Topology, quantum entanglement, and criticality in the...

42
Topology, quantum entanglement, and criticality in the high temperature superconductors Exploring quantum phenomena and quantum matter in ultrahigh magnetic fields, National Science Foundation, Alexandria VA Subir Sachdev September 21, 2017 HARVARD Talk online: sachdev.physics.harvard.edu

Transcript of Topology, quantum entanglement, and criticality in the...

Page 1: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Topology, quantum entanglement, and criticality

in the high temperature superconductors

Exploring quantum phenomena and quantum matter in ultrahigh magnetic fields,

National Science Foundation, Alexandria VA

Subir SachdevSeptember 21, 2017

HARVARD

Talk online: sachdev.physics.harvard.edu

Page 2: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

SM

FL

Figure: K. Fujita and J. C. Seamus Davis

YBa2Cu3O6+x

Page 3: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

“Undoped”insulating

anti-ferromagnet

Page 4: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Anti-ferromagnet

with p mobile holes

per square

Page 5: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

FilledBand

Page 6: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

But relative to the band

insulator, there are 1+ p holes

per square

Anti-ferromagnet

with p mobile holes

per square

Page 7: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Anti-ferromagnet

with p mobile holes

per square

In a conventional metal (a Fermi liquid), with no

broken symmetry, the area enclosed by the Fermi surface must be 1+p But relative to

the band insulator, there are 1+ p holes

per square

Page 8: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

A conventional metal:

the Fermi liquid with Fermi

surface of size 1+p

M. Plate, J. D. F. Mottershead, I. S. Elfimov, D. C. Peets, Ruixing Liang, D. A. Bonn, W. N. Hardy,S. Chiuzbaian, M. Falub, M. Shi, L. Patthey, and A. Damascelli, Phys. Rev. Lett. 95, 077001 (2005)

SM

FL

SM

FL

Page 9: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Figure 1: (a) Schematic phase diagram as a function of doping for hole-doped superconductors.The parent Mott insulating antiferromagnet at zero hole doping evolves to a d-wave super-conducting ground state with increasing hole doping. Superconductivity forms a dome-likeregion, with the superconducting temperature increasing up to an optimal value in the vicinityof p 0.18 hole doping and subsequently decreasing. The region of superconductivity belowoptimal doping is referred to as the underdoped regime, and the region of superconductivityabove optimal doping is referred to as the overdoped regime. A Fermi liquid regime is ob-served beyond the end of the superconducting dome. The elevated temperature region abovethe superconducting dome in the overdoped regime is characterised by unusual properties, suchas an absence of an antinodal density of states at the Fermi energy, and is known as the pseu-dogap regime. (b) Comparison between the small Fermi surface measured in the underdopedregion;26, 35 data from35 and the large Fermi surface measured in the overdoped regime;28, 61 datafrom.28

2

Overdoped

. Vignolle B, Carrington A, Cooper RA, French MMJ, Mackenzie AP, Jaudet C, Vignolles D, Proust C and Hussey NE. 2008. Nature 455:952-955

S. Sebastian and C. Proust, Annual Reviews of Condensed Matter Physics, 6 (2015) 411-30

carried out in a standard 4He cryostat at the LNCMP pulsed fieldfacility in Toulouse using the same contact configuration. For thetorque measurements, small pieces of the same crystals were attachedto sensitive piezoresistive cantilevers and mounted in a dilution refri-gerator in a second magnet cell at LNCMP.

Figure 1a, b shows interlayer resistance (RH) data and magnetictorque (t) data with the field oriented close to the c axis for twodifferent Tl2201 crystals at temperatures below their zero-field super-conducting transitions (Tc < 10 K). For both sets of data, anexpanded view near the field maxima reveals clear oscillations whoseamplitude grows with increasing field strength. Figure 1c shows theoscillatory part of the magnetization Dt/B plotted versus inversemagnetic field 1/B. The observation of oscillations, periodic in 1/B,in both the magnetization and the resistivity, at fields well above theupper critical field Bc2, confirm these as quantum oscillations.Strikingly, as shown in Fig. 1c, these oscillations are more than oneorder of magnitude faster than those found in underdopedYBa2Cu4O8.

The frequency (F) of the oscillations is directly related to theextremal cross-sectional area A of the Fermi surface normal to thefield orientation, via the Onsager relation, A 5 2peF/". Figure 2ashows the fast Fourier transform of the data in Fig. 1c. A single, sharpdHvA frequency of 18,100 6 50 T is obtained, corresponding to aFermi surface extremal cross-sectional area A of 172.8 6 0.5 nm22

and average kF of 7.42 6 0.05 nm21. The temperature dependence of

the Shubnikov–de Haas amplitude is shown in Fig. 2b. Fitting thiswith the standard Lifshitz–Kosevich theory, we obtain a cyclotroneffective mass m*5 4.1 6 1.0 me, where me is the free electron mass.Finally, Fig. 2c shows that the field dependence of the amplitude ofthe dHvA oscillations follows the expected exponential decay, fromwhich we estimate l 5 320 A.

All these numbers are in excellent agreement with those deducedfrom other measurements in the same material with similar dopinglevels. The Fermi surface topology deduced from ADMR6,14 is repro-duced as a solid black line in Fig. 2a inset. Its area agrees extremelywell with the measured dHvA frequency, and corresponds to ,65%of the Brillouin zone of Tl2201 and a doping level p 5 0.30. In addi-tion, given that for a two-dimensional Fermi surface, the electronicspecific heat (Sommerfeld coefficient) is cel 5 (pNAkB

2a2/3"2)m*(where kB is the Boltzmann constant, NA is Avogadro’s number,and a 5 3.86 A is the in-plane lattice constant15), our value of m*corresponds to cel < 6.0 6 1.0 mJ mol21 K22, in excellent agreementwith that measured directly16 for overdoped polycrystalline Tl2201(7 6 2 mJ mol21 K22). Hence, in contrast to the previous work2–5, wecan make quantitative comparisons between quasiparticle propertiesderived from quantum oscillations at high fields and those measureddirectly by transport and thermodynamics at zero field. This goodoverall consistency suggests that the Fermi surface is composedentirely of the single quasi-two-dimensional sheet that we observe.DFT calculations predict that the bare band mass in stoichiometricTl2201 is ,1.2me (ref. 9). The difference between the measured andcalculated masses implies strong electron-correlation-driven renor-malization, even at this elevated doping level.

Despite strong electron–electron interactions, the observation ofquantum oscillations implies that quasiparticles exist at all points onthe Fermi surface of overdoped Tl2201. The observation of genuinequantum oscillations in Tl2201 supports the recognized idea thatgeneralized Fermi-liquid theory can be applied on the overdopedside of the phase diagram. Moreover, the observation of quantumoscillations on both sides of optimal doping, albeit with very differentfrequencies, suggests that the quantum oscillations observed inunderdoped copper oxides directly probe the Fermi surface there,rather than some anomalous vortex physics17, and raises the intri-guing prospect that quasiparticle states may survive, at least at someloci on the Fermi surface, across the entire doping range, from theinsulating/superconducting boundary to the non-superconductingmetal on the heavily overdoped side.

56 58 60

T = 0.7 ± 0.2 K

T = 2.8 ± 0.1 K

54 56 58

5560

0.13

0.14

0.05

0.10

0.15

0.00

–0.10

–0.05

0.00

–0.16

–0.14

600 10 20 30 40 50B (T)

1.7 1.8 1.9100/B (T–1)

2

–0.15

–5

0

5

50

0.05

a

b

R⊥

(Ω)

t (a

.u.)

∆t/B

(a.u

.)

c

YBa2Cu4O8

Tl2201

Figure 1 | Quantum oscillations in Tl2201. a, Raw data on interlayerresistance (RH) with B//c for a Tl2201 single crystal at T 5 2.8 6 0.1 K.RH(B) rises rapidly above the irreversibility field Birr, passes through a smallplateau, then grows quasi-quadratically with field up to 60 T. Inset,magnified view of the high-field region of the down sweep. Small but well-defined oscillations are clearly resolved in the raw data, with an amplitudethat grows with increasing field strength. The maximum amplitude of theoscillations is only 0.5 mV. b, Averaged magnetic torque data (from fivesweeps at temperatures between 0.6 K and 0.8 K) with B close to the c axis fora different Tl2201 crystal. The temperatures of the torque sweeps are subjectto an additional uncertainty of 150 mK due to the weak thermal link insidethe dilution refrigerator5 and the high currents needed to observe theoscillations. Below B 5 14 T, the torque shows hysteretic behaviour due toflux trapping and expulsion in the superconducting mixed state. Again, well-defined oscillations are clearly resolved in the expanded region shown in theinset. The value of Tc for both crystals is 10 K (defined by their zero-resistivestate), compared with the maximal Tc in Tl2201 of 92 K. The torque crystalshowed a very small kink in the zero-field rH data around 20 K, suggestingthat some fraction of the crystal, presumably the surface layer, had a higherTc value. Note that the difference in Birr exhibited by both crystals isamplified in the pulsed magnetic field because the sweep rate dB/dt for thehigher Birr sample is greater. a.u., arbitrary units. c, The oscillatorycomponent of the torque data shown in b plotted as green dots against 1/B(the corresponding B values are shown at the top of the panel), after amonotonic background has been subtracted. The black line superimposedon the data is a fit to the Lifshitz–Kosevich expression for dHvA oscillationsin a two-dimensional metal29. Also shown (blue), for comparison, is theoscillatory component in the torque signal on underdoped YBa2Cu4O8.

NATURE | Vol 455 | 16 October 2008 LETTERS

953 ©2008 Macmillan Publishers Limited. All rights reserved

Page 10: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Figure 1: (a) Schematic phase diagram as a function of doping for hole-doped superconductors.The parent Mott insulating antiferromagnet at zero hole doping evolves to a d-wave super-conducting ground state with increasing hole doping. Superconductivity forms a dome-likeregion, with the superconducting temperature increasing up to an optimal value in the vicinityof p 0.18 hole doping and subsequently decreasing. The region of superconductivity belowoptimal doping is referred to as the underdoped regime, and the region of superconductivityabove optimal doping is referred to as the overdoped regime. A Fermi liquid regime is ob-served beyond the end of the superconducting dome. The elevated temperature region abovethe superconducting dome in the overdoped regime is characterised by unusual properties, suchas an absence of an antinodal density of states at the Fermi energy, and is known as the pseu-dogap regime. (b) Comparison between the small Fermi surface measured in the underdopedregion;26, 35 data from35 and the large Fermi surface measured in the overdoped regime;28, 61 datafrom.28

2

. Vignolle B, Carrington A, Cooper RA, French MMJ, Mackenzie AP, Jaudet C, Vignolles D, Proust C and Hussey NE. 2008. Nature 455:952-955

carried out in a standard 4He cryostat at the LNCMP pulsed fieldfacility in Toulouse using the same contact configuration. For thetorque measurements, small pieces of the same crystals were attachedto sensitive piezoresistive cantilevers and mounted in a dilution refri-gerator in a second magnet cell at LNCMP.

Figure 1a, b shows interlayer resistance (RH) data and magnetictorque (t) data with the field oriented close to the c axis for twodifferent Tl2201 crystals at temperatures below their zero-field super-conducting transitions (Tc < 10 K). For both sets of data, anexpanded view near the field maxima reveals clear oscillations whoseamplitude grows with increasing field strength. Figure 1c shows theoscillatory part of the magnetization Dt/B plotted versus inversemagnetic field 1/B. The observation of oscillations, periodic in 1/B,in both the magnetization and the resistivity, at fields well above theupper critical field Bc2, confirm these as quantum oscillations.Strikingly, as shown in Fig. 1c, these oscillations are more than oneorder of magnitude faster than those found in underdopedYBa2Cu4O8.

The frequency (F) of the oscillations is directly related to theextremal cross-sectional area A of the Fermi surface normal to thefield orientation, via the Onsager relation, A 5 2peF/". Figure 2ashows the fast Fourier transform of the data in Fig. 1c. A single, sharpdHvA frequency of 18,100 6 50 T is obtained, corresponding to aFermi surface extremal cross-sectional area A of 172.8 6 0.5 nm22

and average kF of 7.42 6 0.05 nm21. The temperature dependence of

the Shubnikov–de Haas amplitude is shown in Fig. 2b. Fitting thiswith the standard Lifshitz–Kosevich theory, we obtain a cyclotroneffective mass m*5 4.1 6 1.0 me, where me is the free electron mass.Finally, Fig. 2c shows that the field dependence of the amplitude ofthe dHvA oscillations follows the expected exponential decay, fromwhich we estimate l 5 320 A.

All these numbers are in excellent agreement with those deducedfrom other measurements in the same material with similar dopinglevels. The Fermi surface topology deduced from ADMR6,14 is repro-duced as a solid black line in Fig. 2a inset. Its area agrees extremelywell with the measured dHvA frequency, and corresponds to ,65%of the Brillouin zone of Tl2201 and a doping level p 5 0.30. In addi-tion, given that for a two-dimensional Fermi surface, the electronicspecific heat (Sommerfeld coefficient) is cel 5 (pNAkB

2a2/3"2)m*(where kB is the Boltzmann constant, NA is Avogadro’s number,and a 5 3.86 A is the in-plane lattice constant15), our value of m*corresponds to cel < 6.0 6 1.0 mJ mol21 K22, in excellent agreementwith that measured directly16 for overdoped polycrystalline Tl2201(7 6 2 mJ mol21 K22). Hence, in contrast to the previous work2–5, wecan make quantitative comparisons between quasiparticle propertiesderived from quantum oscillations at high fields and those measureddirectly by transport and thermodynamics at zero field. This goodoverall consistency suggests that the Fermi surface is composedentirely of the single quasi-two-dimensional sheet that we observe.DFT calculations predict that the bare band mass in stoichiometricTl2201 is ,1.2me (ref. 9). The difference between the measured andcalculated masses implies strong electron-correlation-driven renor-malization, even at this elevated doping level.

Despite strong electron–electron interactions, the observation ofquantum oscillations implies that quasiparticles exist at all points onthe Fermi surface of overdoped Tl2201. The observation of genuinequantum oscillations in Tl2201 supports the recognized idea thatgeneralized Fermi-liquid theory can be applied on the overdopedside of the phase diagram. Moreover, the observation of quantumoscillations on both sides of optimal doping, albeit with very differentfrequencies, suggests that the quantum oscillations observed inunderdoped copper oxides directly probe the Fermi surface there,rather than some anomalous vortex physics17, and raises the intri-guing prospect that quasiparticle states may survive, at least at someloci on the Fermi surface, across the entire doping range, from theinsulating/superconducting boundary to the non-superconductingmetal on the heavily overdoped side.

56 58 60

T = 0.7 ± 0.2 K

T = 2.8 ± 0.1 K

54 56 58

5560

0.13

0.14

0.05

0.10

0.15

0.00

–0.10

–0.05

0.00

–0.16

–0.14

600 10 20 30 40 50B (T)

1.7 1.8 1.9100/B (T–1)

2

–0.15

–5

0

5

50

0.05

a

b

R⊥

(Ω)

t (a

.u.)

∆t/B

(a.u

.)

c

YBa2Cu4O8

Tl2201

Figure 1 | Quantum oscillations in Tl2201. a, Raw data on interlayerresistance (RH) with B//c for a Tl2201 single crystal at T 5 2.8 6 0.1 K.RH(B) rises rapidly above the irreversibility field Birr, passes through a smallplateau, then grows quasi-quadratically with field up to 60 T. Inset,magnified view of the high-field region of the down sweep. Small but well-defined oscillations are clearly resolved in the raw data, with an amplitudethat grows with increasing field strength. The maximum amplitude of theoscillations is only 0.5 mV. b, Averaged magnetic torque data (from fivesweeps at temperatures between 0.6 K and 0.8 K) with B close to the c axis fora different Tl2201 crystal. The temperatures of the torque sweeps are subjectto an additional uncertainty of 150 mK due to the weak thermal link insidethe dilution refrigerator5 and the high currents needed to observe theoscillations. Below B 5 14 T, the torque shows hysteretic behaviour due toflux trapping and expulsion in the superconducting mixed state. Again, well-defined oscillations are clearly resolved in the expanded region shown in theinset. The value of Tc for both crystals is 10 K (defined by their zero-resistivestate), compared with the maximal Tc in Tl2201 of 92 K. The torque crystalshowed a very small kink in the zero-field rH data around 20 K, suggestingthat some fraction of the crystal, presumably the surface layer, had a higherTc value. Note that the difference in Birr exhibited by both crystals isamplified in the pulsed magnetic field because the sweep rate dB/dt for thehigher Birr sample is greater. a.u., arbitrary units. c, The oscillatorycomponent of the torque data shown in b plotted as green dots against 1/B(the corresponding B values are shown at the top of the panel), after amonotonic background has been subtracted. The black line superimposedon the data is a fit to the Lifshitz–Kosevich expression for dHvA oscillationsin a two-dimensional metal29. Also shown (blue), for comparison, is theoscillatory component in the torque signal on underdoped YBa2Cu4O8.

NATURE | Vol 455 | 16 October 2008 LETTERS

953 ©2008 Macmillan Publishers Limited. All rights reserved

Underdoped

Overdoped

. Doiron-Leyraud N, Proust C, LeBoeuf D, Levallois J, Bonnemaison JB, Liang R, Bonn DA, Hardy WN, Taillefer L. 2007. Nature 447:565-568

S. Sebastian and C. Proust, Annual Reviews of Condensed Matter Physics, 6 (2015) 411-30

Page 11: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

A conventional metal:

the Fermi liquid with Fermi

surface of size 1+p

SM

FL

SM

FL

carried out in a standard 4He cryostat at the LNCMP pulsed fieldfacility in Toulouse using the same contact configuration. For thetorque measurements, small pieces of the same crystals were attachedto sensitive piezoresistive cantilevers and mounted in a dilution refri-gerator in a second magnet cell at LNCMP.

Figure 1a, b shows interlayer resistance (RH) data and magnetictorque (t) data with the field oriented close to the c axis for twodifferent Tl2201 crystals at temperatures below their zero-field super-conducting transitions (Tc < 10 K). For both sets of data, anexpanded view near the field maxima reveals clear oscillations whoseamplitude grows with increasing field strength. Figure 1c shows theoscillatory part of the magnetization Dt/B plotted versus inversemagnetic field 1/B. The observation of oscillations, periodic in 1/B,in both the magnetization and the resistivity, at fields well above theupper critical field Bc2, confirm these as quantum oscillations.Strikingly, as shown in Fig. 1c, these oscillations are more than oneorder of magnitude faster than those found in underdopedYBa2Cu4O8.

The frequency (F) of the oscillations is directly related to theextremal cross-sectional area A of the Fermi surface normal to thefield orientation, via the Onsager relation, A 5 2peF/". Figure 2ashows the fast Fourier transform of the data in Fig. 1c. A single, sharpdHvA frequency of 18,100 6 50 T is obtained, corresponding to aFermi surface extremal cross-sectional area A of 172.8 6 0.5 nm22

and average kF of 7.42 6 0.05 nm21. The temperature dependence of

the Shubnikov–de Haas amplitude is shown in Fig. 2b. Fitting thiswith the standard Lifshitz–Kosevich theory, we obtain a cyclotroneffective mass m*5 4.1 6 1.0 me, where me is the free electron mass.Finally, Fig. 2c shows that the field dependence of the amplitude ofthe dHvA oscillations follows the expected exponential decay, fromwhich we estimate l 5 320 A.

All these numbers are in excellent agreement with those deducedfrom other measurements in the same material with similar dopinglevels. The Fermi surface topology deduced from ADMR6,14 is repro-duced as a solid black line in Fig. 2a inset. Its area agrees extremelywell with the measured dHvA frequency, and corresponds to ,65%of the Brillouin zone of Tl2201 and a doping level p 5 0.30. In addi-tion, given that for a two-dimensional Fermi surface, the electronicspecific heat (Sommerfeld coefficient) is cel 5 (pNAkB

2a2/3"2)m*(where kB is the Boltzmann constant, NA is Avogadro’s number,and a 5 3.86 A is the in-plane lattice constant15), our value of m*corresponds to cel < 6.0 6 1.0 mJ mol21 K22, in excellent agreementwith that measured directly16 for overdoped polycrystalline Tl2201(7 6 2 mJ mol21 K22). Hence, in contrast to the previous work2–5, wecan make quantitative comparisons between quasiparticle propertiesderived from quantum oscillations at high fields and those measureddirectly by transport and thermodynamics at zero field. This goodoverall consistency suggests that the Fermi surface is composedentirely of the single quasi-two-dimensional sheet that we observe.DFT calculations predict that the bare band mass in stoichiometricTl2201 is ,1.2me (ref. 9). The difference between the measured andcalculated masses implies strong electron-correlation-driven renor-malization, even at this elevated doping level.

Despite strong electron–electron interactions, the observation ofquantum oscillations implies that quasiparticles exist at all points onthe Fermi surface of overdoped Tl2201. The observation of genuinequantum oscillations in Tl2201 supports the recognized idea thatgeneralized Fermi-liquid theory can be applied on the overdopedside of the phase diagram. Moreover, the observation of quantumoscillations on both sides of optimal doping, albeit with very differentfrequencies, suggests that the quantum oscillations observed inunderdoped copper oxides directly probe the Fermi surface there,rather than some anomalous vortex physics17, and raises the intri-guing prospect that quasiparticle states may survive, at least at someloci on the Fermi surface, across the entire doping range, from theinsulating/superconducting boundary to the non-superconductingmetal on the heavily overdoped side.

56 58 60

T = 0.7 ± 0.2 K

T = 2.8 ± 0.1 K

54 56 58

5560

0.13

0.14

0.05

0.10

0.15

0.00

–0.10

–0.05

0.00

–0.16

–0.14

600 10 20 30 40 50B (T)

1.7 1.8 1.9100/B (T–1)

2

–0.15

–5

0

5

50

0.05

a

b

R⊥

(Ω)

t (a

.u.)

∆t/B

(a.u

.)

c

YBa2Cu4O8

Tl2201

Figure 1 | Quantum oscillations in Tl2201. a, Raw data on interlayerresistance (RH) with B//c for a Tl2201 single crystal at T 5 2.8 6 0.1 K.RH(B) rises rapidly above the irreversibility field Birr, passes through a smallplateau, then grows quasi-quadratically with field up to 60 T. Inset,magnified view of the high-field region of the down sweep. Small but well-defined oscillations are clearly resolved in the raw data, with an amplitudethat grows with increasing field strength. The maximum amplitude of theoscillations is only 0.5 mV. b, Averaged magnetic torque data (from fivesweeps at temperatures between 0.6 K and 0.8 K) with B close to the c axis fora different Tl2201 crystal. The temperatures of the torque sweeps are subjectto an additional uncertainty of 150 mK due to the weak thermal link insidethe dilution refrigerator5 and the high currents needed to observe theoscillations. Below B 5 14 T, the torque shows hysteretic behaviour due toflux trapping and expulsion in the superconducting mixed state. Again, well-defined oscillations are clearly resolved in the expanded region shown in theinset. The value of Tc for both crystals is 10 K (defined by their zero-resistivestate), compared with the maximal Tc in Tl2201 of 92 K. The torque crystalshowed a very small kink in the zero-field rH data around 20 K, suggestingthat some fraction of the crystal, presumably the surface layer, had a higherTc value. Note that the difference in Birr exhibited by both crystals isamplified in the pulsed magnetic field because the sweep rate dB/dt for thehigher Birr sample is greater. a.u., arbitrary units. c, The oscillatorycomponent of the torque data shown in b plotted as green dots against 1/B(the corresponding B values are shown at the top of the panel), after amonotonic background has been subtracted. The black line superimposedon the data is a fit to the Lifshitz–Kosevich expression for dHvA oscillationsin a two-dimensional metal29. Also shown (blue), for comparison, is theoscillatory component in the torque signal on underdoped YBa2Cu4O8.

NATURE | Vol 455 | 16 October 2008 LETTERS

953 ©2008 Macmillan Publishers Limited. All rights reserved

B. Vignolle, A. Carrington, R.A. Cooper, M.M.J. French, A.P. Mackenzie,C. Jaudet, D. Vignolles, C. Proust and N.E. Hussey, Nature 455, 952 (2008)

Page 12: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Quantum oscillations of a Fermi surface of size close to p

SM

FL

SM

FL

of inverse field, by subtracting the monotonic background (shown forall temperatures in Supplementary Fig. 2). This shows that the oscilla-tions are periodic in 1/B, as is expected of oscillations that arisefrom Landau quantization. A Fourier transform yields the powerspectrum, displayed in Fig. 3b, which consists of a single frequency,F 5 (530 6 20) T. In Fig. 3c, we plot the amplitude of the oscillationsas a function of temperature, from which we deduce a carrier massm* 5 (1.9 6 0.1)m0, where m0 is the bare electron mass. Within errorbars, both F and m* are the same in sample B, for which the current J isparallel to the b axis (see Supplementary Fig. 1). Oscillations of thesame frequency are also observed in Rxx (in both samples), albeit with asmaller amplitude. We note that while at 7.5 K the oscillations are stillperceptible, they are absent at 11 K, as expected from thermallydamped quantum oscillations (see Supplementary Fig. 5).

While quantum oscillations in YBa2Cu3O61y (YBCO) have beenthe subject of a number of earlier studies8–10, the data reported so fardo not exhibit clear oscillations as a function of 1/B and, as such, havenot been accepted as convincing evidence for a Fermi surface11.Furthermore, we note that all previous work was done on orientedpowder samples as opposed to the high-quality single crystals used inthe present study.

Quantum oscillations are a direct measure of the Fermi surfacearea via the Onsager relation: F 5 (W0/2p2)Ak, where W0 5 (2.07 310215) T m2 is the flux quantum, and Ak is the cross-sectional area ofthe Fermi surface normal to the applied field. A frequency of 530 Timplies a Fermi surface pocket that encloses a k-space area (in the a–bplane) of Ak 5 5.1 nm22, that is, 1.9% of the Brillouin zone (of area4p2/ab). This is only 3% of the area of the Fermi surface cylindermeasured in Tl-2201 (see Fig. 1c), whose radius is kF < 7 nm21. In theremainder, we examine two scenarios to explain the dramatic differ-ence between the small Fermi surface revealed by the low frequency ofquantum oscillations reported here for YBa2Cu3O6.5 and the largecylindrical surface observed in overdoped Tl-2201. The first scenarioassumes that the particular band structure of YBa2Cu3O6.5 is differ-ent and supports a small Fermi surface sheet. In the second, theelectronic structure of overdoped copper oxides undergoes a trans-formation as the doping p is reduced below a value pc associated witha critical point.

Band structure calculations for stoichiometric YBCO (y 5 1.0),

which is slightly overdoped (with p 5 0.2), show a Fermi surfaceconsisting of four sheets12,13, as reproduced in Fig. 4a: two largecylinders derived from the CuO2 bi-layer, one open surface comingfrom the CuO chains, and a small cylinder associated with both chainand plane states. The latter sheet, for example, could account for thelow frequency reported here. ARPES studies on YBCO near optimal

doping14,15 appear to be in broad agreement with this electronicstructure. However, recent band structure calculations16 performedspecifically for YBa2Cu3O6.5, which take into account the unit celldoubling caused by the ortho-II order, give a Fermi surface where thesmall cylinder is absent, as shown in Fig. 4b. This leaves no obviouscandidate Fermi surface sheet for the small orbit reported here.

The fact that the same oscillations are observed for currents along aand b suggests that they are not associated with open orbits in thechain-derived Fermi surface sheet. In YBCO, the CuO chains alongthe b axis are an additional channel of conduction, responsible for ananisotropy in the zero-field resistivity r(T) of the normal state (above

35 40 45 50 55 60

45 50 55B (T)

60 65

J II a

B

0.04

0.05

0.06

1.5 K 2 K 2.5 K 3 K 3.5 K 4.2 K

0.02

0.04

30 65B (T)

–Rxy

(Ω)

–Rxy

(Ω)

0.00

0.06

Figure 2 | Hall resistance of YBa2Cu3O6.5. Rxy as a function of magneticfield B, for sample A, at different temperatures between 1.5 and 4.2 K. Thefield is applied normal to the CuO2 planes (B | | c) and the current is along thea axis of the orthorhombic crystal structure (J | | a). The inset shows a zoomon the data at T 5 2 K, with a fitted monotonic background (dashed line).

0.015 0.020 0.025

1 / B (T–1)

1 / (530 T)

–6

–4

–2

0

2

–∆R

xy (m

Ω)

a 4

0 1 2 3F (kT)

0.0

0.5

1.0

1 3 52 4T (K)

–9

–8

1.5 –7

ln (A

/ T)

c

A (a

.u.)

b

Figure 3 | Quantum oscillations in YBCO. a, Oscillatory part of the Hallresistance, obtained by subtracting the monotonic background (shown inthe inset of Fig. 2 for T 5 2 K), as a function of inverse magnetic field, 1/B.The background at each temperature is given in Supplementary Fig. 2.b, Power spectrum (Fourier transform) of the oscillatory part for the T 5 2 Kisotherm, revealing a single frequency at F 5 (530 6 20) T, whichcorresponds to a k-space area Ak 5 5.1 nm22, from the Onsager relationF 5 (W0/2p2)Ak . Note that the uncertainty of 4% on F is not given by thewidth of the peak (a consequence of the small number of oscillations), but bythe accuracy with which the position of successive maxima in a can bedetermined. c, Temperature dependence of the oscillation amplitude A,plotted as ln(A/T) versus T. The fit is to the standard Lifshitz–Kosevichformula, whereby A/T 5 [sinh(am*T/B)]21, which yields a cyclotron massm* 5 (1.9 6 0.1)m0, where m0 is the free electron mass.

Y

ba

Y

X XΓΓ

Figure 4 | Fermi surface of YBCO from band structure calculations.a, Fermi surface of YBa2Cu3O7 in the kz 5 0 plane (from ref. 13, withpermission from O. K. Andersen), showing the four bands discussed in themain text. b, Fermi surface of ortho-II ordered YBa2Cu3O6.5 in the kz 5 0plane (from ref. 16, with permission from T. M. Rice). In both a and b thegrey shading indicates one quadrant of the first Brillouin zone.

LETTERS NATURE | Vol 447 | 31 May 2007

566Nature ©2007 Publishing Group

N. Doiron-Leyraud, C. Proust, D. LeBoeuf, J. Levallois, J.B. Bonnemaison,R. Liang, D.A. Bonn, W.N. Hardy, L. Taillefer, Nature 447, 565 (2007)

Page 13: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Figure 10: Charge order with wavevectors Qx and Qy of similar or dissimilar amplitudes thatyields a nodal electron pocket (red). A negative Hall effect is expected from such a pocket athigh magnetic fields. The dotted green lines represent antinodal regions, which are suppresseddue likely to a combination of charge correlations in conjunction with effects such as pairingor magnetic correlations.40, 86 Figure adapted from References.40, 86 (b) Quantum mechanicalcalculation of an electron spectral function. Charge order with d-wave symmetry (i.e., bondorder) with wavevectors Qx and Qy gives similar results to those found in panel a, yieldinga nodal electron pocket (red) accompanied by small adjoining hole pockets (blue).87 Figureadapted from Reference.87

17

Harrison N, Sebastian SE. 2011. Phys. Rev. Lett. 106:226402

Great deal of evidence that underdoped quantum oscillations are due to an electron pocket which is created

by field-induced long-range charge density wave order

Page 14: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Reconstructing the large Fermi surface by CDW order inevitably leads to many more Fermi surfaces, apart from

the electron pocket

Allais A, Chowdhury D, Sachdev S. 2014. Nature Commun. 5:5771

2

Our results are in good agreement with some of the observedtrends.

Q2

Q1

-p 0 p-p

0

paL

-p 0 p

bL

0

1

50 100 150 200 250 3000.0

0.5

1.0

1.5

2.0

15203050100

—êeBa02

d.o.

s.

BHTL

FIG. 1. Spectral functions in the presence of static and fluctuatingorder. (a) The color density plot displays the electron spectral func-tion in the presence of long-range bidirectional bond density wave(BDW) at zero magnetic field (in the unfolded Brillouin zone). Suchlong-ranged BDW is likely to be present only in a strong magneticfield and will not be seen in ARPES experiments. Annotations aresuperimposed to highlight aspects of the spectral density. In black,the Fermi surface used for our computation. Dashed arrows markthe wavevectors of the BDW. The BDW causes reconstruction of theFermi surface and the formation of an electron-like pocket, markedin red, and two hole-like pockets, marked in blue. The pocket con-tours are obtained by semiclassical analysis as described in sec. III A.The parameters are t1 = 1.0, t2 = 0.33, t3 = 0.03, µ = 0.9604,p = 10%, Px

0 = Py0 = 0.15, = 0.317. (b) Electron spectral func-

tion in the presence of fluctuating superconducting and bond densitywave correlations. The parameters are p = 11%, 0 = Px

0 = Py0 = 1,

T/t1 = 0.06, g/2 = 0.2, S = 0.05, = 2. The details are dis-cussed in sec. I C (c) Quantum oscillations in the density of statesinduced by an applied magnetic field: red lines mark peaks associ-ated with the electron pocket (frequency 432 T or 1.55% of Brillouinzone), and blue lines those from the hole pockets (frequency 90.9 Tor 0.326% of Brillouin zone).

I. RESULTS

A. Model

We base our analysis on the following model hamiltonian

H =X

r,a

tac†

r+a

cr

+ a

r+a/2c†

r+a,"c†r,# + h.c.

+X

i

Pia

eiQi·(r+a/2)ir+a/2c†

r+a

cr

+ h.c..

(1)

Here r labels the sites of a square lattice and the vector a

runs over first, second and third neighbors, and also on-site(a = 0). The first term is the usual kinetic term, with hoppingparameters ta.

The second term couples the electron to the superconduct-ing order parameter. The coecient

a

specifies the super-conducting form factor and the field is the superconductingorder parameter: it can be short-ranged or acquire an expecta-tion value.

The third term couples the fermion to the bond order. Theindex i labels di↵erent wavevectors Qi, the coecients Pi

a

specify the corresponding form factors and the fields i arethe order parameters, which can also be long-ranged or fluc-tuating. Throughout this work we consider a specific formof a density wave which resides primarily upon the bonds ofthe lattice: this is not crucial for the quantum oscillations, butis important for the electron spectral function at intermediatetemperatures. Building on recent experimental and theoreti-cal work [34–48] we use a bond density wave (BDW) with ad-form factor.

Both interaction terms can be obtained by appropriate de-coupling of the Heisenberg interaction in the particle-hole andparticle-particle channels [35].

We use a d-wave superconducting form factor ±x

=+0/2,±y

= 0/2 and a bond order with the same formfactor Pi

±x

= +Pi0/2, P

i±y

= Pi0/2 [34, 35] which is supported

by recent experimental evidence [47, 48] although a small s-wave component may also be present [49]. We consider a setof two wavevectors Q1 = 2(, 0) and Q2 = 2(0, ), with 0.3, also based on experimental evidence.

A summary of our main results appears in fig.1, whichshows the electronic spectral functions in the presence oflong-range incommensurate BDW (fig.1a) and in the presenceof fluctuating BDW and superconductivity (fig.1b).

B. Quantum oscillations from density wave order

Fig. 1(a) illustrates Fermi surface reconstruction by a bi-directional density wave modulation with wavevectors paral-lel to the Cu-O bonds and period 3 lattice spacings. Weconstrain the Fermi surface by spectroscopy experiments [50],and use the remaining freedom in hopping parameters to ob-tain an electron pocket of about the right size. While our anal-ysis mainly centers around the experiments on YBCO, whichhas a sligthly orthorombic lattice, for simplicity we assume

Page 15: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Zerofield“gamma”

0 10 20 30 40 50 60 70 80 90 1000

10

20

30

40

50

C/T

(mJ

mol

-1 K

-2)

Temperature2 (K2)

YBCO 6.56

0T

γ=1.85

mJ/mol K2

LowestvaluesforYBCO~2mJ/molK2(Wrightetal,HandbookofHighTc)

LSCOvaluesrangefrom0.4to4mJ/molK2.WenetalPhysicalReviewB70,214505(2004)

Specific heat measurements in field-induced normal state of under doped YBCO

Value is only consistent with a single electron pocket

. Riggs SC, Vafek O, Kemper JB, Betts JB, Migliori A, Balakirev FF, Hardy WN, Liang R, Bonn DA, Boebinger GS. 2011. Nature Phys. 7:332-335

Page 16: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Sawtooth waveform characteristic of an ideal 2D Fermi gas

2D electrons in GaAs/AlGaAs heterostructures

A. Potts et al. J. Phys.: Condens. Matter 8, 5189–5207 (1996) Y. –T. Hsu, M. Hartstein, J. Porras, T. Loew et al. (unpublished)

Consistent with a single electron pocket

Page 17: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

arX

iv:1

606.

0277

2v1

[con

d-m

at.su

pr-c

on]

8 Ju

n 20

16

Single reconstructed Fermi surface pocket in an underdoped single layer cupratesuperconductor

M. K. Chan,1, 2, ∗ N. Harrison,1, ∗ R. D. McDonald,1 B. J. Ramshaw,1 K. A. Modic,1 N. Barisic,3,2 and M. Greven2

1Mail Stop E536, Pulsed Field Facility, National High Magnetic Field Laboratory,Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA

2School of Physics and Astronomy, University of Minnesota, Minneapolis, Minnesota 55455, USA3Technishce Universitat Wein, Wiedner Haupstr. 8-10, 1040, Vienna, Austria

The observation of a reconstructed Fermi surface via quantum oscillations in hole-doped cupratesopened a path towards identifying broken symmetry states in the pseudogap regime. However, suchan identification has remained inconclusive due to the multi-frequency quantum oscillation spectraand complications accounting for bilayer effects in most studies. We overcome these impedimentswith high resolution measurements on the structurally simpler cuprate HgBa2CuO4+δ (Hg1201),which features one CuO2 plane per unit cell. We find only a single oscillatory component withno signatures of magnetic breakdown tunneling to additional orbits. Therefore, the Fermi surfacecomprises a single quasi-two-dimensional pocket. Quantitative modeling of these results indicatesthat biaxial charge-density-wave within each CuO2 plane is responsible for the reconstruction, andrules out criss-crossed charge stripes between layers as a viable alternative in Hg1201. Lastly, wedetermine that the characteristic gap between reconstructed pockets is a significant fraction of thepseudogap energy.

I. INTRODUCTION

The identification of broken symmetry states, particu-larly in the pseudogap region, is essential for understand-ing the cuprate phase diagram. The surprising discoveryof a small Fermi surface from quantum oscillations (QOs)in underdoped YBa2Cu3O6+x (Y123) [1] motivated pro-posals for a crystal-lattice-symmetry-breaking order pa-rameter [2–9] that reconstructs either the large Fermisurface identified in overdoped cuprates [10–12] or theFermi arcs of the pseudogap state [13–15].The spectrum of quantum oscillations is, in principle,

a distinct probe of the Fermi surface morphology andis thus a signature of the broken symmetry state [2–9].The ubiquity of short-range charge-density-wave (CDW)in underdoped cuprates [16–22] make CDW a naturalchoice as the order responsible for Fermi surface recon-struction. However, despite the availability of exquisitelydetailed QO studies in Y123 [8, 23–25], its complicatedmulti-frequency spectrum has prevented a consensus onthe exact model for reconstruction [8, 9, 24–28]. Part ofthe difficulty stems from the crystal structure of Y123,particularly the bilayer splitting of the elementary Fermipockets due to the two CuO2 planes per unit cell. Thedifferent models are sensitive to the magnitude, sym-metry, and momentum dependence of the bilayer cou-pling [8, 26–28], which are controversial. Furthermore,neither diffraction nor QO experiments in the cuprateshave yet been able to address the crucial question as towhether the two orthogonal CDW vectors spatially coex-ist in the same CuO2 plane or whether stripes alternate ina criss-cross fashion on consecutive CuO2 planes [29, 30].

∗ Correspondence to [email protected] and [email protected]

Apart from the Y-based bilayer compounds [1, 31],HgBa2CuO4+δ (Hg1201) is the only other hole-dopedcuprate for which QOs have been detected [32] in thepseudogap regime. Importantly, in addition to featur-ing a very high-Tc (≈ 97 K at optimal doping), Hg1201has a tetragonal crystal symmetry consisting of only oneCuO2 plane per unit cell. This means that the analy-sis of the experimental data on this compound is freefrom complications associated with bilayer coupling andorthorombicity.

Here we show that high resolution measurements ofup to 10 cycles of the QOs in Hg1201 permit a reso-lution of the reconstructed electronic structure. Usingpulsed magnetic fields extending to 90 T combined withcontactless resistivity measurements, we find the QOsin Hg1201 to be remarkably simple: a single oscilla-tion frequency exhibiting a monotonic magnetic field de-pendence characteristic of a single Fermi surface pocket.We find quantitative agreement between the observedsingle QO frequency and that from a diamond-shapedelectron pocket resulting from biaxial CDW reconstruc-tion [18, 33]. There are no signatures of the predicted ad-ditional small hole-like pocket [9] reported for Y123 [25].This could be due to the antinodal states, which con-stitute these hole pockets, being gapped out or stronglysupressed by the pseudogap phenomena. We also de-termine a very small c-axis transfer integral for Hg1201,which precludes a model based on an alternating criss-cross pattern of uniaxial charge stripes on consecutiveCuO2 planes [30]. The absence of signatures of magneticbreakdown tunneling to neighboring sections of the Fermisurfaces (such as the putative small hole pockets [9]) pro-vides a lower bound estimate of≈ 20 meV for the relevantgap. Importantly, this is a significant fraction of the anti-nodal pseudogap energy [34]. Overall, our results point

Nature Communications 7, 12244 (2016)

3

0 20 40 60 800

0.1

0.2

0.3

∆f(

MH

z)

HgBa2CuO4+δa

SC Normal

50 60 70 80 900

500

1000

1500

d∆f/dB

(Hz

T-1

)

b

50 60 70 80 90

0.1

0

0.1

B (T)

∆f/f

(%)

c

FIG. 2. Observation of quantum oscillations inHg1201 with contactless resistivity. (a) Evolution ofthe PDO circuit frequency coupled to Hg1201 UD71 withapplied magnetic field B along the c axis of the sample atT = 1.8 K. The sample undergoes a transition from super-conducting (SC, black shaded region) to normal (blue region)at B ∼ 35 T. (b) Derivative of the raw data with respectto magnetic field reveals quantum oscillations in the normalstate. As described in the text, a non-oscillatory polynomialbackground is subtracted from the raw data to extract thequantum oscillations. The derivative of the background isshown as the dashed black line. (c) Quantum oscillations afterthe polynomial background has been removed. The dashedblack line is a fit to the Lifshitz-Kosevitch form discussed inMethods.

Figs. 3). In contrast to the residuals obtained for Hg1201,the residual for Y123 (see Fig. 3b), again obtained onsubtracting a fit to the dominant quantum oscillation fre-quency (of F ≈ 530 T), reveals a distinctive beat patternresulting from the interference between the two remain-ing QO components whose amplitudes are ≈ 40 and 50 %of the dominant frequency (FT of the residual shown

Fig. 3c).Limits on the Fermi surface warping and c-axis

hopping. In quasi-two-dimensional metals, the inter-plane hopping leads to warping of the cylindrical Fermisurface, yielding two oscillation frequencies originatingfrom minimum and maximum extremal cross-sections.While our observation of a single quantum oscillationfrequency rules out very large warping, small warpingcan manifest in observable nodes in the magnetic field-dependent QO amplitude. This is represented by an ad-ditional amplitude factor, Rw, which is parametrized bythe separation between the two frequencies 2∆Fc (seeMethods). To illustrate this point, in Fig. 4 we fit thedata with several different fixed values of ∆Fc in Rw.The absence of nodes in the experimental data enablesus to make an upper bound estimate of ∆Fc < 16 T.Using m∗ ≈ 2.7 me (see Fig.5) and 2∆Fc ≈ 4t⊥m∗/(!e),we obtain a c-axis hopping of t⊥ < 0.35 meV for Hg1201,revealing it to be at least 1000 times smaller than thenearest neighbor hopping (t = 460 meV [45]) within theCuO2 planes. Our upper bound is also 25 times smallerthan the bare value determined from LDA calculations(t⊥ = 10 meV [45]), reflecting a large quasiparticle re-normalization.Our ability to set a firm upper bound estimate for t⊥

in Hg1201 contrasts with the situation in Y123, where es-timates of the c-axis warping are challenging to separatefrom the effects of bilayer coupling. Estimates for ∆Fc

in Y123 range from ≈ 15 to 90 T depending on whetherthe observed beat pattern originates from the combinedeffects of bilayer-splitting and magnetic breakdown tun-neling [8] or Fermi surface warping [23–25].Fermi surface reconstruction by biaxial CDW.

The simple crystalline structure of Hg1201 makes it theideal system for relating the k-space area of the observedFermi surface pocket to prior photoemission [46] and x-ray scattering [18] measurements. The former constrainsthe unreconstructed Fermi-surface while the later pro-vides the magnitude of the reconstruction wave-vector.Following Allais et al. [9], we require that the large un-reconstructed hole-like Fermi surface of area AUFS ac-commodate 1 + p carriers, where p is the hole dopingdefined relative to the half filled band. We then pro-ceed to translate the Fermi surface multiple times by thewavevectors (QCDW, 0) and (0, QCDW) and their combi-nations in Fig. 6a. Here we have assumed a biaxial re-construction scheme. Further details of the calculation isdescribed in the Methods section.The biaxial CDW reconstruction (shown in Fig. 6)

yields a diamond-shaped electron pocket (depicted inred) flanked by smaller hole pockets (depicted in blue)accompanied by additional open Fermi surface sheets(shown in Supplementary Figure 3). While the parame-ters for Hg1201 are slightly different than for Y123, thetopology of the reconstructed Fermi surface is essentiallythe same.Using the Onsager relation Fe,h = !

2πeAe,h, where Ae

is the area of the electron pocket and Ah is the area of

And shape of quantum oscillations also support only a single electron pocket

Page 18: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

The parent “pseudogap metal”, without CDW order, cannot have a large Fermi surface, and should be gapped in the anti-nodal region

Such a Fermi surface violates the Luttinger theorem of Fermi-liquid theory

However, such a Fermi surface can appear in a metal with bulk topological order, which has long-range, many-body quantum entanglement

Page 19: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

The parent “pseudogap metal”, without CDW order, cannot have a large Fermi surface, and should be gapped in the anti-nodal region

Such a Fermi surface violates the Luttinger theorem of Fermi-liquid theory

However, such a Fermi surface can appear in a metal with bulk topological order, which has long-range, many-body quantum entanglement

Page 20: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

The parent “pseudogap metal”, without CDW order, cannot have a large Fermi surface, and should be gapped in the anti-nodal region

Such a Fermi surface violates the Luttinger theorem of Fermi-liquid theory

However, such a Fermi surface can appear in a metal with bulk topological order, which has long-range, many-body quantum entanglement

T. Senthil, M. Vojta and S. Sachdev, PRB 69, 035111 (2004)

A. Paramekanti and A. Vishwanath, PRB 70, 245118 (2004).

Page 21: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

A simple model:

Fluctuating antiferromagnetism leads

to small Fermi surfaces and

bulk topological order with

long-range quantum entanglement

Page 22: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Begin with the “spin-fermion” model. Electrons ci↵ on the square

lattice with dispersion

Hc = X

i,

t

c

†i,↵ci+v,↵

+ c

†i+v,↵

ci,↵

µ

X

i

c

†i,↵ci,↵ +Hint

are coupled to an antiferromagnetic order parameter

`(i),

` = x, y, z

Hint =

X

i

i`(i)c

†i,↵

`↵ci, + V

where i = ±1 on the two sublattices.

When

`(i) =constant independent of i, we have long-range AFM,

and a gap in the fermion spectrum at the anti-nodes.

Page 23: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Begin with the “spin-fermion” model. Electrons ci↵ on the square

lattice with dispersion

Hc = X

i,

t

c

†i,↵ci+v,↵

+ c

†i+v,↵

ci,↵

µ

X

i

c

†i,↵ci,↵ +Hint

are coupled to an antiferromagnetic order parameter

`(i),

` = x, y, z

Hint =

X

i

i`(i)c

†i,↵

`↵ci, + V

where i = ±1 on the two sublattices.

When

`(i) =constant independent of i, we have long-range AFM,

and a gap in the fermion spectrum at the anti-nodes.

Increasing SDW order

Page 24: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Begin with the “spin-fermion” model. Electrons ci↵ on the square

lattice with dispersion

Hc = X

i,

t

c

†i,↵ci+v,↵

+ c

†i+v,↵

ci,↵

µ

X

i

c

†i,↵ci,↵ +Hint

are coupled to an antiferromagnetic order parameter

`(i),

` = x, y, z

Hint =

X

i

i`(i)c

†i,↵

`↵ci, + V

where i = ±1 on the two sublattices.

When

`(i) =constant independent of i, we have long-range AFM,

and a gap in the fermion spectrum at the anti-nodes.

Increasing SDW order

Increasing SDW order

Page 25: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

For fluctuating antiferromagnetism, we transform to arotating reference frame using the SU(2) rotation Ri

ci"ci#

= Ri

i,+

i,

,

in terms of fermionic “chargons” s and a Higgs field Ha(i)

``(i) = Ri aHa(i)R†

i

The Higgs field is the AFM order in the rotating reference frame.Note that this representation is ambiguous up to aSU(2) gauge transformation, Vi

i,+

i,

! Vi

i,+

i,

Ri ! RiV†i

aHa(i) ! Vi bHb(i)V †

i .

Page 26: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Fluctuating antiferromagnetism

The simplest e↵ective Hamiltonian for the fermionic chargons isthe same as that for the electrons, with the AFM order replacedby the Higgs field.

H = X

i,

t

†i,s i+v,s

+ †i+v,s

i,s

µ

X

i

†i,s i,s +Hint

Hint = X

i

iHa(i) †

i,sass0 i,s0 + VH

IF we can transform to a rotating reference frame in whichHa(i) =a constant independent of i and time, THEN the fermions inthe presence of fluctuating AFM will inherit the anti-nodal gap ofthe electrons in the presence of static AFM.

Page 27: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Fluctuating antiferromagnetism

The simplest e↵ective Hamiltonian for the fermionic chargons isthe same as that for the electrons, with the AFM order replacedby the Higgs field.

H = X

i,

t

†i,s i+v,s

+ †i+v,s

i,s

µ

X

i

†i,s i,s +Hint

Hint = X

i

iHa(i) †

i,sass0 i,s0 + VH

IF we can transform to a rotating reference frame in whichHa(i) =a constant independent of i and time, THEN the fermions inthe presence of fluctuating AFM will inherit the anti-nodal gap ofthe electrons in the presence of static AFM.Increasing SDW order

Page 28: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Fluctuating antiferromagnetismWe cannot always find a single-valued SU(2) rotation Ri to makethe Higgs field Ha(i) a constant !

n-fold vortex in

AFM order

A.V. Chubukov, T. Senthil and S. Sachdev, PRL 72, 2089 (1994);

S. Sachdev, E. Berg, S. Chatterjee, and Y. Schattner, PRB 94, 115147 (2016)

Page 29: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Fluctuating antiferromagnetismWe cannot always find a single-valued SU(2) rotation Ri to makethe Higgs field Ha(i) a constant !

n-fold vortex in

AFM order

R

(1)nR

A.V. Chubukov, T. Senthil and S. Sachdev, PRL 72, 2089 (1994);

S. Sachdev, E. Berg, S. Chatterjee, and Y. Schattner, PRB 94, 115147 (2016)

Page 30: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

We cannot always find a single-valued SU(2) rotation Ri to makethe Higgs field Ha(i) a constant !

n-fold vortex in

AFM order

R

(1)nR

Topological order

Vortices with n odd must be suppressed: such a metal with“fluctuating antiferromagnetism” has BULK Z2

TOPOLOGICAL ORDER and fermions which inherit the“pocket” Fermi surfaces of the antiferromagnetic metal i.e. apseudogap. Odd vortices in antiferromagnetism are stable

bulk quasiparticles (‘visons’) which could be observed in STM.

Page 31: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

SM

FL

Pseudogap metal

with bulk topological

order?

Page 32: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

SM

FL

Gauge theory for a topological Higgs-to-

confinement phase

transition

Page 33: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

SM

FL

Gauge theory for a topological Higgs-to-

confinement phase

transition

Page 34: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

increased hole doping, with fits to Eq. 1. (B) shows the same data versus 𝑘𝑘𝐵𝐵𝑇𝑇/ℏ𝜔𝜔𝑐𝑐∗, where

𝜔𝜔𝑐𝑐∗ = 𝑒𝑒𝑒𝑒/𝑚𝑚∗: this scaling with 𝑚𝑚∗ shows the robustness of the fit across the entire doping and

temperature range. (C) The effective mass as a function of hole doping; error bars are the

standard error from regression of Eq. 1 to the data. The dashed line is a guide to the eye.

Fig. 4: A quantum critical point near optimal doping. The blue curves correspond to 𝑇𝑇𝑐𝑐, as

defined by the resistive transition (right axis), at magnetic fields of 0, 15, 30, 50, 70, and 82 T

(some data points taken from [39], [57].) As the magnetic field is increased, the

superconducting 𝑇𝑇𝑐𝑐 is suppressed. By 30 T two separate domes remain, centred around

𝑝𝑝 ≈ 0.08 and 𝑝𝑝 ≈ 0.18; by 82 T only the dome at 𝑝𝑝 ≈ 0.18 remains. The inverse of the

Quantum oscillations and DW order

Phase diagram in a high magnetic field

B. J. Ramshaw, S. E. Sebastian, R. D. McDonald, James Day, B. S. Tan, Z. Zhu, J. B. Betts, Ruixing Liang, D. A. Bonn, W. N. Hardy, N. Harrison, Science 348, 6232 (2105).

Page 35: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

9

0.05 0.1 0.15 0.2 0.25p

0

20

40

60

80T

( K )

YBCO H = 50 T

Tc

CDW

TNMR

SDW

p*

0 0.1 0.2 0.3p

0

100

200

T ( K

)

YBCOH = 0

TN

Tc

TXRD

TSDW

T*

b

a

Phase diagram in a high magnetic field

Badoux, Proust, Taillefer et al., Nature 531, 210 (2016)

Quantum oscillations and DW order

Page 36: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Hall effect measurements in YBCO

Badoux, Proust, Taillefer et al., Nature 531, 210 (2016)

14

0 0.1 0.2 0.3p

0

0.5

1

1.5n H

= V

/ e

RH

p

1 + p

SDW CDW FL

p*

a

b

Fermi liquid (FL) with carrier

density 1+p

Page 37: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Hall effect measurements in YBCO

Badoux, Proust, Taillefer et al., Nature 531, 210 (2016)

14

0 0.1 0.2 0.3p

0

0.5

1

1.5n H

= V

/ e

RH

p

1 + p

SDW CDW FL

p*

a

b

Spin density wave (SDW)

breaks translational invariance,

and the Fermi liquid then has carrier density p

Page 38: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Hall effect measurements in YBCO

Badoux, Proust, Taillefer et al., Nature 531, 210 (2016)

14

0 0.1 0.2 0.3p

0

0.5

1

1.5n H

= V

/ e

RH

p

1 + p

SDW CDW FL

p*

a

b

Charge density wave (CDW)

leads to complex Fermi

surface reconstruction and negative

Hall resistance

Page 39: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Hall effect measurements in YBCO

Badoux, Proust, Taillefer et al., Nature 531, 210 (2016)

14

0 0.1 0.2 0.3p

0

0.5

1

1.5n H

= V

/ e

RH

p

1 + p

SDW CDW FL

p*

a

b

Rapid increase in carrier

density due to a quantum

phase transition at p=0.19 ?

Page 40: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

Scale-invariant magnetoresistance in a cuprate superconductor

P. Giraldo-Gallo,1 J. A. Galvis,1 Z. Stegen,1 K. A. Modic,2 F. F Balakirev,3 J. B. Betts,3

X. Lian,1 C. Moir,1 S. C. Riggs,1 J. Wu,4 A. T. Bollinger,4 X. He,4, 5 I. Bozovic,4, 5

B. J. Ramshaw,6, 3 R. D. McDonald,3 G. S. Boebinger,1, 7 and A. Shekhter1,

1National High Magnetic Field Laboratory,

Florida State University, Tallahassee, FL 32310, USA

2Max-Planck-Institute for Chemical Physics of Solids,

Noethnitzer Strasse 40, D-01187, Dresden, Germany

3Los Alamos National Laboratory, Los Alamos, NM 87545, USA

4Brookhaven National Laboratory, Upton, New York 11973-5000, USA.

5Applied Physics Department, Yale University,

New Haven, Connecticut 06520, USA.

6Laboratory for Atomic and Solid State Physics,

Cornell University, Ithaca, NY 14853, USA

7Department of Physics, Florida State University, Tallahassee, FL 32310, USA

The anomalous metallic state in high-temperature superconducting cuprates is masked

by the onset of superconductivity near a quantum critical point. Use of high magnetic

fields to suppress superconductivity has enabled a detailed study of the ground state in

these systems. Yet, the direct e↵ect of strong magnetic fields on the metallic behavior at

low temperatures is poorly understood, especially near critical doping, x = 0.19. Here we

report a high-field magnetoresistance study of thin films of La2x

Srx

CuO4 cuprates in close

vicinity to critical doping, 0.161 x 0.190. We find that the metallic state exposed

by suppressing superconductivity is characterized by a magnetoresistance that is linear in

magnetic field up to the highest measured fields of 80T. The slope of the linear-in-field

resistivity is temperature-independent at very high fields. It mirrors the magnitude and

doping evolution of the linear-in-temperature resistivity that has been ascribed to Planckian

dissipation near a quantum critical point. This establishes true scale-invariant conductivity

as the signature of the strange metal state in high-temperature superconducting cuprates.

Email: [email protected]

arX

iv:1

705.

0580

6v1

[con

d-m

at.st

r-el

] 16

May

201

7

14

x

T

B

xcrit

Figure 3. Sketch of the x-B-T phase diagram in the vicinity of critical doping xcrit as

suggested by our measurement (superconducting transition is not shown). Solid-lines

indicate the edge of the region of linear-T behavior (marked in red) and linear-B behavior

(marked in blue). Dashed-lines indicate the region of the x-B-T phase space where

temperature and magnetic field compete.

Scale-invariant magnetoresistance in a cuprate superconductor

P. Giraldo-Gallo,1 J. A. Galvis,1 Z. Stegen,1 K. A. Modic,2 F. F Balakirev,3 J. B. Betts,3

X. Lian,1 C. Moir,1 S. C. Riggs,1 J. Wu,4 A. T. Bollinger,4 X. He,4, 5 I. Bozovic,4, 5

B. J. Ramshaw,6, 3 R. D. McDonald,3 G. S. Boebinger,1, 7 and A. Shekhter1,

1National High Magnetic Field Laboratory,

Florida State University, Tallahassee, FL 32310, USA

2Max-Planck-Institute for Chemical Physics of Solids,

Noethnitzer Strasse 40, D-01187, Dresden, Germany

3Los Alamos National Laboratory, Los Alamos, NM 87545, USA

4Brookhaven National Laboratory, Upton, New York 11973-5000, USA.

5Applied Physics Department, Yale University,

New Haven, Connecticut 06520, USA.

6Laboratory for Atomic and Solid State Physics,

Cornell University, Ithaca, NY 14853, USA

7Department of Physics, Florida State University, Tallahassee, FL 32310, USA

The anomalous metallic state in high-temperature superconducting cuprates is masked

by the onset of superconductivity near a quantum critical point. Use of high magnetic

fields to suppress superconductivity has enabled a detailed study of the ground state in

these systems. Yet, the direct e↵ect of strong magnetic fields on the metallic behavior at

low temperatures is poorly understood, especially near critical doping, x = 0.19. Here we

report a high-field magnetoresistance study of thin films of La2x

Srx

CuO4 cuprates in close

vicinity to critical doping, 0.161 x 0.190. We find that the metallic state exposed

by suppressing superconductivity is characterized by a magnetoresistance that is linear in

magnetic field up to the highest measured fields of 80T. The slope of the linear-in-field

resistivity is temperature-independent at very high fields. It mirrors the magnitude and

doping evolution of the linear-in-temperature resistivity that has been ascribed to Planckian

dissipation near a quantum critical point. This establishes true scale-invariant conductivity

as the signature of the strange metal state in high-temperature superconducting cuprates.

Email: [email protected]

arX

iv:1

705.

0580

6v1

[con

d-m

at.st

r-el

] 16

May

201

7

arXiv:1705.05806

Page 41: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

High field studies of cuprates have been crucial to unraveling their phase diagram

Plethora of evidence for an exotic metal underlying the underdoped regime

A metal with bulk topological order (i.e. long-range quantum entanglement) can explain existing experiments

Novel quantum criticality likely associated with a deconfined-confined transition in a gauge theory

Page 42: Topology, quantum entanglement, and criticality in the ...qpt.physics.harvard.edu/talks/nsf17.pdfamplitude grows with increasing field strength. Figure 1c shows the oscillatory part

explain existing experiments

Novel quantum criticality likely associated with a deconfined-confined transition in a gauge theory

Higher field studies, with more experimental probes (STM…), promise to answer many open questions, and could lead to direct detection of topological order.

An understanding of these issues is crucial to understanding the origin of the high critical temperature for superconductivity