ArticlePDF Available

Nematic fluctuations in the cuprate superconductor Bi2Sr2CaCu2O8+δ

Springer Nature
Nature Communications
Authors:

Abstract and Figures

Establishing the presence and the nature of a quantum critical point in their phase diagram is a central enigma of the high-temperature superconducting cuprates. It could explain their pseudogap and strange metal phases, and ultimately their high superconducting temperatures. Yet, while solid evidences exist in several unconventional superconductors of ubiquitous critical fluctuations associated to a quantum critical point, in the cuprates they remain undetected until now. Here using symmetry-resolved electronic Raman scattering in the cuprate Bi2Sr2CaCu2O8+δ, we report the observation of enhanced electronic nematic fluctuations near the endpoint of the pseudogap phase. While our data hint at the possible presence of an incipient nematic quantum critical point, the doping dependence of the nematic fluctuations deviates significantly from a canonical quantum critical scenario. The observed nematic instability rather appears to be tied to the presence of a van Hove singularity in the band structure. Solid evidence of quantum fluctuations associated to a quantum critical point in cuprate superconductors remains elusive. Here, Auvray et al. report Raman scattering evidence of enhanced electronic nematic fluctuations near the endpoint of the pseudogap phase in Bi2Sr2CaCu2O8+δ.
Dynamical nematic fluctuations in Bi2Sr2CaCu2O8+δ\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${\mathrm{Bi}}_2{\mathrm{Sr}}_2{\mathrm{CaCu}}_2{\mathrm{O}}_{8+\delta}$$\end{document}a Temperature-doping generic phase diagram of hole-doped cuprates. The pseudogap phase ends at a putative quantum critical point (QCP) located at the doping p*\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${p}^{* }$$\end{document}. b Nematic order breaking the C4\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${{\mathrm{C}}}_{4}$$\end{document} rotational symmetry of the Cu square lattice down to C2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${{\mathrm{C}}}_{2}$$\end{document} symmetry. The corresponding order parameter has B1g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{1g}$$\end{document} symmetry: in reciprocal space it transforms as kx2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{x}^{2}$$\end{document} − ky2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{y}^{2}$$\end{document} and switches sign upon 90 degrees rotation x→\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\to$$\end{document}y (color scale is defined as blue: negative values, red: positive values and white: 0). c Temperature dependence of the B1g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{1g}$$\end{document} Raman spectrum in the normal state for several doping levels in Bi2Sr2CaCu2O8+δ\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${\mathrm{Bi}}_{2}{\mathrm{{Sr}}}_{2}{\mathrm{CaCu}}_{2}{{\mathrm{O}}}_{8+\delta }$$\end{document}. The B1g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{1g}$$\end{document} symmetry is obtained using cross-photon polarizations at 45 degrees of the Cu–O–Cu direction (see insets)
… 
Symmetry resolved static susceptibilities. a Temperature dependences of the static susceptibility in 3 different symmetries, B1g,B2g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{1g},{B}_{2g}$$\end{document}, and A1g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${A}_{1g}$$\end{document}, as a function of doping. The form factors for each symmetry are depicted in reciprocal space in insets. They are given in terms of the lowest order Brillouin zone harmonics: cos(kx\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{x}$$\end{document}) − cos(ky\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{y}$$\end{document}) for B1g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{1g}$$\end{document}, sin(kx\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{x}$$\end{document})sin(ky\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{y}$$\end{document}) for B2g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{2g}$$\end{document}, cos(kx\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{x}$$\end{document}) + cos(ky\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{y}$$\end{document}) and cos(kx\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{x}$$\end{document})cos(ky\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{y}$$\end{document}) for A1g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${A}_{1g}$$\end{document}. The error bars correspond to the standard error of the low energy fits used for the low energy extrapolation (see supplementary note 1). b Curie-Weiss fits of the inverse B1g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{1g}$$\end{document} nematic susceptibility for temperatures above max(TcT*\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${T}_{c}\,{T}^{*}$$\end{document}). The inset shows the full temperature dependence of the inverse susceptibility of OD74 and UD85 where deviation from Curie-Weiss law are observed at T*\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${T}^{* }$$\end{document}, and an additional upturn is observed at Tc\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${T}_{c}$$\end{document}. Full and open symbol correspond to data above and below T*\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${T}^{*}$$\end{document}, respectively
… 
Phase diagram of critical nematic fluctuations. Color-coded plot summarizing the evolution of the B1g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{1g}$$\end{document} nematic susceptibility as a function of doping and temperature in Bi2212. The nematic Curie-Weiss temperature T0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${T}_{0}$$\end{document} is also shown along with the superconducting Tc\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${T}_{c}$$\end{document} and pseudogap T*\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${T}^{* }$$\end{document} temperatures. The lines are guide to the eye. The error bars for T0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${T}_{0}$$\end{document} correspond to the standard error of the Curie-Weiss fits. The inset shows the Fermi surface deformation associated to the incipient Pomeranchuk instability which breaks the C4\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${C}_{4}$$\end{document} symmetry
… 
This content is subject to copyright. Terms and conditions apply.
ARTICLE
Nematic uctuations in the cuprate superconductor
Bi
2
Sr
2
CaCu
2
O
8+δ
N. Auvray 1, B. Loret1, S. Benhabib1, M. Cazayous1, R.D. Zhong 2, J. Schneeloch2, G.D. Gu2, A. Forget3,
D. Colson3, I. Paul1, A. Sacuto 1& Y. Gallais 1*
Establishing the presence and the nature of a quantum critical point in their phase diagram is
a central enigma of the high-temperature superconducting cuprates. It could explain their
pseudogap and strange metal phases, and ultimately their high superconducting tempera-
tures. Yet, while solid evidences exist in several unconventional superconductors of ubiqui-
tous critical uctuations associated to a quantum critical point, in the cuprates they remain
undetected until now. Here using symmetry-resolved electronic Raman scattering in the
cuprate Bi2Sr2CaCu2O8þδ, we report the observation of enhanced electronic nematic uc-
tuations near the endpoint of the pseudogap phase. While our data hint at the possible
presence of an incipient nematic quantum critical point, the doping dependence of the
nematic uctuations deviates signicantly from a canonical quantum critical scenario. The
observed nematic instability rather appears to be tied to the presence of a van Hove sin-
gularity in the band structure.
https://doi.org/10.1038/s41467-019-12940-w OPEN
1Université de Paris, Matériaux et Phénomènes Quantiques, CNRS UMR 7162, F-75205 Paris, France. 2Condensed Matter Physics and Materials Science
Department, Brookhaven National Laboratory, Upton, NY 11973, USA. 3Service de Physique de lÉtat Condensé, DRF/IRAMIS/SPEC (UMR 3680 CNRS), CEA
Saclay, 91191 Gif-sur-Yvette cedex, France. *email: yann.gallais@univ-paris-diderot.fr
NATURE COMMUNICATIONS | (2019) 10:5209 | https://doi.org/10.1038/s41467-019-12940-w | www.nature.com/naturecommunications 1
1234567890():,;
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Unconventional superconductivity (SC) is often linked to
the proximity of an electronically ordered phase whose
termination at a quantum critical point (QCP) is located
inside a superconducting dome-like region1,2. This observation
suggests quantum criticality as an organizing principle of their
phase diagram. The associated critical uctuations could act as
possible source for enhanced superconducting pairing35and
explain their ubiquitous strange metal phases which often
show non-Fermi liquid behavior, like a linear in temperature
resistivity7. However whether the quantum critical point scenario
holds for high-Tcsuperconducting cuprates has remained largely
unsettled68. This can be traced back to two fundamental reasons.
First the exact nature of the pseudogap (PG) state, the most
natural candidate for the ordered phase, remains mysterious.
Experimentally a wealth of broken symmetry phases have been
reported at, or below, the somewhat the loosely dened PG
temperature T9. It is currently unclear which, if any, of these
orders is the main driver of the PG phase as they could be all
secondary instabilities of a pre-existing PG state. Second there is
little direct evidence for critical uctuations above T, questioning
the very existence of a QCP associated to the termination of a
second order phase transition at a critical doping p(see Fig. 1a).
AF uctuations do not appear to be critical close to the putative
QCP, at least for hole-doped cuprates10,11, and critical CDW
uctuations are only observed below T12. Fluctuations associated
to the more subtle intra-unit-cell orders1317 are more elusive and
have not been reported up to now.
Recently nematicity, an electronic state with broken rotational
symmetry but preserved translational invariance of the under-
lying lattice18 (see Fig. 1b), has emerged as a potential candidate
for the origin of the PG phase. A second order phase transition to
a nematic phase, breaking the C4rotational symmetry of the
CuO2plane, has been reported by torque magnetometry in
YBa2Cu3O6þδclose to the TTreported by other techniques19.In
a separate study a divergent electronic specic heat coefcient was
observed at the endpoint of the PG phase in several cuprates and
was interpreted as a thermodynamical hallmark of a QCP20. The
nature of the ordered state associated to this putative QCP is
however not yet settled, and nematicity stands as a potential
candidate. To assess its relevance and the role of nematic degrees
of freedom in driving the PG order, probing the associated
uctuations is thus essential.
Because it probes uniform (q=0) dynamical electronic uctua-
tions in a symmetry selective way, electronic Raman scattering can
access nematic uctuations without applying any strain even in
nominally tetragonal systems21. For metallic systems, the nematic
uctuations probed by Raman scattering can be thought as dyna-
mical Fermi surface deformations which break the lattice point
group symmetry. In the context of iron-based superconductors (Fe
SC) ubiquitous critical nematic uctuations were observed by
Raman scattering in several compounds2224.Theywereshownto
drive the C4symmetry breaking structural transition from the
tetragonal to the orthorhombic lattice, and to persist over a sig-
nicant portion of their phase diagram21. In the context of cuprates
nematicity along Cu-O-Cu bonds has been reported via several
techniques, mostly in YBa2Cu3O6þδ1417,19,25,26. The associated
order parameter is an uniform traceless tensor of B1g(or x2y2)
symmetry, which switches signs upon interchanging the xand y
axis of the CuO2square plane (Fig. 1b). Nematicity along different
directions has also been found in one layer mercury-based cuprate
HgBa2CuO427, and overdoped La1xSrxCuO428. In the former case
nematicity develops along the diagonal of the CuO2plane and thus
transforms as the B2g(or xy) symmetry. The ability of Raman
scattering to resolve the symmetry of the associated order parameter
is therefore crucial.
Results
Doping and symmetry dependent Raman spectra. We present
Raman scattering measurements on 6 single crystals of the cuprate
Bi2Ca2SrCu2O8þδ(Bi2212) covering a doping range between p=
0.12 and p=0.23. A particular emphasis was put in the doping
region bracketing p0:22 close to where the PG was shown to
terminate, i.e., between p=0.20 and p=0.232936. At these dop-
ings a relatively wide temperature range is accessible above both T
and Tcto probe these uctuations, and look for ngerprints of a
nematic QCP. The polarization resolved Raman experiments were
performed in several congurations of in-plane incoming and
outgoing photon polarizations in order to extract the relevant
irreducible representations, or symmetries, of the D4hgroup: B1g
which transforms as x2y2
;B2g(xy)andA1g. As indicated above
while the former two correspond to nematic orders along and at 45
degrees of the CuOCu bonds, respectively, the latter one is fully
symmetric and is not associated to any symmetry breaking. The
recorded spectra were corrected by the Bose factor and are thus
proportional to the imaginary part of the frequency dependent
Raman response function χ00
μðωÞin the corresponding symmetry μ
where μ=B1g;B2g;A1g(see Methods section for more details on
the Raman scattering set-up and polarization congurations).
In Fig. 1c is displayed the evolution of the normal state Raman
spectrum in the B1gsymmetry as a function of doping. From
previous Raman studies, OD60 (Tc=60K) sample lies very close
to the termination point of the PG, p0.22 and no signature of
PG is seen for OD60, OD55 (Tc=55K), and OD52 (Tc=52K)
compositions36. Other samples, OD74 (Tc=74K), OD80 (Tc=
80K), and UD85 (Tc=85K) display PG behavior. The normal
state spectra are consistent with previously published Raman data
for the doping compositions where they overlap37. In particular
while the spectrum shows little temperature dependence in the
underdoped composition UD85, it acquires a signicant tem-
perature dependence in the overdoped regime where the overall
B1gRaman response increases upon cooling. The low-frequency
slope of the Raman response being proportional to the lifetime of
the quasiparticles, this evolution was previously attributed to an
increase metallicity of anti-nodal quasiparticles, located at (π,0)
and equivalent points of the Brillouin zone with overdoping37.
However the increase of intensity upon cooling observed in
overdoped compositions, p>0:2, is not conned to low
frequencies as expected in a naive Drude model, but extends
over wide energy range up to at least 500 cm1. This suggests that
it is not a simple quasiparticle lifetime effect but, as we show
immediately below, it is rather associated to a strongly
temperature dependent static nematic susceptibility. It is also
evident in Fig. 1c that this evolution is non-monotonic with
doping as the OD60 spectra shows signicantly more temperature
dependence than at any other dopings.
Symmetry-resolved susceptibilities. To analyze the observed
temperature dependence and its link to a nematic instability, it is
useful to extract the symmetry resolved static susceptibility
χμðω¼0Þ=χ0
μfrom the measured nite frequency response
χ00
μðωÞusing Kramers-Kronig relations:
χ0
μ¼ZΛ
0
dωχ00ðωÞ
ωð1Þ
where μstands for the symmetry and Λa high-energy cut-off. In
order to perform the integration, the spectra were interpolated to
zero frequency either linearly, or using a Drude lineshape (see
Supplementary Note 2 and Supplementary Figs. 2 and 3).
The integration was performed up to Λ=800 cm1, above
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12940-w
2NATURE COMMUNICATIONS | (2019) 10:5209 | https://doi.org/10.1038/s41467-019-12940-w | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
which the spectra do not show any appreciable temperature
dependence in the normal state (see Supplementary Fig. 4). For
B1gsymmetry χ0
B1gis directly proportional to the static electronic
nematic susceptibility, and its evolution as a function of doping
and temperature is shown in Fig. 2a. For a comparison the same
quantity, extracted for 3 crystals in the complementary symme-
tries (see Supplementary Note 1 for the spectra), B2gand A1g,is
also shown. In order to compare different compositions and since
3
0
SC
PG
ab
c
T
T*
p*
es
el
p
y
x
C4C2
0
kx
ky
UD85
OD80
OD74
OD52
OD55
OD60
61 K 83 K 63 K 70 K
88 K
108 K
128 K
148 K
168 K
188 K
208 K
228 K
248 K
268 K
288 K
308 K
73 K
162 K
191 K
221 K
250 K
275 K
300 K
82 K
97 K
113 K
137 K
128 K
148 K
168 K
188 K
208 K
248 K
268 K
303 K
81 K 120 K 102 K 99 K
113 K
137 K
162 K
187 K
211 K
236 K
260 K
281 K
301 K
112 K
139 K
164 K
193 K
222 K
250 K
276 K
302 K
128 K
158 K
188 K
218 K
248 K
268 K
298 K
85 K
90 K
95 K
100 K
105 K
110 K
115 K
88 K
93 K
98 K
103 K
108 K
65 K
70 K
74 K
79 K
2
B1g
(norm. at 800 cm–1)
1
0
3
2
1
0
0 200 400 600 0 200 400
Raman shift (cm–1)
600 0 200 400 600
UD85
OD80
OD74
OD52
B1g
OD55 OD60
Fig. 1 Dynamical nematic uctuations in Bi2Sr2CaCu2O8þδaTemperature-doping generic phase diagram of hole-doped cuprates. The pseudogap phase
ends at a putative quantum critical point (QCP) located at the doping p.bNematic order breaking the C4rotational symmetry of the Cu square lattice
down to C2symmetry. The corresponding order parameter has B1gsymmetry: in reciprocal space it transforms as k2
x
k2
yand switches sign upon
90 degrees rotation x!y(color scale is dened as blue: negative values, red: positive values and white: 0). cTemperature dependence of the B1gRaman
spectrum in the normal state for several doping levels in Bi2Sr2CaCu2O8þδ. The B1gsymmetry is obtained using cross-photon polarizations at 45 degrees of
the CuOCu direction (see insets)
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12940-w ARTICLE
NATURE COMMUNICATIONS | (2019) 10:5209 | https://doi.org/10.1038/s41467-019-12940-w | www.nature.com/naturecommunications 3
Content courtesy of Springer Nature, terms of use apply. Rights reserved
we do not have access to the susceptibilities in absolute units, all
extracted susceptibilities have been normalized to their average
value close to 300 K, and we focus on their temperature depen-
dences. The temperature dependence of the B1gnematic sus-
ceptibility is strongly doping dependent. In the normal state it is
only weakly temperature dependent for UD85 (p=0.13), while
for overdoped compositions it displays a signicant enhancement
upon cooling. The pronounced divergent-like behavior for OD60
is only cut-off by the entrance to the SC state. This effect is
however reduced for p>p
(OD55 and OD52), mirroring the
non-monotonic behavior already apparent in the raw spectra.
While our focus here is on the normal state, it is notable that the
nematic susceptibility is suppressed upon entering the super-
conducting state for all doping studied, suggesting that the
nematic instability is suppressed by the emergence of the super-
conducting order. In addition a weak but distinct suppression of
χ0
μis also observed above Tcfor UD85 (200 K), close to the value
of Tdetermined by other techniques in Bi2212 for similar doping
levels33,38. By contrast the static susceptibilities extracted in B2g
and A1gsymmetries, while displaying some mild enhancement,
are essentially doping independent above Tc. This symmetry
selective behavior unambiguously demonstrates the nematic nat-
ure of the critical uctuations observed close to p0.22.
Curie-Weiss analysis of the B1gnematic susceptibility. Further
insight into the doping dependence of these critical nematic
uctuations can be gained by tting the B1gstatic nematic sus-
ceptibility using a Curie-Weiss law:
1
χ0
B1g
¼A´ðTT0Þð2Þ
In a mean-eld picture of the electronic nematic transition, such a
behavior is expected on the high temperature tetragonal side of a
second order phase transition which would set in at T0.A
negative T0implies that the ground state is on the symmetry
unbroken side of the phase transition. Fig. 2(b) shows linear ts
of the inverse susceptibility for all doping studied. Since clear
deviations to linear behavior for 1
χ0
B1g
are seen below Tand Tc(see
inset of Fig. 2b), we restrict our ts to temperatures above Tfor
doping levels below p, and above Tcfor doping level above p*.
The ts allows us to extract T0, the mean-eld nematic transition
temperature, which quanties the strength of the nematic
instability: graphically T0corresponds to the zero temperature
intercept of the inverse susceptibility.
The evolution of T0as a function of doping is summarized in
Fig. 3in a phase diagram showing the corresponding evolution of
the nematic susceptibility in a color-coded plot. Coming from the
strongly overdoped regime, p0.23, T0increases towards
p0.22 but upon further reducing doping, instead of
crossing-over to positive temperatures T0reverses its behavior
and decreases strongly, suggesting a signicant weakening of the
nematic instability below p. While the T0values remain negative
at all doping suggesting the absence of true nematic quantum
criticality, two aspects should be borne in mind. First the three T0
values above pextrapolate to T=0 K at a doping level above the
one corresponding to OD74, indicating the possible presence of a
nematic QCP located between the doping levels corresponding to
OD74 and OD60 crystals. Second, as shown in the context of Fe
SC, the extracted susceptibility from Raman measurements does
not include the contribution of the electron-lattice coupling21.In
particular the linear nemato-elastic coupling is expected to
increase the nematic transition temperature above T0and shift
accordingly the location of the QCP39. This lattice-induced shift
2
a
1.0
0.5
1
2
1
2A1g0
UD85 K
OD80 K
OD74 K
UD85 K
OD80 K
OD74 K
OD60 K
OD55 K
OD52 K
OD60 K
OD55 K
OD52 K
B2g
B1g
1
0 100 200 300 100 200
Temperature (K)
Temperature (K)
0
1.5
1/B1g
(norm.)
1.0
100
Tc = 85 K
Tc = 74 K
T* = 200 K
T* = 125 K
200 300
Temperature (K)
300
0
0
1/0 (norm. at 300 K)
B1g
(300 K)
0 (T)/0
b
Fig. 2 Symmetry resolved static susceptibilities. aTemperature dependences of the static susceptibility in 3 different symmetries, B1g;B2g, and A1g,asa
function of doping. The form factors for each symmetry are depicted in reciprocal space in insets. They are given in terms of the lowest order Brillouin zone
harmonics: cos(kx)cos(ky) for B1g, sin(kx)sin(ky) for B2g, cos(kx)+cos(ky) and cos(kx)cos(ky) for A1g. The error bars correspond to the standard error of
the low energy ts used for the low energy extrapolation (see supplementary note 1). bCurie-Weiss ts of the inverse B1gnematic susceptibility for
temperatures above max(TcT). The inset shows the full temperature dependence of the inverse susceptibility of OD74 and UD85 where deviation from
Curie-Weiss law are observed at T, and an additional upturn is observed at Tc. Full and open symbol correspond to data above and below T, respectively
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12940-w
4NATURE COMMUNICATIONS | (2019) 10:5209 | https://doi.org/10.1038/s41467-019-12940-w | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
of T0is on the order of 3060 K in Fe SC21, but is not known in
the case of cuprates. Recent elasto-resistance measurements
suggest that the nemato-elastic coupling might be weaker in
cuprates3941. Irrespective of the presence of a nematic QCP, it is
clear that the doping evolution of T0contrasts with the canonical
behavior of a QCP where T0would evolve monotonically as a
function of the tuning parameter.
Discussion
We now discuss the possible origin of the observed nematic
uctuations, and then elaborate on the implications of our nd-
ings for the phase diagram of the cuprates. Theoretically two
routes for nematicity have been proposed in the context of cup-
rates. The rst route is via the melting of an uni-axial density
wave order, like stripe or charge density wave, and is expected to
apply to the underdoped regime where the tendency towards
these orders is strongest42. It is unlikely to be relevant to our
ndings since the nematic susceptibility is almost featureless
above Tin the underdoped crystal, and only shows signicant
enhancement in the overdoped regime. A second route is via a
Pomeranchuk instability of the Fermi liquid, where the Fermi
surface spontaneously deforms and breaks the underlying lattice
rotational symmetry (see inset of Fig. 3). Theoretically, such an
instability was indeed shown to be relevant to the cuprates close
to doping levels where the density of states passes through a van
Hove singularity (vHs) at the (π, 0) and equivalent points of the
Brillouin zone4345. This is consistent with our results since p=
0.22 corresponds to the doping level where one of the Fermi
surface sheets of Bi2212 changes from electron-like to hole-like,
and passes through a van Hove singularity (vHs) at (π,0)
29,36.
The link between the nematic instability and the closeness of the
vHs singularity also naturally explains the non-monotonic
behavior of T0as a function of doping, which is also found in
mean eld theories of vHs induced nematicity46,47. Interestingly
non-Fermi liquid behavior has been argued to occur generically
near a nematic QCP if the coupling to the lattice is weak
enough39,48,49, and the observed critical nematic uctuations may
therefore play a central role in the the strange metal properties
observed in this doping range. While our results suggest that the
nematic instability is linked to the proximity of vHs, we stress
that the observed enhancement cannot be merely a consequence
of a high-density of states at the (π, 0) points: both B1gand A1g
form factors have nite weight at these points, but only the
nematic B1gsusceptibility shows ngerprints of critical behavior
at p. Thus electronic interactions in the nematic channel appears
to be essential to explain our observation. We note that the key
role of interactions, beyond density of state effects, was also
argued to explain the divergence of the electronic specic heat
coefcient which was observed at the PG end-point of several
one-layer cuprates20.
What are the consequences of our ndings for the nature of the
PG state? The above discussion and the doping dependence of T0
allow us to conclude that the PG is likely not driven by a nematic
instability. If this was the case one would expect strong nematic
uctuations close to Tin the underdoped composition and
a monotonic increase of T0, crossing-over to positive values,
as observed in the case of Fe SC22,50,51. It therefore appears that
nematic and PG instabilities are distinct, and possibly even
compete. Note however that since our analysis of T0has been
restricted to temperature above Tthe weakening of the nematic
uctuations below pcannot be a simple consequence of the PG
order. Intriguingly, the states at the (π;0) point appears to be
critical for both orders: while the strength of nematic uctuations
is closely tied to the closeness to the vHs at these points, it was
suggested that the PG regime is characterized by a strong deco-
herence at these hot-spots due to AF uctuations that set in once
the Fermi surface reaches the (π, 0) points corresponding to the
AF zone boundary. This decoherence ultimately drives the Fermi
surface hole-like and induces a cross-over to a PG state at low
temperature52,53. Our results thus indicate that both nematicity
and the PG state depend critically on the Fermi surface topology
in the case of Bi2212. Further measurements on cuprates com-
pounds where the PG endpoint and the change in the Fermi
surface topology occur at distinct doping levels, like
Tl2Ba2CuO6þδ54,55 and YBa2Cu3O6þδ56,57, are needed in order to
clarify the nature of this link and conrm the connection between
the nematic instability and the presence of a vHs.
Methods
Samples. The Bi-2212 single crystals were grown by using a oating zone method.
First optimal doped samples with Tc=90 K were grown at a velocity of 0.2 mm per
hour in air. In order to get overdoped samples down to Tc=65 K, the as-grown
single crystal was put into a high oxygen pressured cell between 1000 and 2000 bars
and then was annealed from 350 C to 500C during 3 days. The overdoped
samples below Tc=60 K were obtained from as-grown Bi-2212 single crystals put
into a pressure cell (Autoclave France) with 100 bars oxygen pressure and annealed
from 9 to 12 days at 350 C . Then the samples were rapidly cooled down to room
temperature by maintaining a pressure of 100 bars. The underdoped sample was
obtained by annealing the as-grown sample in vacuum. The critical temperature Tc
for each crystal has been determined from magnetization susceptibility measure-
ments at a 10 Gauss eld parallel to the c-axis of the crystal. The selected crystals
exhibit a quality factor of Tc
ΔTclarger than 7. ΔTcis the full width of the super-
conducting transition. A complementary estimate of Tcwas achieved from elec-
tronic Raman scattering measurements by dening the temperature from which
the B1gsuperconducting pair breaking peak collapses. Special care has been
devoted to select single crystals which exhibit the same SC pair-breaking peak
energy in the Raman spectra measured from distinct laser spots on a freshly cleaved
surface. The level of doping p was dened from Tcusing Presland and Tallons
equation58:
1T
Tc;max
¼82:6´ðp0:16Þ2ð3Þ
Details of the Raman spectroscopy experiments. Raman experiments have been
carried out using a triple grating JY-T64000 spectrometer in subtractive mode
300
250
200
150
Temperature (K)
1.9
1.7
1.6
1.4
1.2
1.0
1.8
0.7
100
50
Hole doping p
0
PG
SC
–500
–1000
0.14 0.16 0.18 0.20 0.22
T
c
T
0
T*
0
0
0
B
1g
Fig. 3 Phase diagram of critical nematic uctuations. Color-coded plot
summarizing the evolution of the B1gnematic susceptibility as a function of
doping and temperature in Bi2212. The nematic Curie-Weiss temperature
T0is also shown along with the superconducting Tcand pseudogap T
temperatures. The lines are guide to the eye. The error bars for T0
correspond to the standard error of the Curie-Weiss ts. The inset shows
the Fermi surface deformation associated to the incipient Pomeranchuk
instability which breaks the C4symmetry
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12940-w ARTICLE
NATURE COMMUNICATIONS | (2019) 10:5209 | https://doi.org/10.1038/s41467-019-12940-w | www.nature.com/naturecommunications 5
Content courtesy of Springer Nature, terms of use apply. Rights reserved
using two 1800 grooves/mm gratings in the pre-monochromator stage and
600 grooves/mm or 1800 groove/mm grating in the spectrograph stage. The
600 grooves/mm grating was used for all measurements except those carried on the
OD80 sample, for which a 1800 grooves/mm grating was used. For several crystals,
both congurations were used at selected temperatures to check for consistency.
The 600 grooves/mm conguration allows us to cover the low-energy part of the
spectrum down to 50 cm1and up to 900 cm1in a single frame. With the 1800
grooves/mm grating, measurements could be performed down to 15 cm1, but
spectra must then be obtained in two frames. The resolution is set at 5 cm1when
using the 600 grooves/mm conguration. The spectrometer is equipped with a
nitrogen cooled back illuminated CCD detector. We use the 532 nm excitation line
from a diode pump solid state laser. Measurements between 10 and 300 K have
been performed using an ARS closed-cycle He cryostat.
All the raw spectra have been corrected for the Bose factor and the instrumental
spectral response. They are thus proportional to the imaginary part of the Raman
response function χ00ðω;TÞ. A potential concern when correcting the raw spectra
with the Bose factor is the potential presence of non-Raman signal in the raw
spectra. To assess this potential non-Raman signal we note that a single effective
spot temperature was able to reproduce the measured Stokes spectrum from the
anti-Stokes spectrum between 20 and 600 cm1at room temperature. In addition
the raw Raman spectra were found to extrapolate very close to zero at zero Raman
shift at the lowest temperatures measured. Both facts indicates negligible non-
Raman background in the measured spectra.
The direction of incoming and outgoing electric elds are contained in the ðabÞ
plane. The A1gþB2ggeometries are obtained from parallel polarizations at 45
from the CuO bond directions; the B2gand B1ggeometries are obtained from
crossed polarizations along and at 45from the CuO bond directions,
respectively. The crystal was rotated in-situ using an Attocube piezo-rotator to
align the electric eld with respect to the crystallographic axis. A1gspectra are
obtained from the previously listed geometries (see Supplementary Note 1 and
Supplementary Fig. 1).
Data availability
All data generated or analyzed during this study are included in the published
manuscript and the supplementary information les. The relevant raw data le are
available at the following url: https://doi.org/10.6084/m9.gshare.9906275.v1
Received: 25 April 2019; Accepted: 11 October 2019;
References
1. Lohneysen, H. V., Rosch, A., Vojta, M. & Wole, P. Fermi-liquid instabilities
at magnetic quantum phase transitions. Rev. Mod. Phys. 79, 1015 (2007).
2. Shibauchi, T., Carrington, A. & Matsuda, Y. Quantum critical point lying
beneath the superconducting dome in iron-pnictides. Annu. Rev. Condens.
Matter Phys. 5, 113 (2014).
3. Monthoux, P., Pines, D. & Lonzarich, G. G. Superconductivity without
phonons. Nature 450, 11771183 (2007).
4. Metlitski, M. A., Mross, D. F., Sachdev, S. & Senthil, T. Cooper pairing in non-
fermi liquids. Phys. Rev. B 91, 115111 (2015).
5. Labat, D. & Paul, I. Pairing instability near a lattice-inuenced nematic
quantum critical point. Phys. Rev. B 96, 195146 (2017).
6. Broun, D. M. What lies beneath the dome? Nat. Phys. 4, 170172 (2008).
7. Taillefer, L. Scattering and pairing in High-Tccuprates. Annu. Rev. Condens.
Matter Phys. 1, 51 (2010).
8. Varma, C. M. Quantum-critical uctuations in 2D metals: strange metals and
superconductivity in antiferromagnets and in cuprates. Rep. Prog. Phys. 79,
082501 (2016).
9. Keimer, B., Kivelson, S. A., Norman, M. R., Uchida, S. & Zaanen, J. From
quantum matter to high-temperature superconductivity in copper oxides.
Nature 518, 179186 (2015).
10. Reznik, D. et al. Local-moment uctuations in the optimally doped high-Tc
superconductor YBa2Cu3O6þδ.Phys. Rev. B 78, 132503 (2008).
11. Li, Y. et al. Low-energy antiferromagnetic spin uctuations limit the coherent
superconducting gap in cuprates. Phys. Rev. B 98, 224508 (2018).
12. Arpaia, R. et al. Dynamical charge density uctuations pervading the phase
diagram of a Cu-based high-Tc superconductor. Science 365, 906910 (2019).
13. Fauqué, B. et al. Magnetic order in the Pseudogap phase of high-Tc
superconductors. Phys. Rev. Lett. 96, 197001 (2006).
14. Ando, Y., Segawa, K., Komiya, S. & Lavrov, A. N. Electrical resistivity
anisotropy from self-organized one dimensionality in high-temperature
superconductors. Phys. Rev. Lett. 88, 137005 (2002).
15. Hinkov, V. et al. Electronic liquid crystal state in the high-temperature
superconductor YBa2Cu3O6þδ.Science 319, 597600 (2008).
16. Lawler, M. J. et al. Intra-unit-cell electronic nematicity of the high-Tc copper-
oxide pseudogap states. Nature 466, 347 (2010).
17. Daou, R. et al. Broken rotational symmetry in the pseudogap phase of a high-
Tc superconductor. Nature 463, 519 (2010).
18. Fradkin, E., Kivelson, S. A., Lawler, M. J., Eisenstein, J. P. & Mackenzie, A. P.
Nematic fermi uids in condensed matter physics. Annu. Rev. Cond. Matt. 1,
153 (2010).
19. Sato, Y. et al. Thermodynamic evidence for a nematic phase transition at the
onset of the pseudogap in YBa2Cu3O6þδ.Nat. Phys. 13, 10741087 (2017).
20. Michon, B. et al. Thermodynamic signatures of quantum criticality in cuprate
superconductors. Nature 567, 218222 (2019).
21. Gallais, Y. & Paul, I. Charge nematixity and electronic Raman scattering in
Iron-based superconductors. C. R. Phys. 17, 113 (2016).
22. Gallais, Y. et al. Observation of incipient charge nematicity in Ba(Fe1xCox
As)2single crystals. Phys. Rev. Lett. 111, 267001 (2013).
23. Thorsmlle, V. K. et al. Critical quadrupole uctuations and collective modes in
iron pnictide superconductors. Phys. Rev. B 93, 054515 (2016).
24. Massat, P. et al. Charge-induced nematicity in FeSe. Proc. Natl Acad. Sci. 113,
9177 (2016).
25. Cyr-Choinière, O. et al. Two types of nematicity in the phase diagram of the
cuprate superconductor YBa2Cu3O6þδ.Phys. Rev. B 92, 224502 (2015).
26. Achkar, A. J. et al. Nematicity in stripe-ordered cuprates probed via resonant
x-ray scattering. Science 351, 576 (2016).
27. Murayama, H. et al. Diagonal nematicity in the Pseudogap phase of HgBa2
CuO4þδ.Nat. Commun. 10, 3282 (2019).
28. Wu, J., Bollinger, A. T., He, X. & Bozovic, I. Spontaneous breaking of
rotational symmetry in copper oxide superconductors. Nature 547, 432435
(2017).
29. Benhabib, S. et al. Collapse of the normal-state Pseudogap at a lifshitz
transition in the Bi2Sr2CaCu2O8þδcuprate superconductor. Phys. Rev. Lett.
114, 147001 (2015).
30. Ishida, K. et al. Pseudogap behavior in single-crystal Bi2Sr2CaCu2O8þδ
probed by Cu NMR. Phys. Rev. B 58, R5960 (1997).
31. Oda, M. et al. Strong pairing interactions in the underdoped region of Bi2Sr2
CaCu2O8þδ.Phys. C 281, 135142 (1997).
32. Dipasupil, R. M., Oda, M., Momono, N. & Ido, M. Energy gap evolution in the
tunneling spectra of Bi2Sr2CaCu2O8þδ.J. Phys. Soc. Jpn. 71, 15351540
(2002).
33. Vishik, I. M. et al. Phase competition in trisected superconducting dome. Proc.
Natl Acad. Sci 109, 18332 (2012).
34. TomohiroUsui et al. Doping dependencies of onset temperatures for the
Pseudogap and superconductive uctuation in Bi2Sr2CaCu2O8þδ, studied
from both in-plane and out-of-plane magnetoresistance measurements. J.
Phys. Soc. Japan 83, 064713 (2014).
35. Hashimoto, M. et al. Direct spectroscopic evidence for phase competition
between the pseudogap and superconductivity in Bi2Sr2CaCu2O8þδ.Nat.
Mater 14,3742 (2015).
36. Loret, B. et al. Raman and ARPES combined study on the connection between
the existence of the pseudogap and the topology of the Fermi surface in Bi2Sr2
CaCu2O8þδ.Phys. Rev. B 97, 174521 (2018).
37. Venturini, F. et al. Observation of an unconventional metal-insulator
transition in overdoped CuO2compounds. Phys. Rev. Lett. 89, 107003 (2002).
38. Mangin-Thro, L. et al. Characterization of the intra-unit-cell magnetic order
in Bi2Sr2CaCu2O8þδ.Phys. Rev. B 89, 094523 (2014).
39. Paul, I. & Garst, M. Lattice effects on nematic quantum criticality in metals.
Phys. Rev. Lett. 118, 227601 (2017).
40. Xie, T. et al. Spontaneous nematic transition within the pseudogap state in
cuprates. Preprint at https://arxiv.org/abs/1805.00666 (2018).
41. Ishida, Y. et al. Divergent nematic susceptibility near the pseudogap critical
point in a cuprate superconductor. Preprint at https://arxiv.org/abs/
1908.07167 (2019).
42. Kivelson, S. A., Fradkin, E. & Emery, V. J. Electronic liquid-crystal phases of a
doped Mott insulator. Nature 393, 550 (1998).
43. Yamase, H. & Kohno, H. Instability toward formation of quasi-one-
dimensional fermi surface in two-dimensional t-J model. J. Phys. Soc. Jpn 69,
332 (2000).
44. Halbroth, C. J. & Metzner, W. d-Wave Superconductivity and Pomeranchuk
Instability in the Two-Dimensional Hubbard Model. Phys. Rev. Lett. 85, 5162
(2000).
45. Bulut, S., Atkinson, W. A. & Kampf, A. P. Spatially modulated electronic
nematicity in the three-band model of cuprate superconductors. Phys. Rev. B
88, 155132 (2013).
46. Khavkine, I., Chung, C.-H., Oganesyan, V. & Kee, H.-Y. Formation of an
electronic nematic phase in interacting fermion systems. Phys. Rev. B 70,
155110 (2004).
47. Yamase, H., Oganesyan, V. & Metzner, W. Mean-eld theory for symmetry-
breaking Fermi surface deformation on a square lattice. Phys. Rev. B 72,
035114 (2005).
48. Oganesyan, V., Kivelson, S. A. & Fradkin Quantum theory of a nematic Fermi
uid. Phys. Rev. B 64, 195109 (2001).
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12940-w
6NATURE COMMUNICATIONS | (2019) 10:5209 | https://doi.org/10.1038/s41467-019-12940-w | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
49. Metzner, W., Rohe, D. & Andergassen, S. Soft fermi surfaces and breakdown
of Fermi-liquid behavior. Phys. Rev. Lett. 91, 066402 (2003).
50. Chu, J.-H., Kuo, H. H. J., Analytis, J. G. & Fisher, I. R. Divergent nematic
susceptibility in an iron arsenide superconductor. Science 337, 710712
(2012).
51. Hosoi, S. et al. Nematic quantum critical point without magnetism in
FeSe1xSxsuperconductors. Proc. Natl Acad. Sci. USA 113, 8139 (2016).
52. Wu, W. et al. Pseudogap and fermi-surface topology in the two-dimensional
Hubbard model. Phys. Rev. X 8, 021048 (2018).
53. Braganca, H., Sakai, S., Aguiar, M. C. O. & Civelli, M. Correlation-driven
Lifshitz transition at the emergence of the Pseudogap Phase in the two-
dimensional Hubbard model. Phys. Rev. Lett. 120, 067002 (2018).
54. Proust, C., Boaknin, E., Hill, R. W., Taillefer, L. & Mackenzie, A. P. Heat
transport in a strongly overdoped cuprate: Fermi Liquid and a Pure d-Wave
BCS Superconductor. Phys. Rev. Lett. 89, 147003 (2002).
55. Plate, M. et al. Fermi surface and quasiparticle excitations of overdoped Tl2
Ba2CuO6þδ.Phys. Rev. Lett. 95, 077001 (2005).
56. Tallon, J. L. & Loram, J. W. The doping dependence of T*-what is the real
high-Tc phase diagram? Phys. C 349,5368 (2001).
57. Fournier, D. et al. Loss of nodal quasiparticle integrity in underdoped YBa2
Cu3O6þδ.Nat. Phys. 6, 905911 (2010).
58. Presland, M., Tallon, J., Buckley, R., Liu, R. & Flower, N. Phys. C 176,95
(1991).
Acknowledgements
We thank T. Shibauchi for fruitful discussions. The work at Brookhaven National
Laboratory was supported by the Ofce of Science, U.S. Department of Energy under
Contract No. DE-SC0012704.
Author contributions
N.A., B.L., and S.B. performed the Raman scattering experiments with the help of M.C.,
A.S., and Y.G. N.A. performed the data analysis and prepared the gures. R.D.Z., J.S., G.
Gu. grew the single crystals and the annealing procedure to obtain underdoped and
overdoped compositions. A.F. and D.C. performed the high-pressure annealing of the
crystals for the strongly overdoped compositions. I.P. provided theoretical insights. A.S.
and Y.G. supervised the project. Y.G. wrote the paper with inputs from all the authors.
Competing interests
The authors declare no competing interests
Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/s41467-
019-12940-w.
Correspondence and requests for materials should be addressed to Y.G.
Peer review information Nature Communications thanks Bruce Gaulin and the other,
anonymous, reviewer(s) for their contribution to the peer review of this work. Peer
reviewer reports are available.
Reprints and permission information is available at http://www.nature.com/reprints
Publishers note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional afliations.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative
Commons license, and indicate if changes were made. The images or other third party
material in this article are included in the articles Creative Commons license, unless
indicated otherwise in a credit line to the material. If material is not included in the
articles Creative Commons license and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this license, visit http://creativecommons.org/
licenses/by/4.0/.
© The Author(s) 2019
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12940-w ARTICLE
NATURE COMMUNICATIONS | (2019) 10:5209 | https://doi.org/10.1038/s41467-019-12940-w | www.nature.com/naturecommunications 7
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... The study of critical fluctuations near the transition temperature in high-temperature cuprate superconductors, HTSC, has attracted much interest since the discovery of these materials [1][2][3][4][5][6][7][8]. In HTSC, these critical effects are especially significant due, mainly, to the short coherence lengths and corresponding reduced-dimensionality enhancements when competing with the size of the intrinsic layered nanostructure formed by the CuO 2 superconducting planes [1][2][3][4][9][10][11]. ...
... In HTSC, these critical effects are especially significant due, mainly, to the short coherence lengths and corresponding reduced-dimensionality enhancements when competing with the size of the intrinsic layered nanostructure formed by the CuO 2 superconducting planes [1][2][3][4][9][10][11]. It was quite early noted that the temperature behavior of the critical fluctuations (including both critical exponents and amplitudes) could provide information about HTSC such as, e.g., the locus where superconductivity occurs, the symmetry of the pairing wave function, or the possible influence of phase fluctuations on the high value of transition temperature itself [1][2][3][4][5][6][7][8][9][10][11][12][13]. ...
Preprint
Full-text available
We study the critical fluctuations near the resistive transition of very thin films of high-temperature cuprate superconductors composed of a number N of only a few unit cells of superconducting bilayers. For that, we solve the fluctuation spectrum of a Gaussian-Ginzburg-Landau model for few-bilayers superconductors considering two alternating Josephson interlayer interaction strengths, and we obtain the corresponding paraconductivity above the transition. Then, we extend these calculations to temperatures below the transition through expressions for the Ginzburg number and Kosterlitz--Thouless-like critical region. When compared with previously available data in YBa2_2Cu3_3O7δ_{7-\delta} few-bilayers systems, with N=1 to 4, our results seem to provide a plausible scenario for their critical regime.
... Notably, the obtained scaling law near T c ∼ 0 is incompatible with the mean-field theory such as the dwave Bardeen-Cooper-Schrieffer theory even with the dirty limit, pointing to the local pairing and massive phase fluctuation in the real space [10]. Note that nematic fluctuation in Bi 2 Sr 2 CaCu 2 O 8+δ (Bi2212) [11] and reentrant charge order in Pb-Bi2201 [12] have been reported in the overdoped region. ...
Article
Full-text available
The real and imaginary parts of the self-energy in the wide energy range have been evaluated from high-resolution angle-resolved photoemission spectroscopy (ARPES) spectra of a heavily overdoped single-layer cuprate superconductor ( Bi , Pb ) 2 Sr 2 CuO 6 + δ in the normal state. The real part of the self-energy has the zero-point crossing at − 0.6 eV, where the lifetime broadening takes maximum and the ARPES spectral intensity is significantly suppressed forming a high-energy anomaly (HEA). The self-energy part responsible for the HEA is almost temperature independent and the coupling parameter is λ HEA ∼ 1 at 300 K, indicating strongly correlated nature of the strange metal state. Meanwhile, the self-energy part responsible for the low-energy kink (LEK) near the Fermi level shows significant temperature dependence: it is λ LEK ∼ 0.1 at 300 K but enhances steeply below ∼ 150 K up to λ LEK ∼ 0.8 , leading to the total coupling strength of λ tot = λ HEA + λ LEK = 1.8 at 20 K. The temperature-dependent enhancement cannot be explained by the electron-phonon interaction with a fixed magnitude of the Eliashberg function, which suggests an unexplored mechanism that enhances the coupling parameter at lower temperatures. Our results clearly indicate distinct energy scales in the self-energy, providing insight into the strange metal state as well as the temperature-dependent interplay of many-body interactions. Published by the American Physical Society 2025
... In this device, breaking C 4 symmetry is essential to generate the JDE. Besides the C 4 symmetry breaking in the s-d interlayer couplings due to the lattice deviates from standard square shape, it can also arise from the C 4 symmetry breaking in the d-wave pairing function, which is indicated in relative experimentally works [36,41,[66][67][68][69]. The JDE can also exist for the C 4 symmetry breaking in d-wave pairing function, ∆ d k = ∆ d (cos k x − cos k y ) + ∆ ds , with ∆ ds /∆ d represent the C 4 symmetry breaking [43]. ...
Preprint
Motivated by recent progress in both the Josephson diode effect (JDE) and the high-temperature Josephson junction, we propose to realize the JDE in an s-wave/d-wave/s-wave (s-d-s) superconductor junction and investigate the high-temperature superconducting order parameters. The interlayer coupling between s-wave and d-wave superconductors can induce an effective d+is superconducting state, spontaneously breaking time-reversal symmetry. The asymmetric s-d interlayer couplings break the inversion symmetry. Remarkably, the breaking of these two symmetries leads to a ϕ0\phi_0-junction but does not generate JDE. We find that the emergence of the JDE in this junction depends on the C4C_4 rotational symmetry of the system. Although breaking C4C_4 rotational symmetry does not affect time-reversal and inversion symmetries, it can control the magnitude and polarity of diode efficiency. Furthermore, we propose observing C4_{4} symmetry breaking controlled JDE through asymmetric Shapiro steps. Our work suggests a JDE mechanism that relies on high-temperature d-wave pairing, which could inversely contribute to a potential experimental method for detecting the unconventional pairing symmetry in superconductors.
... One of the most enigmatic phases of hole-doped cuprates is charge nematicity [1]. This long-ranged charge order has been investigated quite frequently in the past two decades [2][3][4][5][6][7][8][9][10][11][12][13][14][15][16][17][18]. It manifests itself as the spontaneous breaking, by the electronic structure, of the four-fold rotational symmetry of the CuO 2 unit cell [11] common to all cuprates. ...
Preprint
Full-text available
Recent scanning-tunneling microscopy on hole-doped Bi2_2Sr2_2CaCu2_2O8_8, one of the materials of the cuprate family, finds a long-range ordered spontaneous splitting of the energy levels of oxygen orbitals inside the CuO2_2 unit cells [S. Wang et al., Nat. Mat. 23, 492-498 (2024)]. This spontaneous intra-unit-cell orbital ordering, also known as electronic nematicity, breaks C4C_4 symmetry and is thought to arise from the Coulomb interaction (denoted by VppV_{pp}) between oxygen pxp_x and pyp_y electrons. In this work, we study the spontaneous emergence of electronic nematicity within the three-band Hubbard (aka the Emery-VSA model), using cluster dynamical mean field theory. This method incorporates short-range electronic correlations and gives us access to the density of states, a quantity that is directly probed in experiments. We argue that there is a delicate competition between VppV_{pp} and VpdV_{pd} (the latter being the Coulomb interaction between copper dx2y2d_{x^2-y^2} and oxygen px,yp_{x,y} electrons) that must be taken into account in order to find a Zhang-Rice singlet band well-resolved from the upper Hubbard band, and a splitting of the charge-transfer band (one of the signatures of charge nematicity) by roughly 50 meV, as observed recently.
... These fluctuations might give rise to non-Fermi liquid behaviors [27,28] and they might provide a pairing glue between electrons leading to superconductivity, and this becomes particularly relevant in the proximity of a nematic QCP [29][30][31][32]. The poorly-understood relation between nematicity and unconventional superconductivity explains the huge interest about the former phase: It might explain the origins of the latter [33][34][35][36]. Coming to a more concrete example, in several of the iron-based superconductors [37], one observes the follow-ing phenomenology: Below a critical temperature T nem , an Ising Z 2 symmetry is broken and the nematic character seems to have electronic origin [33]. ...
Preprint
Full-text available
Starting from a low-energy model for the band dispersion of the 2×22 \times 2 charge-ordered phase of the kagome metals AV3_3Sb5_5 (A= K, Rb, Cs), we show that nematicity can develop in this state driven either by 4×44 \times 4 charge fluctuations preemptive of a 1×41 \times 4 charge order (CO), or by an actual zero momentum d-wave charge Pomeranchuk instability (PI). We perform an analysis that starts from a Kohn-Luttinger theory in the particle-hole sector, which allows us to establish a criterion for the development of an attractive nematic channel near the onset of the 1×41 \times 4 CO and near the d-wave charge PI, respectively. We derive an effective charge-fermion model for the d-wave PI with a nematic susceptibility given via a random phase approximation (RPA) summation. By contrast, for the finite momentum CO, the RPA scheme breaks down and needs to be improved upon by including Aslamazov-Larkin contributions to the nematic pairing vertex. We then move to the derivation of the Ginzburg-Landau potentials for the 4×44 \times 4 CO and for the d-wave PI, and we obtain the corresponding expression for the nematic susceptibility at the nematic transition temperature T Tnem \sim \text{T}_\text{nem} in both cases. Our work establishes a relation between the nematicity observed in some of the iron-based superconductors, where the nematic phase might be driven by spin fluctuations, and the vanadium-based kagome metals, where charge fluctuations likely induce nematicity. The two microscopic mechanisms we propose for the stabilization of the nematic state in AV3_3Sb5_5 are distinguishable by diffusive scattering experiments, meaning that it is possible to gauge which of the two theories, if any, is the most likely to describe this phase. Both mechanisms might also be relevant for the recently discovered titanium-based family ATi3_3Sb5_5.
Article
Recent scanning-tunneling microscopy on hole-doped Bi 2 Sr 2 CaCu 2 O 8 , one of the materials of the cuprate family, finds a long-range ordered spontaneous splitting of the energy levels of oxygen orbitals inside the CuO 2 unit cells [S. Wang et al ., Nat. Mat. 23, 492–498 (2024)]. This spontaneous intra-unit-cell orbital ordering, also known as electronic nematicity, breaks C 4 symmetry and is thought to arise from the Coulomb interaction (denoted by V pp ) between oxygen p x and p y electrons. In this work, we study the spontaneous emergence of electronic nematicity within the three-band Hubbard [aka the Emery-VSA (Varma-Schmitt-Rink-Abrahams) model], using cluster dynamical mean-field theory. This method incorporates short-range electronic correlations and gives us access to the density of states, a quantity that is directly probed in experiments. We argue that there is a delicate competition between V pp and V pd (the latter being the Coulomb interaction between copper d x 2 − y 2 and oxygen p x , y electrons) that must be taken into account in order to find a Zhang-Rice singlet band well-resolved from the upper Hubbard band, and a splitting of the charge-transfer band (one of the signatures of charge nematicity) by roughly 50 meV, as observed recently.
Article
Motivated by the pseudogap state of the cuprates, we introduce the concept of an “exceptional” van Hove singularity that appears when a strong electron-electron interaction splits an otherwise simply connected Fermi surface into multiply connected pieces. The singularity describes the touching of two pieces of the split Fermi surface. We show that this singularity is proximate to a second-order van Hove singularity, which can be accessed by tuning a dispersion parameter. We argue that, in a wide class of cuprates, the endpoint of the pseudogap is accessed only by triggering the exceptional van Hove singularity. The resulting Lifshitz transition is characterized by enhanced specific heat and nematic susceptibility, as seen in experiments.
Article
Magnetic and bond-charge interactions can arise from the same microscopic interaction. Motivated by this observation, we compute magnetic and bond-charge excitation spectra on an equal footing by introducing a simple effective model on a square lattice, which describes antiferromagnetic and d-wave superconducting phases around half-filling on the electron-doped side. The magnetic excitation spectrum Imχ(q,ω) has strong weight around q=(π,π) in low energy, and its intensity map exhibits a pencil-tip-like shape in q−ω space. Around q=(0,0) magnetic excitations show a steep dispersion toward the (π,π) and (π,0) directions, which is very similar to a spin-wave dispersion although the system is nonmagnetic. Bond-charge excitations are characterized by four different symmetries and studied for all possible couplings. Bond-charge fluctuations with three different symmetries have large spectral weight around q=(π,π) in a relatively low-energy region and extend widely more than the magnetic excitation spectrum. The d-wave symmetry of bond-charge excitations also has sizable spectral weight along the direction (π/2,π/2)−(0,0)−(π/2,0) in a low-energy region and exhibits softening around q≈(0.5π,0), whereas no such softening is present in the other symmetries. These results capture the essential features observed in electron-doped cuprates and may motivate an experimental test of bond-charge excitations around q=(π,π) on top of the strong magnetic excitations there as well as additional softening in the d-wave channel in the (π,π)−(π/2,π/2) region at low temperatures near the magnetic phase. We extend the present analysis to the hole-doped side and highlight a contrast to the electron-doped side, which includes incommensurate correlations, electronic nematic correlations, and spin and bond-charge resonance modes in the superconducting state.
Article
Full-text available
Charge density modulations have been observed in all families of high-critical temperature (Tc) superconducting cuprates. Although they are consistently found in the underdoped region of the phase diagram and at relatively low temperatures, it is still unclear to what extent they influence the unusual properties of these systems. Using resonant x-ray scattering, we carefully determined the temperature dependence of charge density modulations in YBa2Cu3O7-δ and Nd1+ x Ba2- x Cu3O7-δ for several doping levels. We isolated short-range dynamical charge density fluctuations in addition to the previously known quasi-critical charge density waves. They persist up to well above the pseudogap temperature T*, are characterized by energies of a few milli-electron volts, and pervade a large area of the phase diagram.
Article
Full-text available
The pseudogap phenomenon in the cuprates is arguably the most mysterious puzzle in the field of high-temperature superconductivity. The tetragonal cuprate HgBa2CuO4+δ, with only one CuO2 layer per primitive cell, is an ideal system to tackle this puzzle. Here, we measure the magnetic susceptibility anisotropy within the CuO2 plane with exceptionally high-precision magnetic torque experiments. Our key finding is that a distinct two-fold in-plane anisotropy sets in below the pseudogap temperature T*, which provides thermodynamic evidence for a nematic phase transition with broken four-fold symmetry. Surprisingly, the nematic director orients along the diagonal direction of the CuO2 square lattice, in sharp contrast to the bond nematicity along the Cu-O-Cu direction. Another remarkable feature is that the enhancement of the diagonal nematicity with decreasing temperature is suppressed around the temperature at which short-range charge-density-wave formation occurs. Our result suggests a competing relationship between diagonal nematic and charge-density-wave order in HgBa2CuO4+δ.
Article
Full-text available
Motivated by recent attention to a potential antiferromagnetic quantum critical point at xc∼0.19, we have used inelastic neutron scattering to investigate the low-energy spin excitations in crystals of La2−xSrxCuO4 bracketing xc. We observe a peak in the normal-state spin-fluctuation weight at ∼20meV for both x=0.21 and 0.17, inconsistent with quantum critical behavior. The presence of the peak raises the question of whether low-energy spin fluctuations limit the onset of superconducting order. Empirically evaluating the spin gap Δspin in the superconducting state, we find that Δspin is equal to the coherent superconducting gap Δc determined by electronic spectroscopies. To test whether this is a general result for other cuprate families, we have checked through the literature and find that Δc≤Δspin for cuprates with uniform d-wave superconductivity. We discuss the implications of this result.
Article
Full-text available
The three central phenomena of cuprate superconductors are linked by a common doping pp^{\star}, where the enigmatic pseudogap phase ends, around which the superconducting phase forms a dome, and at which the resistivity exhibits an anomalous linear dependence on temperature as T0T \to 0. However, the fundamental nature of pp^{\star} remains unclear, in particular whether it marks a true quantum phase transition. We have measured the specific heat C of the cuprates Eu-LSCO and Nd-LSCO at low temperature in magnetic fields large enough to suppress superconductivity, over a wide doping range across pp^{\star}. As a function of doping, we find that the electronic term CelC_{\rm el} is strongly peaked at pp^{\star}, where it exhibits a T-TlogT dependence as T0T \to 0. These are the classic signatures of a quantum critical point, as observed in heavy-fermion and iron-based superconductors where their antiferromagnetic phase ends. We conclude that the pseudogap phase of cuprates ends at a quantum critical point, whose associated fluctuations are most likely involved in the d-wave pairing and the anomalous scattering.
Article
Full-text available
One of the distinctive features of hole-doped cuprate superconductors is the onset of a `pseudogap' below a temperature TT^*. Recent experiments suggest that there may be a connection between the existence of the pseudogap and the topology of the Fermi surface. Here, we address this issue by studying the two-dimensional Hubbard model. We show that electronic correlations can strongly modify the Fermi surface as compared to its non-interacting shape. We find that the pseudogap only exists when the Fermi surface is hole-like and that, for a broad range of parameters, its opening is concomitant with a Fermi surface topology change from electron- to hole-like. The pseudogap is shown to be associated with a pole-like feature in the electronic self-energy, which also controls the location of the Fermi surface topology transition through particle-hole asymmetry, hence providing a common link between these observations. We discuss the relevance of our results to experiments on cuprates.
Article
Full-text available
We study the relationship between the pseudogap and Fermi-surface topology in the two-dimensional Hubbard model by means of the cellular dynamical mean-field theory. We find two possible mean-field metallic solutions on a broad range of interaction, doping and frustration: a conventional renormalized metal and an unconventional pseudogap metal. At half-filling, the conventional metal is more stable and displays an interaction-driven Mott metal-insulator transition. However, for large interaction and small doping, region that is relevant for cuprates, the pseudogap phase becomes the ground state. By increasing doping, we show that a first-order transition from the pseudogap to the conventional metal is tight to a change of the Fermi surface from hole to electron like, unveiling a correlation-driven mechanism for a Lifshitz transition. This explains the puzzling link between pseudogap phase and Fermi surface topology which has been pointed out in recent experiments.
Article
We study the behavior of the pseudogap in overdoped Bi2Sr2CaCu2O8+δ by electronic Raman scattering (ERS) and angle-resolved photoemission spectroscopy (ARPES) on the same single crystals. Using both techniques we find that, unlike the superconducting gap, the pseudogap related to the antibonding band vanishes above the critical doping pc=0.22. Concomitantly, we show from ARPES measurements that the Fermi surface of the antibonding band is holelike below pc and becomes electronlike above pc. This reveals that the existence of the pseudogap depends on the Fermi surface topology in Bi2Sr2CaCu2O8+δ, and more generally, puts strong constraint on theories of the pseudogap phase.
Article
The origin of high-temperature superconductivity in copper oxides and the nature of the 'normal' state above the critical temperature are widely debated. In underdoped copper oxides, this normal state hosts a pseudogap and other anomalous features; and in the overdoped materials, the standard Bardeen-Cooper-Schrieffer description fails, challenging the idea that the normal state is a simple Fermi liquid. To investigate these questions, we have studied the behaviour of single-crystal La(2-x)Sr(x)CuO"4 films through which an electrical current is being passed. Here we report that a spontaneous voltage develops across the sample, transverse (orthogonal) to the electrical current. The dependence of this voltage on probe current, temperature, in-plane device orientation and doping shows that this behaviour is intrinsic, substantial, robust and present over a broad range of temperature and doping. If the current direction is rotated in-plane by an angle Φ, the transverse voltage oscillates as sin(2Φ), breaking the four-fold rotational symmetry of the crystal. The amplitude of the oscillations is strongly peaked near the critical temperature for superconductivity and decreases with increasing doping. We find that these phenomena are manifestations of unexpected in-plane anisotropy in the electronic transport. The films are very thin and epitaxially constrained to be tetragonal (that is, with four-fold symmetry), so one expects a constant resistivity and zero transverse voltage, for every Φ. The origin of this anisotropy is purely electronic - the so-called electronic nematicity. Unusually, the nematic director is not aligned with the crystal axes, unless a substantial orthorhombic distortion is imposed. The fact that this anisotropy occurs in a material that exhibits high-temperature superconductivity may not be a coincidence. © 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Article
A central issue in the quest to understand the superconductivity in cuprates is the nature and origin of the pseudogap state, which harbours anomalous electronic states such as Fermi arc, charge density wave (CDW), and d-wave superconductivity. A fundamentally important, but long-standing controversial problem has been whether the pseudogap state is a distinct thermodynamic phase characterized by broken symmetries below the onset temperature TT^*. Electronic nematicity, a fourfold (C4C_4) rotational symmetry breaking, has emerged as a key feature inside the pseudogap regime, but the presence or absence of a nematic phase transition and its relationship to the pseudogap remain unresolved. Here we report thermodynamic measurements of magnetic torque in the underdoped regime of orthorhombic YBa2_2Cu3_3Oy_y with a field rotating in the CuO2_2 plane, which allow us to quantify magnetic anisotropy with exceptionally high precision. Upon entering the pseudogap regime, the in-plane anisotropy of magnetic susceptibility increases after exhibiting a distinct kink at TT^*. Our doping dependence analysis reveals that this anisotropy is preserved below TT^* even in the limit where the effect of orthorhombicity is eliminated. In addition, the excess in-plane anisotropy data show a remarkable scaling behaviour with respect to T/TT/T^* in a wide doping range. These results provide thermodynamic evidence that the pseudogap onset is associated with a second-order nematic phase transition, which is distinct from the CDW transition that accompanies translational symmetry breaking. This suggests that nematic fluctuations near the pseudogap phase boundary have a potential link to the strange metallic behaviour in the normal state, out of which high-TcT_c superconductivity emerges.