YBCO atB ( 13 T, the 1/T1 of 17O rises rapidly as thecore is approached and then diminishes insidethe core. These experiments are all consistentwith vortex-induced spin fluctuations occurringoutside the core.

Theoretical attention was first focused on theregions outside the core by a phenomenological

1Department of Physics, University of California, Berke-ley, CA 94720–7300, USA. 2Materials Sciences Divi-sion, Lawrence Berkeley National Laboratory. Berke-ley, CA 94720, USA. 3Department of Superconductiv-ity, University of Tokyo, Tokyo 113-8656, Japan.

*Present address: Department of Physics, Massachu-setts Institute of Technology, Cambridge, MA 02139–4307, USA.†Present address: Department of Applied Physics,Stanford University, Stanford, CA 94305–4060, USA.‡To whom correspondence should be addressed. E-mail: [email protected]

R E P O R T S

18 JANUARY 2002 VOL 295 SCIENCE www.sciencemag.org466

on

Mar

ch 1

, 201

3w

ww

.sci

ence

mag

.org

Dow

nloa

ded

from

model that proposed that the circulating super-currents weaken the superconducting order pa-rameter and allow the local appearance of acoexisting spin density wave (SDW) and HTSCphase (23) surrounding the core. In a morerecent model, which is an extension of (5) and(22), the effective mass associated with spinfluctuations results in an AF localization lengththat might be substantially greater than the coreradius (30). An associated appearance of chargedensity wave order was also predicted (31)whose effects on the HTSC quasi-particlesshould be detectable in the regions surroundingthe vortex core (23).

To test these ideas, we apply our recentlydeveloped techniques of low-energy quasi-par-ticle imaging at HTSC vortices (21). We chooseto study Bi-2212, because YBCO and LSCOhave proven nonideal for spectroscopic studiesbecause their cleaved surfaces often exhibitnonsuperconducting spectra. Our “as-grown”Bi-2212 crystals are generated by the floatingzone method, are slightly overdoped with Tc !89 K, and contain 0.5% of Ni impurity atoms.They are cleaved (at the BiO plane) in cryogen-ic ultrahigh vacuum below 30 K and immedi-ately inserted into the STM head. Figure 1Ashows a topographic image of the 560 Å squarearea where all the STM measurements reportedhere were carried out. The atomic resolutionand the supermodulation (with wavelength"26 Å oriented at 45° to the Cu-O bond direc-tions) are evident throughout.

To study effects of the magnetic field B onthe superconducting electronic structure, wefirst acquire zero-field maps of the differentialtunneling conductance (G ! dI/dV) measuredat all locations (x, y) in the field of view (FOV)of Fig. 1A. Because LDOS(E ! eV) # G(V ),where V is the sample bias voltage, this resultsin a two-dimensional map of the local densityof states LDOS(E, x, y, B ! 0). We acquirethese LDOS maps at energies ranging from –12meV to $12 meV in 1-meV increments. The Bfield is then ramped to its target value, and, afterany drift has stabilized, we remeasure the topo-graph with the same resolution. The FOVwhere the high-field LDOS measurements areto be made is then matched to that in Fig. 1Awithin 1 Å ("0.25a0) by comparing character-istic topographic/spectroscopic features. Final-ly, we acquire the high-field LDOS maps,LDOS(E, x, y, B), at the same series of energiesas the zero-field case.

To focus preferentially on B field effects,we define a type of two-dimensional map:

S E1

E2(x, y, B) ! !E1

E2

%LDOS&E, x, y, B'

! LDOS&E, x, y, 0'(dE (1)

which represents the integral of all additionalspectral density induced by the B field be-tween the energies E1 and E2 at each location

(x, y). We use this technique of combinedelectronic background subtraction and energyintegration to enhance the signal-to-noise ra-tio of the vortex-induced states. In Bi-2212,these states are broadly distributed in energyaround )7 meV (21), so S )1

)12(x, y, B) effec-tively maps the additional spectral strengthunder their peaks.

Figure 1B is an image of S112(x, y, 5)

measured in the FOV of Fig. 1A. The loca-tions of seven vortices are evident as thedarker regions of dimension "100 Å. Eachvortex displays a spatial structure in the inte-grated LDOS consisting of a checkerboardpattern oriented along Cu-O bonds. We haveobserved spatial structure with the same pe-riodicity and orientation, in the vortex-in-duced LDOS on multiple samples and atfields ranging from 2 to 7 T. In all 35 vorticesstudied in detail, this spatial and energeticstructure exists, but the checkerboard is moreclearly resolved by the positive-bias peak.

We show the power spectrum from thetwo-dimensional Fourier transform ofS1

12(x, y, 5),PS[S112(x, y, 5)]!{FT%S1

12&x, y, 5)]}2,in Fig. 2A and a labeled schematic of theseresults in Fig. 2B. In these k-space images,the atomic periodicity is detected at the pointslabeled by A, which by definition are at(0,)1) and ()1,0). The harmonics of thesupermodulation are identified by the sym-bols B1 and B2. These features (A, B1, andB2) are observed in the Fourier transforms ofall LDOS maps, independent of magneticfield, and they remain as a small backgroundsignal in PS[S1

12(x, y, 5)] because the zero-field and high-field LDOS images can onlybe matched to within 1 Å before subtraction.Most importantly, PS[S1

12(x, y, 5)] revealsnew peaks at the four k-space points, whichcorrespond to the spatial structure of the vor-tex-induced quasi-particle states. We labeltheir locations C. No similar peaks in thespectral weight exist at these points in thetwo-dimensional Fourier transform of thesezero-field LDOS maps.

To quantify these results, we fit a Lorent-zian to PS[S1

12(x, y, 5)] at each of the fourpoints labeled C in Fig. 2B. We find that theyoccur at k-space radius 0.062 Å*1 with width+ ! 0.011 ) 0.002 Å*1. Figure 2C showsthe value of PS[S1

12(x, y, 5)] measured alongthe dashed line in Fig. 2B. The central peakassociated with long-wavelength structure,the peak associated with the atoms, and thepeak due to the vortex-induced quasi-particlestates are all evident. The vortex-inducedstates identified by this means occur at ()1/4,0) and (0, )1/4) to within the accuracy of themeasurement. Equivalently, the checkerboardpattern evident in the LDOS has spatial peri-odicity 4a0 oriented along the Cu-O bonds.Furthermore, the width + of the Lorentzianyields a spatial correlation length for theseLDOS oscillations of L ! (1/,+) - 30 ) 5

Å (or L - 7.8 ) 1.3a0). This is substantiallygreater than the measured (21) core radius. Itis also evident in Figs. 1B and 2A that theLDOS oscillations have stronger spectralweight in one Cu-O direction than in the

Fig. 1. Topographic and spectroscopic images ofthe same area of a Bi-2212 surface. (A) A topo-graphic image of the 560 Å field of view (FOV ) inwhich the vortex studies were carried out. Thesupermodulation can be seen clearly along withsome effects of electronic inhomogeneity. TheCu–O–Cu bonds are oriented at 45° to the su-permodulation. Atomic resolution is evidentthroughout, and the inset shows a 140 Å squareFOV at .2 magnification to make this easier tosee. The mean Bi-Bi distance apparent here isa0 ! 3.83 Å and is identical to the mean Cu-Cudistance in the CuO plane "5 Å below. (B) A mapof S1

12(x, y, 5) showing the additional LDOS in-duced by the seven vortices. Each vortex is ap-parent as a checkerboard at 45° to the pageorientation. Not all are identical, most likely be-cause of the effects of electronic inhomogeneity.The units of S1

12(x, y, 5) are picoamps because itrepresents /dI/dV!0V. In this energy range, themaximum integrated LDOS at a vortex is "3pA, as compared with the zero field integratedLDOS of "1 pA. The latter is subtracted fromthe former to give a maximum contrast of "2pA. We also note that the integrated differen-tial conductance between 0 and *200 meV is200 pA because all measurements reported inthis paper were obtained at a junction resis-tance of 1 gigaohm set at a bias voltage of–200 mV.

R E P O R T S

www.sciencemag.org SCIENCE VOL 295 18 JANUARY 2002 467

on

Mar

ch 1

, 201

3w

ww

.sci

ence

mag

.org

Dow

nloa

ded

from

model that proposed that the circulating super-currents weaken the superconducting order pa-rameter and allow the local appearance of acoexisting spin density wave (SDW) and HTSCphase (23) surrounding the core. In a morerecent model, which is an extension of (5) and(22), the effective mass associated with spinfluctuations results in an AF localization lengththat might be substantially greater than the coreradius (30). An associated appearance of chargedensity wave order was also predicted (31)whose effects on the HTSC quasi-particlesshould be detectable in the regions surroundingthe vortex core (23).

To test these ideas, we apply our recentlydeveloped techniques of low-energy quasi-par-ticle imaging at HTSC vortices (21). We chooseto study Bi-2212, because YBCO and LSCOhave proven nonideal for spectroscopic studiesbecause their cleaved surfaces often exhibitnonsuperconducting spectra. Our “as-grown”Bi-2212 crystals are generated by the floatingzone method, are slightly overdoped with Tc !89 K, and contain 0.5% of Ni impurity atoms.They are cleaved (at the BiO plane) in cryogen-ic ultrahigh vacuum below 30 K and immedi-ately inserted into the STM head. Figure 1Ashows a topographic image of the 560 Å squarearea where all the STM measurements reportedhere were carried out. The atomic resolutionand the supermodulation (with wavelength"26 Å oriented at 45° to the Cu-O bond direc-tions) are evident throughout.

To study effects of the magnetic field B onthe superconducting electronic structure, wefirst acquire zero-field maps of the differentialtunneling conductance (G ! dI/dV) measuredat all locations (x, y) in the field of view (FOV)of Fig. 1A. Because LDOS(E ! eV) # G(V ),where V is the sample bias voltage, this resultsin a two-dimensional map of the local densityof states LDOS(E, x, y, B ! 0). We acquirethese LDOS maps at energies ranging from –12meV to $12 meV in 1-meV increments. The Bfield is then ramped to its target value, and, afterany drift has stabilized, we remeasure the topo-graph with the same resolution. The FOVwhere the high-field LDOS measurements areto be made is then matched to that in Fig. 1Awithin 1 Å ("0.25a0) by comparing character-istic topographic/spectroscopic features. Final-ly, we acquire the high-field LDOS maps,LDOS(E, x, y, B), at the same series of energiesas the zero-field case.

To focus preferentially on B field effects,we define a type of two-dimensional map:

S E1

E2(x, y, B) ! !E1

E2

%LDOS&E, x, y, B'

! LDOS&E, x, y, 0'(dE (1)

which represents the integral of all additionalspectral density induced by the B field be-tween the energies E1 and E2 at each location

(x, y). We use this technique of combinedelectronic background subtraction and energyintegration to enhance the signal-to-noise ra-tio of the vortex-induced states. In Bi-2212,these states are broadly distributed in energyaround )7 meV (21), so S )1

)12(x, y, B) effec-tively maps the additional spectral strengthunder their peaks.

Figure 1B is an image of S112(x, y, 5)

measured in the FOV of Fig. 1A. The loca-tions of seven vortices are evident as thedarker regions of dimension "100 Å. Eachvortex displays a spatial structure in the inte-grated LDOS consisting of a checkerboardpattern oriented along Cu-O bonds. We haveobserved spatial structure with the same pe-riodicity and orientation, in the vortex-in-duced LDOS on multiple samples and atfields ranging from 2 to 7 T. In all 35 vorticesstudied in detail, this spatial and energeticstructure exists, but the checkerboard is moreclearly resolved by the positive-bias peak.

We show the power spectrum from thetwo-dimensional Fourier transform ofS1

12(x, y, 5),PS[S112(x, y, 5)]!{FT%S1

12&x, y, 5)]}2,in Fig. 2A and a labeled schematic of theseresults in Fig. 2B. In these k-space images,the atomic periodicity is detected at the pointslabeled by A, which by definition are at(0,)1) and ()1,0). The harmonics of thesupermodulation are identified by the sym-bols B1 and B2. These features (A, B1, andB2) are observed in the Fourier transforms ofall LDOS maps, independent of magneticfield, and they remain as a small backgroundsignal in PS[S1

12(x, y, 5)] because the zero-field and high-field LDOS images can onlybe matched to within 1 Å before subtraction.Most importantly, PS[S1

12(x, y, 5)] revealsnew peaks at the four k-space points, whichcorrespond to the spatial structure of the vor-tex-induced quasi-particle states. We labeltheir locations C. No similar peaks in thespectral weight exist at these points in thetwo-dimensional Fourier transform of thesezero-field LDOS maps.

To quantify these results, we fit a Lorent-zian to PS[S1

12(x, y, 5)] at each of the fourpoints labeled C in Fig. 2B. We find that theyoccur at k-space radius 0.062 Å*1 with width+ ! 0.011 ) 0.002 Å*1. Figure 2C showsthe value of PS[S1

12(x, y, 5)] measured alongthe dashed line in Fig. 2B. The central peakassociated with long-wavelength structure,the peak associated with the atoms, and thepeak due to the vortex-induced quasi-particlestates are all evident. The vortex-inducedstates identified by this means occur at ()1/4,0) and (0, )1/4) to within the accuracy of themeasurement. Equivalently, the checkerboardpattern evident in the LDOS has spatial peri-odicity 4a0 oriented along the Cu-O bonds.Furthermore, the width + of the Lorentzianyields a spatial correlation length for theseLDOS oscillations of L ! (1/,+) - 30 ) 5

Å (or L - 7.8 ) 1.3a0). This is substantiallygreater than the measured (21) core radius. Itis also evident in Figs. 1B and 2A that theLDOS oscillations have stronger spectralweight in one Cu-O direction than in the

Fig. 1. Topographic and spectroscopic images ofthe same area of a Bi-2212 surface. (A) A topo-graphic image of the 560 Å field of view (FOV ) inwhich the vortex studies were carried out. Thesupermodulation can be seen clearly along withsome effects of electronic inhomogeneity. TheCu–O–Cu bonds are oriented at 45° to the su-permodulation. Atomic resolution is evidentthroughout, and the inset shows a 140 Å squareFOV at .2 magnification to make this easier tosee. The mean Bi-Bi distance apparent here isa0 ! 3.83 Å and is identical to the mean Cu-Cudistance in the CuO plane "5 Å below. (B) A mapof S1

12(x, y, 5) showing the additional LDOS in-duced by the seven vortices. Each vortex is ap-parent as a checkerboard at 45° to the pageorientation. Not all are identical, most likely be-cause of the effects of electronic inhomogeneity.The units of S1

12(x, y, 5) are picoamps because itrepresents /dI/dV!0V. In this energy range, themaximum integrated LDOS at a vortex is "3pA, as compared with the zero field integratedLDOS of "1 pA. The latter is subtracted fromthe former to give a maximum contrast of "2pA. We also note that the integrated differen-tial conductance between 0 and *200 meV is200 pA because all measurements reported inthis paper were obtained at a junction resis-tance of 1 gigaohm set at a bias voltage of–200 mV.

R E P O R T S

www.sciencemag.org SCIENCE VOL 295 18 JANUARY 2002 467

on

Mar

ch 1

, 201

3w

ww

.sci

ence

mag

.org

Dow

nloa

ded

from

Monday, April 29, 13

Page 96: Quantum criticality in condensed matter

LETTERSPUBLISHED ONLINE: 14 OCTOBER 2012 | DOI: 10.1038/NPHYS2456

Direct observation of competition betweensuperconductivity and charge density waveorder in YBa2Cu3O6.67

J. Chang1,2*, E. Blackburn3, A. T. Holmes3, N. B. Christensen4, J. Larsen4,5, J. Mesot1,2,Ruixing Liang6,7, D. A. Bonn6,7, W. N. Hardy6,7, A. Watenphul8, M. v. Zimmermann8, E. M. Forgan3

and S. M. Hayden9

Superconductivity often emerges in the proximity of, or incompetition with, symmetry-breaking ground states such asantiferromagnetism or charge density waves1–5 (CDW). Anumber of materials in the cuprate family, which includes thehigh transition-temperature (high-Tc) superconductors, showspin and charge density wave order5–7. Thus a fundamentalquestion is to what extent do these ordered states existfor compositions close to optimal for superconductivity.Here we use high-energy X-ray diffraction to show thata CDW develops at zero field in the normal state ofsuperconducting YBa2Cu3O6.67 (Tc = 67K). This sample hasa hole doping of 0.12 per copper and a well-ordered oxygenchain superstructure8. Below Tc, the application of a magneticfield suppresses superconductivity and enhances the CDW.Hence, the CDW and superconductivity in this typical high-Tcmaterial are competing orders with similar energy scales,and the high-Tc superconductivity forms from a pre-existingCDW environment. Our results provide a mechanism for theformation of small Fermi surface pockets9, which explain thenegative Hall and Seebeck effects10,11 and the ‘Tc plateau’12 inthis material when underdoped.

Charge density waves in solids are periodic modulations of con-duction electron density. They are often present in low-dimensionalsystems such as NbSe2 (ref. 4). Certain cuprate materials such asLa2�x�yNdySrxCuO4 (Nd-LSCO) and La2�xBaxCuO4 (LBCO) alsoshow charge modulations that suppress superconductivity near x =1/8 (refs 6,7). In some cases, these are believed to be unidirectionalin the CuO2 plane, and have been dubbed ‘stripes’2,3. There is now amounting body of indirect evidence that charge and/or spin densitywaves (static modulations) may be present at high magnetic fieldsin samples with high Tc: quantum oscillation experiments on un-derdoped YBa2Cu3Oy (YBCO) have revealed the existence of at leastone small Fermi surface pocket9,10, whichmay be created by a chargemodulation11. More recently, nuclear magnetic resonance (NMR)studies have shown a magnetic-field-induced splitting of the Cu2Flines of YBCO (ref. 13). An important issue is the extent towhich thetendency towards charge order exists in high-Tc superconductors2,3.

Here we report a hard (100 keV) X-ray diffraction study, inmagnetic fields up to 17 T, of a detwinned single crystal of

1Institut de la Matière Complexe, Ecole Polytechnique Fedérale de Lausanne (EPFL), CH-1015 Lausanne, Switzerland, 2Paul Scherrer Institut, Swiss LightSource, CH-5232 Villigen PSI, Switzerland, 3School of Physics and Astronomy, University of Birmingham, Birmingham, B15 2TT, UK, 4Department ofPhysics, Technical University of Denmark, DK-2800 Kongens Lyngby, Denmark, 5Laboratory for Neutron Scattering, Paul Scherrer Institut, CH-5232Villigen PSI, Switzerland, 6Department of Physics and Astronomy, University of British Columbia, Vancouver, Canada, 7Canadian Institute for AdvancedResearch, Toronto, Canada, 8Hamburger Synchrotronstrahlungslabor (HASYLAB) at Deutsches Elektronen-Synchrotron (DESY), 22603 Hamburg,Germany, 9H. H. Wills Physics Laboratory, University of Bristol, Bristol, BS8 1TL, UK. *e-mail: [email protected].

YBa2Cu3O6.67 (with ortho-VIII oxygen ordering8,12, Tc = 67Kand p = 0.12, where p is the hole concentration per planarCu). We find that a CDW forms in the normal state belowTCDW ⇡ 135K. The charge modulation has two fundamentalwave vectors qCDW = q1 = (�1,0,0.5) and q2 = (0,�2,0.5), where�1 ⇡ 0.3045(2) and �2 ⇡ 0.3146(7), with no significant field- ortemperature-dependence of these values. The CDW gives riseto satellites of the parent crystal Bragg peaks at positions suchas Q = (2 ± �1,0,0.5). Although the satellite intensities have astrong temperature and magnetic field dependence, the CDW isnot field-induced and is unaffected by field in the normal state.Below Tc it competes with superconductivity, and a decrease ofthe CDW amplitude in zero field becomes an increase whensuperconductivity is suppressed by field. A very recent paper14reports complementary resonant soft X-ray scattering experimentsperformed on (Y,Nd)Ba2Cu3O6+x as a function of doping and inthe absence of amagnetic field. The results are broadly in agreementwith our zero field data.

Figure 1a,g shows scans through the (2� �1,0,0.5) and (0,2��2,0.5) positions at T = 2K. Related peaks were observed at(2+�1,0,0.5) and (4��1,0,0.5) (see Supplementary Fig. S3). Theincommensurate peaks are not detected above 150K (Fig. 1c). Fromthe peak width we estimate that the modulation has an in-planecorrelation length ⇠a ⇡ 95± 5Å (at 2 K and 17 T—see Methods).The existence of four similar in-plane modulations (±�1,0) and(0,±�2) indicates that the modulation is associated with the (nearlysquare) CuO2 planes rather than the CuO chains. The presentexperiment cannot distinguish between 1�q and 2�q structures,that is, we cannot tell directly whether modulations along the a andb directions co-exist in space or occur in different domains of thecrystal. However, Bragg peaks from the twoCDWcomponents havesimilar intensities and widths (Fig. 1b,g) despite the orthorhombiccrystal structure, which breaks the symmetry between them. Thissuggests that q1 and q2 are coupled, leading to the co-existence ofmultiple wave vectors, as seen in other CDW systems such as NbSe2(ref. 4). The scan along the c⇤ direction in Fig. 1d has broad peaksclose to l = ±0.5 reciprocal lattice units (r.l.u.), indicating that theCDW is weakly correlated along the c direction, with a correlationlength ⇠c of approximately 0.6 lattice units.

NATURE PHYSICS | VOL 8 | DECEMBER 2012 | www.nature.com/naturephysics 871

LETTERSPUBLISHED ONLINE: 14 OCTOBER 2012 | DOI: 10.1038/NPHYS2456

Direct observation of competition betweensuperconductivity and charge density waveorder in YBa2Cu3O6.67

J. Chang1,2*, E. Blackburn3, A. T. Holmes3, N. B. Christensen4, J. Larsen4,5, J. Mesot1,2,Ruixing Liang6,7, D. A. Bonn6,7, W. N. Hardy6,7, A. Watenphul8, M. v. Zimmermann8, E. M. Forgan3

and S. M. Hayden9

Superconductivity often emerges in the proximity of, or incompetition with, symmetry-breaking ground states such asantiferromagnetism or charge density waves1–5 (CDW). Anumber of materials in the cuprate family, which includes thehigh transition-temperature (high-Tc) superconductors, showspin and charge density wave order5–7. Thus a fundamentalquestion is to what extent do these ordered states existfor compositions close to optimal for superconductivity.Here we use high-energy X-ray diffraction to show thata CDW develops at zero field in the normal state ofsuperconducting YBa2Cu3O6.67 (Tc = 67K). This sample hasa hole doping of 0.12 per copper and a well-ordered oxygenchain superstructure8. Below Tc, the application of a magneticfield suppresses superconductivity and enhances the CDW.Hence, the CDW and superconductivity in this typical high-Tcmaterial are competing orders with similar energy scales,and the high-Tc superconductivity forms from a pre-existingCDW environment. Our results provide a mechanism for theformation of small Fermi surface pockets9, which explain thenegative Hall and Seebeck effects10,11 and the ‘Tc plateau’12 inthis material when underdoped.

Charge density waves in solids are periodic modulations of con-duction electron density. They are often present in low-dimensionalsystems such as NbSe2 (ref. 4). Certain cuprate materials such asLa2�x�yNdySrxCuO4 (Nd-LSCO) and La2�xBaxCuO4 (LBCO) alsoshow charge modulations that suppress superconductivity near x =1/8 (refs 6,7). In some cases, these are believed to be unidirectionalin the CuO2 plane, and have been dubbed ‘stripes’2,3. There is now amounting body of indirect evidence that charge and/or spin densitywaves (static modulations) may be present at high magnetic fieldsin samples with high Tc: quantum oscillation experiments on un-derdoped YBa2Cu3Oy (YBCO) have revealed the existence of at leastone small Fermi surface pocket9,10, whichmay be created by a chargemodulation11. More recently, nuclear magnetic resonance (NMR)studies have shown a magnetic-field-induced splitting of the Cu2Flines of YBCO (ref. 13). An important issue is the extent towhich thetendency towards charge order exists in high-Tc superconductors2,3.

Here we report a hard (100 keV) X-ray diffraction study, inmagnetic fields up to 17 T, of a detwinned single crystal of

1Institut de la Matière Complexe, Ecole Polytechnique Fedérale de Lausanne (EPFL), CH-1015 Lausanne, Switzerland, 2Paul Scherrer Institut, Swiss LightSource, CH-5232 Villigen PSI, Switzerland, 3School of Physics and Astronomy, University of Birmingham, Birmingham, B15 2TT, UK, 4Department ofPhysics, Technical University of Denmark, DK-2800 Kongens Lyngby, Denmark, 5Laboratory for Neutron Scattering, Paul Scherrer Institut, CH-5232Villigen PSI, Switzerland, 6Department of Physics and Astronomy, University of British Columbia, Vancouver, Canada, 7Canadian Institute for AdvancedResearch, Toronto, Canada, 8Hamburger Synchrotronstrahlungslabor (HASYLAB) at Deutsches Elektronen-Synchrotron (DESY), 22603 Hamburg,Germany, 9H. H. Wills Physics Laboratory, University of Bristol, Bristol, BS8 1TL, UK. *e-mail: [email protected].

YBa2Cu3O6.67 (with ortho-VIII oxygen ordering8,12, Tc = 67Kand p = 0.12, where p is the hole concentration per planarCu). We find that a CDW forms in the normal state belowTCDW ⇡ 135K. The charge modulation has two fundamentalwave vectors qCDW = q1 = (�1,0,0.5) and q2 = (0,�2,0.5), where�1 ⇡ 0.3045(2) and �2 ⇡ 0.3146(7), with no significant field- ortemperature-dependence of these values. The CDW gives riseto satellites of the parent crystal Bragg peaks at positions suchas Q = (2 ± �1,0,0.5). Although the satellite intensities have astrong temperature and magnetic field dependence, the CDW isnot field-induced and is unaffected by field in the normal state.Below Tc it competes with superconductivity, and a decrease ofthe CDW amplitude in zero field becomes an increase whensuperconductivity is suppressed by field. A very recent paper14reports complementary resonant soft X-ray scattering experimentsperformed on (Y,Nd)Ba2Cu3O6+x as a function of doping and inthe absence of amagnetic field. The results are broadly in agreementwith our zero field data.

Figure 1a,g shows scans through the (2� �1,0,0.5) and (0,2��2,0.5) positions at T = 2K. Related peaks were observed at(2+�1,0,0.5) and (4��1,0,0.5) (see Supplementary Fig. S3). Theincommensurate peaks are not detected above 150K (Fig. 1c). Fromthe peak width we estimate that the modulation has an in-planecorrelation length ⇠a ⇡ 95± 5Å (at 2 K and 17 T—see Methods).The existence of four similar in-plane modulations (±�1,0) and(0,±�2) indicates that the modulation is associated with the (nearlysquare) CuO2 planes rather than the CuO chains. The presentexperiment cannot distinguish between 1�q and 2�q structures,that is, we cannot tell directly whether modulations along the a andb directions co-exist in space or occur in different domains of thecrystal. However, Bragg peaks from the twoCDWcomponents havesimilar intensities and widths (Fig. 1b,g) despite the orthorhombiccrystal structure, which breaks the symmetry between them. Thissuggests that q1 and q2 are coupled, leading to the co-existence ofmultiple wave vectors, as seen in other CDW systems such as NbSe2(ref. 4). The scan along the c⇤ direction in Fig. 1d has broad peaksclose to l = ±0.5 reciprocal lattice units (r.l.u.), indicating that theCDW is weakly correlated along the c direction, with a correlationlength ⇠c of approximately 0.6 lattice units.

NATURE PHYSICS | VOL 8 | DECEMBER 2012 | www.nature.com/naturephysics 871

NATURE PHYSICS DOI: 10.1038/NPHYS2456 LETTERS

Q = (1.695, 0, 0.5)

T CDW

T C

T cusp

Q = (0, 3.691, 0.5) (x4)

17 T

2 K

2 K

66 K

66 K

150 K

7.5 T

0 T

10¬3

)

T (K) H (T)

Wid

th

(r.l

.u.)

of (

2-

, 0, 0

.5)

Inte

nsity

(cn

ts s

¬1)

Inte

nsity

(cn

ts s

¬1)

15 T

200

150

100

50

0

12

11

10

9

8

7

60 50 100 150

T (K)

0 5 10 15

H (T)

12

11

10

9

8

7

6

200

150

100

50

0

a

c

b

d

0 50 100 150 0 5 10 15

δ

10¬3

)W

idth

(

r.l.u

.) o

f (2-

, 0

, 0.5

σ σ

Figure 2 | Competition between charge–density-wave order and superconductivity. a, Temperature dependence of the peak intensity at (1.695, 0, 0.5)(circles) and (0, 3.691, 0.5) (squares) for different applied magnetic fields. The square data points have been multiplied by a factor of four. In the normalstate, there is a smooth onset of the CDW order. In the absence of an applied magnetic field there is a decrease in the peak intensity below Tc. This trendcan be reversed by the application of a magnetic field. b, Magnetic field dependence of the lattice modulation peak intensity at (1.695,0,0.5) for differenttemperatures. At T = 2 K, the peak intensity grows approximately linearly with magnetic field up to the highest applied field. c,d, Gaussian linewidth of the(1.695, 0, 0.5) CDW modulation plotted versus temperature and field respectively. The raw linewidth, including a contribution from the instrumentalresolution, is field-independent in the normal state (T > Tc). In contrast, the CDW order becomes more coherent below Tc, once a magnetic field is applied.This effect ceases once the amplitude starts to be suppressed owing to competition with superconductivity. The vertical dashed lines in a,c illustrate theconnection between these two features of the data that define the Tcusp temperatures. All other lines are guides to the eye. Error bars indicate standarddeviations of the fit parameters described in Methods.

The intensities of the incommensurate Bragg peaks are sensitiveto atomic displacements parallel to the total scattering vectorQ. The comparatively small contribution to Q along the c⇤

direction from l = 0.5 r.l.u. means that our signal for a (h,0, 0.5) peak is dominated by displacements parallel to the adirection. (There will also be displacements parallel to the cdirection but we are essentially insensitive to them in our presentscattering geometry). Our data indicate that the incommensuratepeaks are much stronger if they are satellites of strong Braggpeaks of the form (⌧ = (2n,0,0)) at positions such as ⌧ ± q1.This indicates that the satellites are caused by a modulationof the parent crystal structure. The fact that the scattering ispeaked at l = ±0.5 r.l.u. means that neighbouring bilayers aremodulated in antiphase. The two simplest structures (Fig. 3a,b)compatible with our data (see Supplementary Information) involvethe neighbouring CuO2 planes in the bilayer being displaced inthe same (bilayer-centred) or opposite (chain-centred) directions,resulting in the maximum amplitude of the modulation being onthe CuO2 planes or CuO chains respectively. In their 2�q form,these structures would lead to the in-plane ‘checkerboard’ patternshown in Fig. 3c. Scanning tunnelling microscopy studies of otherunderdoped cuprates16 and of field-induced CDW correlations invortex cores17 also support the tendency towards checkerboardformation18, although disorder can cause small stripe domainsto mimic checkerboard order19. Our observation of a CDW

may be related to phonon anomalies20, which suggest that inYBCO near p⇡ 1/8 there are anomalies in the underlying chargesusceptibility for q⇡ (0,0.3).

Cuprate superconductors show strong spin correlations, andthe interplay between spin and charge correlations may be at theheart of the high-Tc phenomenon. The spin correlations are largelydynamic, with energies up to several hundred meV. YBa2Cu3O6+xand La2�x(Ba,Sr)xCuO4+� show incommensurate magnetic order,which can be enhanced by suppressing superconductivity with anapplied magnetic field21–24; this has some analogies with the CDWorder observed here. The magnetic order is static on the ⇠1meVfrequency scale of neutron diffraction and has been detected inlightly doped YBa2Cu3O6+x for p 0.082 (ref. 21), and moderatelydoped La2�xSrxCuO4 for p 0.14 (ref. 24). The YBa2Cu3O6.67(p⇡ 1/8) sample studied here is expected to have a relatively largespin gap, h! ⇡ 20meV (ref. 25), in its magnetic excitations atlow temperature, making it unlikely that it orders magnetically.As discussed earlier, this is confirmed by other measurements13,14,so the CDW does not seem to be accompanied by spin order.Moreover, there is no obvious relationship between qCDW and thewave vector of the incipient spin fluctuations qSF ⇡ (0.1,0) ofsimilarly doped samples25.

It is interesting to note that TCDW corresponds approximatelywithTH (Fig. 4), the temperature at whichHall effectmeasurementssuggest that Fermi surface reconstruction begins26. A CDW that

NATURE PHYSICS | VOL 8 | DECEMBER 2012 | www.nature.com/naturephysics 873

NATURE PHYSICS DOI: 10.1038/NPHYS2456 LETTERS

0

50

100

150

Tem

pera

ture

(K

)

Tem

pera

ture

(K

)200

250

300

0.00 0.05 0.10Doping, p (holes/Cu) 0H (T)

T∗TN

TH

TH

TSDW

TCDW

Tcusp

TVL

TCDW

TNMR

0.15

SCVortexsolid

YBCO YBCO p = 0.12

Vortex liquidAF

0.20 0 10 20 30 40 500

40

80

120

160a b

µ

Figure 4 | Phase diagram of YBa2Cu3O7�x

. a, Doping dependence of the antiferromagnetic ordering temperature TN, the incommensurate spin-densitywave order TSDW (green triangles; ref. 21), the superconducting temperature Tc and the pseudogap temperature T⇤ as determined from the Nernst effect30

(black squares) and neutron diffraction29 (purple squares). Notice that the Nernst effect30 indicates a broken rotational symmetry inside the pseudogapregion, whereas a translational symmetry preserving magnetic order is found by neutron scattering29. Below temperature scale TH (black circles), a largerand negative Hall coefficient was observed26 and interpreted in terms of a Fermi surface reconstruction. Our X-ray diffraction experiments show that inYBCO p = 0.12 incommensurate CDW order spontaneously breaks the crystal translational symmetry at a temperature TCDW that is twice as large as Tc.TCDW is also much larger than TNMR (red squares), the temperature scale below which NMR observes field-induced charge order13. b, Field dependence ofTCDW (filled red circles) and Tcusp (open squares), the temperature below which the CDW is suppressed by superconductivity, compared with TH (openblack circle) and TVL (filled blue circles), the temperature where the vortex liquid state forms26. Error bars on TSDW, TH, TNMR, and T⇤ are explained inrefs 21,26,30,33. The error bars on TCDW and Tcusp reflect the uncertainty in determining the onset and suppression temperature of CDW order from Fig. 2.

these various orders are ‘intertwined’31. In this context, we canview our present results as indicating that the electron systemhas a tendency towards two ground states: a charge densitywave, which breaks translational symmetry and involves electron–hole correlations, versus superconductivity, which breaks gaugesymmetry and involves electron–electron correlations.We note thatthe q-vectors of the CDW lie close to the separation of pieces ofFermi surface that have maximum superconducting gap at optimaldoping and have the same sign of the order parameter.

MethodsOur experiments used 100 keV hard X-ray synchrotron radiation from theDORIS-III storage ring at DESY, Hamburg, Germany. We installed a recentlydeveloped 17 T horizontal cryomagnet designed for beamline use on the triple-axisdiffractometer at beamline BW5. The sample was mounted by gluing it over a holein a temperature-controlled aluminium plate within the cryomagnet vacuum andwas thermally shielded by thin Al and aluminizedmylar foils glued to this plate. Thesample temperature could be controlled over the range ⇠2–300K. The incomingand outgoing beams passed through 1mm thick aluminium cryostat vacuumwindows, which gave a maximum of ⇠ ±10� input and output angles relative tothe field direction, which was parallel to the sample c axis within <1�. Betweenthe beam access windows and the sample plate, there were further aluminiumfoil thermal radiation shields at liquid nitrogen temperature. A 2mm squareaperture collimated the incoming beam, so that it passed mainly through the partof the sample over the hole in the aluminium plate, greatly reducing backgroundscattering by the plate. Further slits before the analyser and the detector removedscattering by the cryostat windows and nitrogen shields. The scattering plane(a⇤–c⇤) was horizontal. The cryomagnet was mounted on a rotation stage with agoniometer giving � tilt about the field axis. The sample was initially mountedwith its a axis nearly horizontal. The � goniometer allowed the exact alignment ofthis axis using the (2 0 0) Bragg peak and could also be used for low-resolutionscans in the b⇤ direction. Magnetic fields were applied with the sample heatedabove Tc; it was then field-cooled to base temperature. When fields were applied,minor changes in the position and angle of the sample holder were observed; thesewere corrected by use of horizontal and vertical motion stages under the cryostatrotation stage, and by realigning on the (2 0 0) Bragg peak. During temperaturescans, realignment on the (2 0 0) Bragg peak was performed automatically at everytemperature point to ensure that all measurements were centred. After resultshad been obtained with the a axis horizontal, the sample was remounted withthe b axis horizontal for further measurements. The YBa2Cu3O6.67 sample haddimensions a⇥b⇥ c = 3.1⇥1.7⇥0.6mm3 and mass 18mg. The superconductingtransition temperature Tc = 67K (width: 10%–90%= 1.1K) was derived froma zero-field-cooled magnetization curve at 0.1mT. The single crystal was 99%

detwinned and the Cu–O chains were ordered with the ortho-VIII structure bystandard procedures12.

The diffracted intensities from the CDW, shown in Fig. 1, are composed ofan incommensurate lattice modulation peak on a smoothly varying background.The background along (h, 0, 0.5) mainly originates from the tails of the ortho-VIIIpeaks (see Supplementary Information). It varies strongly from one Brillouinzone to another; for example, the background around (2.7, 0, 0.5) is an order ofmagnitude larger than around (1.7, 0, 0.5). The background has essentially nofield dependence (Fig. 1a–c) so subtracting the zero-field from high-field data isa simple way to eliminate the background. This reveals the field-enhanced signalinside the superconducting state (Fig. 1a–d).

As there is a weak temperature dependence in the background (Fig. 1a–c), itis not possible to eliminate it by subtracting a high-temperature curve. Therefore,to obtain the temperature dependences shown in Fig. 2, we fitted the data toa Gaussian function G(Q) and modelled the background by a second-orderpolynomial B(Q)= c0 + c1Q+ c2Q2. The constants c0, c1 and c2 have a smallbut significant temperature dependence. The low counting statistics resulted inGaussians fitting equally well as other possible lineshapes such as Lorentzians.

The signal-to-background ratio is best for the (2� �1, 0, 0.5) peak duethe weaker structural ortho-VIII peak (see Supplementary Fig. S2). From theGaussian fits to the (2��1, 0, 0.5) satellite peak at 2 K and 17 T we can estimatethe correlation length ⇠ along the three crystal axis directions. We define ⇠ = 1/� ,where � = (� 2

meas�� 2R)0.5 is the measured Gaussian standard deviation corrected for

the instrument resolution �R and expressed in Å�1. Along the a axis direction, wefind � = 6.4⇥10�3 r.l.u. ⌘ 1.1⇥10�2 Å�1, and hence ⇠a = 95±5Å. Deconvolvingthe poor instrumental resolution along the b axis direction for the (2��1, 0, 0.5)peak yields a similar correlation length ⇠b ⇠ ⇠a.

Received 18 June 2012; accepted 18 September 2012;published online 14 October 2012

References1. Mathur, N. D. et al. Magnetically mediated superconductivity in heavy fermion

compounds. Nature 394, 39–43 (1998).2. Kivelson, S. A. et al. How to detect fluctuating stripes in the high-temperature

superconductors. Rev. Mod. Phys. 75, 1201–1241 (2003).3. Vojta, M. Lattice symmetry breaking in cuprate superconductors: Stripes,

nematics, and superconductivity. Adv. Phys. 58, 699–820 (2009).4. Moncton, D. E., Axe, J. D. & DiSalvo, F. J. Neutron scattering study of the

charge–density wave transitions in 2H–TaSe2 and 2H–NbSe2. Phys. Rev. B 16,801–819 (1977).

5. Demler, E., Sachdev, S. & Zhang, Y. Spin-ordering quantum transitions ofsuperconductors in a magnetic field. Phys. Rev. Lett. 87, 067202 (2001).

6. Tranquada, J. M. et al. Evidence for stripe correlations of spins and holes incopper oxide superconductors. Nature 375, 561–563 (1995).

NATURE PHYSICS | VOL 8 | DECEMBER 2012 | www.nature.com/naturephysics 875

LETTERS NATURE PHYSICS DOI: 10.1038/NPHYS2456

17 T | |c0 T

17 T | |c0 T

17 T | |c0 T

T = 150 K

Cut 2

Cut 4 Cut 4 Cut 4

I (cn

ts s

¬1)

h in (h, 0, 0.5) (r.l.u.) h in (h, 0, 0.5) (r.l.u.) h in (h, 0, 0.5) (r.l.u.)

k in (0, k, 0.5) (r.l.u.) k in (0, k, 0.5) (r.l.u.)

I (17

T)-

I (0

T) (

cnts

s¬1

)

Cut 3

200

I (cn

ts s

¬1)

2 K

65 K

0 T

65 K0 T

65 K0 T

x0.625

¬1.0 ¬0.5l in (1.696, 0, 1) (r.l.u.)

450

400

350

300

250

200

150

I (cn

ts s

¬1)

450

400

350

300

250

0.5Cut 1

200

150

I (cn

ts s

¬1)

450

400

350

300

250

200

150

1.64 1.66 1.68 1.70 1.72 1.74

T = 2 K T = 66 K

1.64 1.66 1.68

0.3 (0, 0, 0)

0.3

k

h

1.70 1.72 1.74 1.64 1.66 1.68 1.70 1.72 1.74

Cut 1 Cut 1 Cut 1

0

50

100

150

0.0 0.5 1.0

100

150

200

I (cn

ts s

¬1)

100

150

200

I (cn

ts s

¬1)

100

150

200

1.64 1.66 1.68 1.70 1.72 1.74

Cut 2

Cut 4

I (cn

ts s

¬1)

350

300

250

200

150

100

2.26 2.28 2.30 2.32 2.34 2.36k in (0, k, 0.5) (r.l.u.)

3.64 3.66 3.68 3.70 3.72 3.74

h i

e f

b c

–0.05 0.00k in (1.694, k, 0.5) (r.l.u.)

0.05

Cut 350 K0 T

a

d

g

l

Figure 1 | Incommensurate charge–density-wave order. Diffracted intensity in reciprocal space Q= (h,k,l) = ha⇤ +kb⇤ + lc⇤ where a⇤ = 2⇡/a, b⇤ = 2⇡/band c⇤ = 2⇡/c, with lattice parameters a = 3.81 Å, b = 3.87 Å (Supplementary Fig. S1), c = 11.72 Å. Four different scans in reciprocal space, projected intothe first Brillouin zone, are shown schematically in e. a–c, Scans along (h,0,0.5) for temperatures and magnetic fields (applied along the crystalc-direction) as indicated. An incommensurate lattice modulation, peaked at (2��1, 0, 0), where �1 = 0.3045(2), emerges as the temperature is loweredbelow 135 K. The intensity of the satellite in b is of the order 2⇥ 10�6 weaker than the (2, 0, 0) reflection. This becomes field-dependent below thezero-field superconducting transition temperature Tc = 67 K. The full-width half-maximum instrumental resolution is shown by horizontal lines in b,f. Bydeconvolving the resolution from the Gaussian fits to the data taken at 17 T and 2 K, an h-width of �a = 6.4⇥ 10�3 r.l.u. corresponding to a correlationlength ⇠a = 1/�a of 95±5 Å was found (see Methods). d, The field-induced signal I(17 T)� I(0 T) at T = 2 K is modulated along (1.695, 0, l) and peaks atapproximately l = ±0.5. f, Scan along (1.695, k, 0.5). The poor resolution along the k-direction did not allow accurate determination of the width along(1.695, k, 0.5), but we estimate a value of 0.01 r.l.u., comparable to that along (h, 0, 0.5), indicating similar coherence lengths along a- and b-axis directions.g–i, Scans along (0, k, 0.5). Incommensurate peaks are found in several Brillouin zones, for example, at positions Q= (0,2±�2,0.5) and (0,4��2,0.5),where �2 = 0.3146(7), see also Supplementary Fig. S3. The vertical dashed line in g indicates �1 whereas the line in a indicates �2. The lattice modulationwas fitted to a Gaussian function (solid lines in a–d,f–i) on a background (dashed lines) modelled by a second-order polynomial. Error bars are determinedby counting statistics.

In zero field, the intensity of the CDWBragg peak (Fig. 2) growson cooling to Tc, below which it is partially suppressed. For T >Tc,a magnetic field applied along the c direction has no effect. BelowTc it causes an increase of the intensity of the CDW signal (Figs 1aand 2). At T = 2K, the intensity grows with applied magnetic field(Fig. 2b) and shows no signs of saturation up to 17 T. The magneticfield also makes the CDW more long-range ordered (Fig. 2c). Inzero magnetic field, the q-width varies little with temperature.However, below Tc in a field, the CDW order not only becomesstronger, but also becomes more coherent, down to a temperatureTcusp below which the intensity starts to decrease (Figs 2 and 4).All of this is clear evidence for competition between CDW andsuperconducting orders.

Non-resonant X-ray diffraction is sensitive to modulations ofcharge density and magnetic moments. In our case, the expectedmagnetic cross-section is several orders of magnitude smaller thanour observed signal, which must therefore be due to charge scatter-ing. NMR measurements on a sample of the same composition as

ours13 indicate that the CDW is not accompanied by magnetic or-der, and this is confirmed by soft X-raymeasurements, whichwouldalso be sensitive to fluctuating order14. Charge density modulationsin solids will always involve both a modulation of the electroniccharge and a periodic displacement of the atomic positions15. Weare more sensitive to the atomic displacements than to the chargemodulation because ions with large numbers of electrons (as inYBCO)dominate the scattering (see Supplementary Information).

NMR data13 suggest that CDW order only appears belowT ⇡ 67K andH > 9 T, whereas with X-rays we observe CDW orderin zero field up to 135K. This apparent discrepancy may arisefrom differing timescales of various probes (see SupplementaryInformation for further discussion). X-ray diffraction experimentsare usually interpreted as measuring the static order of a givenstructure, but, if performed with wide energy acceptance, arealso sensitive to short-lived structures. Thus, it is possible thatthe observed CDW is quasi-static and only frozen on the NMRtimescale (⇡3 ns) at high fields and lower temperatures.

872 NATURE PHYSICS | VOL 8 | DECEMBER 2012 | www.nature.com/naturephysics

Monday, April 29, 13

Page 97: Quantum criticality in condensed matter

LETTERdoi:10.1038/nature10345

Magnetic-field-induced charge-stripe order in thehigh-temperature superconductor YBa2Cu3OyTaoWu1, Hadrien Mayaffre1, Steffen Kramer1, Mladen Horvatic1, Claude Berthier1, W. N. Hardy2,3, Ruixing Liang2,3, D. A. Bonn2,3

& Marc-Henri Julien1

Electronic charges introduced in copper-oxide (CuO2) planesgenerate high-transition-temperature (Tc) superconductivity but,under special circumstances, they can also order into filamentscalled stripes1. Whether an underlying tendency towards chargeorder is present in all copper oxides and whether this has anyrelationship with superconductivity are, however, two highly con-troversial issues2,3. To uncover underlying electronic order, mag-netic fields strong enough to destabilize superconductivity can beused. Such experiments, including quantum oscillations4–6 inYBa2Cu3Oy (an extremely clean copper oxide in which chargeorder has not until now been observed) have suggested that super-conductivity competes with spin, rather than charge, order7–9. Herewe report nuclear magnetic resonance measurements showing thathigh magnetic fields actually induce charge order, without spinorder, in the CuO2 planes of YBa2Cu3Oy. The observed static, uni-directional, modulation of the charge density breaks translationalsymmetry, thus explaining quantum oscillation results, and weargue that it is most probably the same 4a-periodic modulationas in stripe-ordered copper oxides1. That it develops only whensuperconductivity fades away and near the same 1/8 hole dopingas in La22xBaxCuO4 (ref. 1) suggests that charge order, althoughvisibly pinned by CuO chains in YBa2Cu3Oy, is an intrinsic pro-pensity of the superconducting planes of high-Tc copper oxides.The ortho II structure of YBa2Cu3O6.54 (p5 0.108, where p is the

hole concentration per planar Cu) leads to two distinct planar CuNMR sites: Cu2F are those Cu atoms located below oxygen-filledchains, and Cu2E are those below oxygen-empty chains10. The maindiscovery of ourwork is that, on cooling in a fieldH0 of 28.5 T along thec axis (that is, in the conditions for which quantum oscillations areresolved; see Supplementary Materials), the Cu2F lines undergo aprofound change, whereas theCu2E lines do not (Fig. 1). To first order,this change can be described as a splitting of Cu2F into two sites havingboth different hyperfine shiftsK5 Æhzæ/H0 (where Æhzæ is the hyperfinefield due to electronic spins) and quadrupole frequencies nQ (related tothe electric field gradient). Additional effects might be present (Fig. 1),but they areminor in comparisonwith the observed splitting. Changesin field-dependent and temperature-dependent orbital occupancy (forexample dx2{y2 versus dz2{r2 ) without on-site change in electronicdensity are implausible, and any change in out-of-plane charge densityor lattice would affect Cu2E sites as well. Thus, the change in nQ canonly arise from a differentiation in the charge density between Cu2Fsites (or at the oxygen sites bridging them). A change in the asymmetryparameter and/or in the direction of the principal axis of the electricfield gradient could also be associated with this charge differentiation,but these are relatively small effects.The charge differentiation occurs below Tcharge5 506 10K for

p5 0.108 (Fig. 1 and Supplementary Figs 9 and 10) and 676 5K forp5 0.12 (Supplementary Figs 7 and 8). Within error bars, for each ofthe samples Tcharge coincides with T0, the temperature at which theHall constant RH becomes negative, an indication of the Fermi surface

reconstruction11–13. Thus, whatever the precise profile of the staticcharge modulation is, the reconstruction must be related to the trans-lational symmetry breaking by the charge ordered state.The absence of any splitting or broadening of Cu2E lines implies a

one-dimensional character of the modulation within the planes andimposes strong constraints on the charge pattern. Actually, only twotypes of modulation are compatible with a Cu2F splitting (Fig. 2). Thefirst is a commensurate short-range (2a or 4a period) modulationrunning along the (chain) b axis. However, this hypothesis is highlyunlikely: to the best of our knowledge, no such modulation has everbeen observed in the CuO2 planes of any copper oxide; it would there-fore have to be triggered by a charge modulation pre-existing in thefilled chains. A charge-density wave is unlikely because the finite-sizechains are at best poorly conducting in the temperature and dopingrange discussed here11,14. Any inhomogeneous charge distributionsuch as Friedel oscillations around chain defects would broaden ratherthan split the lines. Furthermore, we can conclude that charge orderoccurs only for high fields perpendicular to the planes because theNMR lines neither split at 15T nor split in a field of 28.5 T parallelto the CuO2 planes (along either a or b), two situations in whichsuperconductivity remains robust (Fig. 1). This clear competitionbetween charge order and superconductivity is also a strong indicationthat the charge ordering instability arises from the planes.Theonlyother patterncompatiblewithNMRdata is an alternationof

more and less charged Cu2F rows defining a modulation with a periodof four lattice spacings along the a axis (Fig. 2). Strikingly, this corre-sponds to the (site-centred) charge stripes found in La22xBaxCuO4 atdoping levels near p5 x5 0.125 (ref. 1). Being a proven electronicinstability of the planes, which is detrimental to superconductivity2,stripe ordernot onlyprovides a simple explanationof theNMRsplittingbut also rationalizes the striking effect of the field. Stripe order is alsofully consistent with the remarkable similarity of transport data inYBa2Cu3Oy and in stripe-ordered copper oxides (particularly thedome-shaped dependence ofT0 around p5 0.12)11–13. However, stripesmust be parallel from plane to plane in YBa2Cu3Oy, whereas they areperpendicular in, for example, La22xBaxCuO4. We speculate that thisexplains why the charge transport along the c axis in YBa2Cu3Oy

becomes coherent in high fields below T0 (ref. 15). If so, stripe fluctua-tions must be involved in the incoherence along c above T0.Once we know the doping dependence of nQ (ref. 16), the difference

DnQ5 3206 50 kHz for p5 0.108 implies a charge density variationas small as Dp5 0.036 0.01 hole between Cu2Fa and Cu2Fb. Acanonical stripe description (Dp5 0.5 hole) is therefore inadequateat the NMR timescale of ,1025 s, at which most (below T0) or all(above T0) of the charge differentiation is averaged out by fluctuationsfaster than 105 s21. This should not be a surprise: themetallic nature ofthe compound at all fields is incompatible with full charge order, evenif this order is restricted to the direction perpendicular to the stripes17.Actually, there is compelling evidence of stripe fluctuations down tovery low temperatures in stripe-ordered copper oxides18, and indirect

1Laboratoire National des Champs Magnetiques Intenses, UPR 3228, CNRS-UJF-UPS-INSA, 38042 Grenoble, France. 2Department of Physics and Astronomy, University of British Columbia, Vancouver,British Columbia V6T1Z1, Canada. 3Canadian Institute for Advanced Research, Toronto, Ontario M5G1Z8, Canada.

8 S E P T E M B E R 2 0 1 1 | V O L 4 7 7 | N A T U R E | 1 9 1

Macmillan Publishers Limited. All rights reserved©2011

LETTERdoi:10.1038/nature10345

Magnetic-field-induced charge-stripe order in thehigh-temperature superconductor YBa2Cu3OyTaoWu1, Hadrien Mayaffre1, Steffen Kramer1, Mladen Horvatic1, Claude Berthier1, W. N. Hardy2,3, Ruixing Liang2,3, D. A. Bonn2,3

& Marc-Henri Julien1

Electronic charges introduced in copper-oxide (CuO2) planesgenerate high-transition-temperature (Tc) superconductivity but,under special circumstances, they can also order into filamentscalled stripes1. Whether an underlying tendency towards chargeorder is present in all copper oxides and whether this has anyrelationship with superconductivity are, however, two highly con-troversial issues2,3. To uncover underlying electronic order, mag-netic fields strong enough to destabilize superconductivity can beused. Such experiments, including quantum oscillations4–6 inYBa2Cu3Oy (an extremely clean copper oxide in which chargeorder has not until now been observed) have suggested that super-conductivity competes with spin, rather than charge, order7–9. Herewe report nuclear magnetic resonance measurements showing thathigh magnetic fields actually induce charge order, without spinorder, in the CuO2 planes of YBa2Cu3Oy. The observed static, uni-directional, modulation of the charge density breaks translationalsymmetry, thus explaining quantum oscillation results, and weargue that it is most probably the same 4a-periodic modulationas in stripe-ordered copper oxides1. That it develops only whensuperconductivity fades away and near the same 1/8 hole dopingas in La22xBaxCuO4 (ref. 1) suggests that charge order, althoughvisibly pinned by CuO chains in YBa2Cu3Oy, is an intrinsic pro-pensity of the superconducting planes of high-Tc copper oxides.The ortho II structure of YBa2Cu3O6.54 (p5 0.108, where p is the

hole concentration per planar Cu) leads to two distinct planar CuNMR sites: Cu2F are those Cu atoms located below oxygen-filledchains, and Cu2E are those below oxygen-empty chains10. The maindiscovery of ourwork is that, on cooling in a fieldH0 of 28.5 T along thec axis (that is, in the conditions for which quantum oscillations areresolved; see Supplementary Materials), the Cu2F lines undergo aprofound change, whereas theCu2E lines do not (Fig. 1). To first order,this change can be described as a splitting of Cu2F into two sites havingboth different hyperfine shiftsK5 Æhzæ/H0 (where Æhzæ is the hyperfinefield due to electronic spins) and quadrupole frequencies nQ (related tothe electric field gradient). Additional effects might be present (Fig. 1),but they areminor in comparisonwith the observed splitting. Changesin field-dependent and temperature-dependent orbital occupancy (forexample dx2{y2 versus dz2{r2 ) without on-site change in electronicdensity are implausible, and any change in out-of-plane charge densityor lattice would affect Cu2E sites as well. Thus, the change in nQ canonly arise from a differentiation in the charge density between Cu2Fsites (or at the oxygen sites bridging them). A change in the asymmetryparameter and/or in the direction of the principal axis of the electricfield gradient could also be associated with this charge differentiation,but these are relatively small effects.The charge differentiation occurs below Tcharge5 506 10K for

p5 0.108 (Fig. 1 and Supplementary Figs 9 and 10) and 676 5K forp5 0.12 (Supplementary Figs 7 and 8). Within error bars, for each ofthe samples Tcharge coincides with T0, the temperature at which theHall constant RH becomes negative, an indication of the Fermi surface

reconstruction11–13. Thus, whatever the precise profile of the staticcharge modulation is, the reconstruction must be related to the trans-lational symmetry breaking by the charge ordered state.The absence of any splitting or broadening of Cu2E lines implies a

one-dimensional character of the modulation within the planes andimposes strong constraints on the charge pattern. Actually, only twotypes of modulation are compatible with a Cu2F splitting (Fig. 2). Thefirst is a commensurate short-range (2a or 4a period) modulationrunning along the (chain) b axis. However, this hypothesis is highlyunlikely: to the best of our knowledge, no such modulation has everbeen observed in the CuO2 planes of any copper oxide; it would there-fore have to be triggered by a charge modulation pre-existing in thefilled chains. A charge-density wave is unlikely because the finite-sizechains are at best poorly conducting in the temperature and dopingrange discussed here11,14. Any inhomogeneous charge distributionsuch as Friedel oscillations around chain defects would broaden ratherthan split the lines. Furthermore, we can conclude that charge orderoccurs only for high fields perpendicular to the planes because theNMR lines neither split at 15T nor split in a field of 28.5 T parallelto the CuO2 planes (along either a or b), two situations in whichsuperconductivity remains robust (Fig. 1). This clear competitionbetween charge order and superconductivity is also a strong indicationthat the charge ordering instability arises from the planes.Theonlyother patterncompatiblewithNMRdata is an alternationof

more and less charged Cu2F rows defining a modulation with a periodof four lattice spacings along the a axis (Fig. 2). Strikingly, this corre-sponds to the (site-centred) charge stripes found in La22xBaxCuO4 atdoping levels near p5 x5 0.125 (ref. 1). Being a proven electronicinstability of the planes, which is detrimental to superconductivity2,stripe ordernot onlyprovides a simple explanationof theNMRsplittingbut also rationalizes the striking effect of the field. Stripe order is alsofully consistent with the remarkable similarity of transport data inYBa2Cu3Oy and in stripe-ordered copper oxides (particularly thedome-shaped dependence ofT0 around p5 0.12)11–13. However, stripesmust be parallel from plane to plane in YBa2Cu3Oy, whereas they areperpendicular in, for example, La22xBaxCuO4. We speculate that thisexplains why the charge transport along the c axis in YBa2Cu3Oy

becomes coherent in high fields below T0 (ref. 15). If so, stripe fluctua-tions must be involved in the incoherence along c above T0.Once we know the doping dependence of nQ (ref. 16), the difference

DnQ5 3206 50 kHz for p5 0.108 implies a charge density variationas small as Dp5 0.036 0.01 hole between Cu2Fa and Cu2Fb. Acanonical stripe description (Dp5 0.5 hole) is therefore inadequateat the NMR timescale of ,1025 s, at which most (below T0) or all(above T0) of the charge differentiation is averaged out by fluctuationsfaster than 105 s21. This should not be a surprise: themetallic nature ofthe compound at all fields is incompatible with full charge order, evenif this order is restricted to the direction perpendicular to the stripes17.Actually, there is compelling evidence of stripe fluctuations down tovery low temperatures in stripe-ordered copper oxides18, and indirect

1Laboratoire National des Champs Magnetiques Intenses, UPR 3228, CNRS-UJF-UPS-INSA, 38042 Grenoble, France. 2Department of Physics and Astronomy, University of British Columbia, Vancouver,British Columbia V6T1Z1, Canada. 3Canadian Institute for Advanced Research, Toronto, Ontario M5G1Z8, Canada.

8 S E P T E M B E R 2 0 1 1 | V O L 4 7 7 | N A T U R E | 1 9 1

Macmillan Publishers Limited. All rights reserved©2011

evidence (explaining the rotational symmetry breaking) over a broadtemperature range in YBa2Cu3Oy (refs 14, 19–22). Therefore, insteadof being a defining property of the ordered state, the small amplitude ofthe charge differentiation is more likely to be a consequence of stripeorder (the smectic phase of an electronic liquid crystal17) remainingpartly fluctuating (that is, nematic).In stripe copper oxides, charge order at T5Tcharge is always accom-

panied by spin order at Tspin,Tcharge. Slowing down of the spin

fluctuations strongly enhances the spin–lattice (1/T1) and spin–spin(1/T2) relaxation rates between Tcharge and Tspin for

139La nuclei. Forthemore strongly hyperfine-coupled 63Cu, the relaxation rates becomeso large that the Cu signal is gradually ‘wiped out’ on cooling belowTcharge (refs 18, 23, 24). In contrast, the 63Cu(2) signal here inYBa2Cu3Oy does not experience any intensity loss and 1/T1 does notshow any peak or enhancement as a function of temperature (Fig. 3).Moreover, the anisotropy of the linewidth (SupplementaryInformation) indicates that the spins, although staggered, align mostlyalong the field (that is, c axis) direction, and the typical width of thecentral lines at base temperature sets an uppermagnitude for the staticspin polarization as small as gÆSzæ# 23 1023mB for both samples infields of,30T. These consistent observations rule out the presence ofmagnetic order, in agreement with an earlier suggestion based on thepresence of free-electron-like Zeeman splitting6.In stripe-ordered copper oxides, the strong increase of 1/T2 on

cooling below Tcharge is accompanied by a crossover of the time decayof the spin-echo from the high-temperature Gaussian formexp(2K(t/T2G)2) to an exponential form exp(2t/T2E)18,23. A similarcrossover occurs here, albeit in a less extreme manner because of theabsence ofmagnetic order: 1/T2 sharply increases belowTcharge and thedecay actually becomes a combination of exponential and Gaussiandecays (Fig. 3). In Supplementary Information we provide evidencethat the typical values of the 1/T2E below Tcharge imply that antiferro-magnetic (or ‘spin-density-wave’) fluctuations are slow enough toappear frozen on the timescale of a cyclotron orbit 1/vc< 10212 s.In principle, such slow fluctuations could reconstruct the Fermi sur-face, provided that spins are correlated over large enough distances25,26

(see also ref. 9). It is unclear whether this condition is fulfilled here. The

0.04 0.08 0.12 0.160

40

80

120

Superconducting

Spinorder

T (K

)

p (hole/Cu)

Field-inducedcharge order

Figure 4 | Phase diagram of underdoped YBa2Cu3Oy. The charge orderingtemperature Tcharge (defined as the onset of the Cu2F line splitting; blue opencircles) coincides with T0 (brown plus signs), the temperature at which the Hallconstant RH changes its sign. T0 is considered as the onset of the Fermi surfacereconstruction11–13. The continuous line represents the superconductingtransition temperature Tc. The dashed line indicates the speculative nature ofthe extrapolation of the field-induced charge order. The magnetic transitiontemperatures (Tspin) are frommuon-spin-rotation (mSR) data (green stars)27.T0and Tspin vanish close to the same critical concentration p5 0.08. A scenario offield-induced spin order has been predicted for p. 0.08 (ref. 8) by analogy withLa1.855Sr0.145CuO4, for which the non-magnetic ground state switches toantiferromagnetic order in fields greater than a few teslas (ref. 7 and referencestherein).Ourwork, however, shows that spin order does not occur up to,30T.In contrast, the field-induced charge order reported here raises the question ofwhether a similar field-dependent charge order actually underlies the fielddependence of the spin order in La22xSrxCuO4 and YBa2Cu3O6.45. Error barsrepresent the uncertainty in defining the onset of theNMR line splitting (Fig. 1fand Supplementary Figs 8–10).

0 20 40 60 80 1000

4

8

100

10–1

10–2

1/T 1

(ms–

1 )1/T 2

(�s–

1 )

33.5 T28.5 T

15 T15 T

T (K)

Inte

nsity

(arb

. uni

ts)

15 T

0

0.02

0.04

0.06

15 T

0 50 100 0 50 1001.0

1.5

2.0

33.5 T

28.5 T

T (K)

T (K)

a

c

e

b

d

f

g

Figure 3 | Slow spin fluctuations instead of spin order. a, b, Temperaturedependence of the planar 63Cu spin-lattice relaxation rate 1/T1 for p5 0.108(a) and p5 0.12 (b). The absence of any peak/enhancement on cooling rulesout the occurrence of a magnetic transition. c, d, Increase in the 63Cu spin–spinrelaxation rate 1/T2 on cooling below,Tcharge, obtained from a fit of the spin-echo decay to a stretched form s(t) / exp(2(t/T2)

a), for p5 0.108 (c) andp5 0.12 (d). e, f, Stretching exponent a for p5 0.108 (e) and p5 0.12 (f). Thedeviation from a5 2 on cooling arises mostly from an intrinsic combination ofGaussian and exponential decays, combined with some spatial distribution ofT2 values (Supplementary Information). The grey areas define the crossovertemperature Tslow below which slow spin fluctuations cause 1/T2 to increaseand to become field dependent; note that the change of shape of the spin-echodecay occurs at a slightly higher (,115K) temperature than Tslow. Tslow isslightly lower thanTcharge, which is consistentwith the slow fluctuations being aconsequence of charge-stripe order. The increase of a at the lowesttemperatures probably signifies that the condition cÆhz2æ1/2tc= 1, where tc isthe correlation time, is no longer fulfilled, so that the associated decay is nolonger a pure exponential. We note that the upturn of 1/T2 is already present at15T, whereas no line splitting is detected at this field. The field therefore affectsthe spin fluctuations quantitatively but not qualitatively. g, Plot of NMR signalintensity (corrected for a temperature factor 1/T and for the T2 decay) againsttemperature. Open circles, p5 0.108 (28.5T); filled circles, p5 0.12 (33.5T).The absence of any intensity loss at low temperatures also rules out the presenceof magnetic order with any significant moment. Error bars represent the addeduncertainties in signal analysis, experimental conditions andT2measurements.All measurements are with H | | c.

LETTER RESEARCH

8 S E P T E M B E R 2 0 1 1 | V O L 4 7 7 | N A T U R E | 1 9 3

Macmillan Publishers Limited. All rights reserved©2011

Monday, April 29, 13

Page 98: Quantum criticality in condensed matter

New insights and solvable models for diffusion and transport of strongly interacting systems near quantum critical points using the methods of gauge-gravity duality.

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

Summary

Conformal quantum matter

Monday, April 29, 13

Page 99: Quantum criticality in condensed matter

Antiferromagnetic quantum criticality leads to d-wave superconductivity (supported by sign-problem-free Monte Carlo simulations)

Metals with antiferromagnetic spin correlations have nearly degenerate instabilities: to d-wave superconductivity, and to a charge density wave with a d-wave form factor. This is a promising explanation of the pseudogap regime.

Summary

Antiferromagnetism in metals and the high temperature superconductors

Monday, April 29, 13