This is the html version of the file https://arxiv.org/abs/cond-mat/0609250.
Google automatically generates html versions of documents as we crawl the web.
arXiv:cond-mat/0609250v1 [cond-mat.str-el] 11 Sep 2006
  Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Page 1
arXiv:cond-mat/0609250v1 [cond-mat.str-el] 11 Sep 2006
Tl2Ba2CuO6+δ Brings Spectroscopic Probes Deep
Into the Overdoped Regime of the High-Tc Cuprates
DC Peets, JDF Mottershead, B Wu, IS Elfimov, R Liang, WN
Hardy, DA Bonn, M Raudsepp, NJC Ingle and A Damascelli
Department of Physics and Astronomy, University of British Columbia,
Vancouver, BC V6T 1Z1 Canada
† Department of Earth and Ocean Sciences, University of British Columbia,
Vancouver, BC V6T 1Z4 Canada
E-mail: bonn@physics.ubc.ca and damascelli@physics.ubc.ca
Abstract. Single-particle spectroscopic probes, such as scanning tunneling and
angle-resolved photoemission spectroscopy (ARPES), have provided us with crucial
insights into the complex electronic structure of the high-Tc cuprates, in particular
for the under and optimally doped regimes where high-quality crystals suitable for
surface-sensitive experiments are available. Conversely, the elementary excitations on
the heavily overdoped side of the phase diagram remain largely unexplored. Important
breakthroughs could come from the study of Tl2Ba2CuO6+δ (Tl2201), a structurally
simple system whose doping level can be tuned from optimal to extreme overdoping
by varying the oxygen content. Using a self-flux method and encapsulation, we have
grown single crystals of Tl2201, which were then carefully annealed under controlled
oxygen partial pressures. Their high quality and homogeneity are demonstrated by
narrow rocking curves and superconducting transition widths. For higher dopings, the
crystals are orthorhombic, a lattice distortion stabilized by O interstitials in the TlO
layer. These crystals have enabled the first successful ARPES study of both normal
and superconducting-state electronic structure in Tl2201, allowing a direct comparison
with the Fermi surface from magnetoresistance and the gap from thermal conductivity
experiments. This establishes Tl2201 as the first high-Tc cuprate for which a surface-
sensitive single-particle spectroscopy and a comparable bulk transport technique have
arrived at quantitative agreement on a major feature such as the normal state Fermi
surface. The momentum dependence of the ARPES lineshape reveals, however, an
unexpected phenomenology: in contrast to the case of under- and optimally-doped
cuprates, quasiparticles are sharp near (π, 0), the antinodal region where the gap is
maximum, and broad at (π/2, π/2), the nodal region where the gap vanishes. This
reversed quasiparticle anisotropy past optimal doping, and its relevance to scattering,
many-body, and quantum-critical phenomena in the high-Tc cuprates, are discussed.
Submitted to New Journal of Physics (September 4, 2006).

Page 2
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
2
1. Introduction
The cuprate superconductors can be tuned through a remarkable progression of states
of matter by doping charge carriers into the CuO2 planes [1]. The most generic feature
of this tuning is a sequence from an antiferromagnetic insulator, through a doping range
where the superconducting critical temperature builds to a maximum (optimal doping)
and then dies away again at higher doping, in the so-called overdoped regime (Fig.1).
This sequence contains three separate touchstones for an understanding of the cuprates:
the Mott insulator, which is known to result from very strong electron-electron repulsion
and is antiferromagnetically ordered, the d-wave superconductor, and the overdoped
metal. The latter is widely believed to exhibit less exotic normal-state properties
and might be understood through Fermi liquid theory. Over the last two decades, a
great deal of experimental work has been done on undoped, underdoped, and optimally
doped cuprates, aiming at elucidating the connection between the antiferromagnetic
insulator and high-Tc superconductor (HTSC), and the role of electronic correlations in
the emergence of superconductivity. However, the testing of Fermi liquid theory in the
overdoped regime has been severely hampered by a lack of compounds that can actually
be doped to this high level and are suitable for a wide range of experimental techniques.
In the struggle to understand high-Tc superconductivity, the quest for better
materials to study remains key. Researchers want samples that can be grown
very cleanly, can be doped over a very wide range of the phase diagram, and are
suitable for a variety of bulk and surface measurements. Surface-sensitive techniques
in particular, such as angle-resolved photoemission spectroscopy (ARPES) [2, 3]
and scanning tunneling microscopy (STM) and spectroscopy (STS) [4], introduce
additional constraints with regard to sample requirements. These highly sophisticated
spectroscopic probes can be used to obtain information on the single-particle excitation
spectrum of a solid — i.e. the spectral function A(k,ω) that can be calculated in terms
of the electron Green’s function starting from the microscopic many-body Hamiltonian
of the system [5] — but only if the material can be cleaved to expose a well-defined and
stable surface (i.e. which does not reconstruct, obscuring the bulk electronic structure).
A summary of the approximate doping range explored by ARPES on different HTSC
families is given in Fig.1. On the undoped side of the phase diagram (p=0), the parent
compounds Ca2CuO2Cl2 and Sr2CuO2Cl2 yield excellent cleaved [001] faces that have
allowed ARPES studies of the electronic structure of the correlated antiferromagnetic
insulator and of the magneto-elastic interactions associated with the propagation of a
single hole in the two-dimensional spin background [6, 7, 8]. Via hole doping, one can
drive these materials through a metal-insulator transition and enter the superconducting
dome (in this paper we will focus on the hole-doped side of the phase diagram).
For instance, recent STM, STS [9], and ARPES [10] experiments on the underdoped
superconductor Ca2−xNaxCuO2Cl2 have suggested the existence of new electronic
ordering phenomena in the doping region between the insulator and the superconductor,
possibly in the form of glassy electronic behavior or a surface-nucleated phase transition

Page 3
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
3
Figure 1. Generic temperature-doping phase diagram for both electron (p < 0) and
hole-doped (p>0) cuprates. The doping range actually explored by ARPES for those
material families more extensively studied is indicated by the corresponding arrows.
as more conventional long-range charge order was not observed in scattering studies [11].
Near optimal doping, where the cuprates exhibit their highest superconducting critical
temperatures, YBa2Cu3O7−δ stands out as the cleanest material and has been used
extensively in bulk-sensitive studies of the normal and superconducting properties, such
as the symmetry of the order parameter [12, 13, 14], superfluid density [15], charge and
thermal transport [16], low-energy electrodynamics [17, 18], vibrational and magnetic
excitation spectra [19, 20]. Unfortunately, this material is complicated by the presence
of CuO-chain layers and does not have a neutral [001] cleavage plane for surface-sensitive
techniques. In particular, YBa2Cu3O7−δ cleaves between the chain layer and the BaO
layer. BaO surfaces have had the nearest source of dopants (the CuO chains) removed,
so that their hole-doping is uncertain. STS shows that the CuO chain surfaces are
characterized by prominent surface density waves [21] and differ substantially from the
bulk. They exhibit surface states, as seen in ARPES [22], and possess unavoidable
doping disorder. These problems have severely limited the availability of single-particle
spectroscopy data from ARPES [23, 24], STM, and STS [21, 25].
Extensive ARPES investigations have been performed on La2−xSrxCuO4 [26], over
a broad doping range extending from the undoped insulator (x = 0) to the overdoped
superconductor (x=0.22; note that x=p for this compound). This system, however, is
affected by severe cation disorder associated with La-Sr substitution right next to the
CuO2 plane, as well as lattice, charge, and magnetic instabilities, which might provide
an explanation for why the maximum Tc in this compound is suppressed with respect to
other single-layer materials [27]. In turn, many of the key bulk-sensitive measurements
requiring perfect crystals and long mean free paths could not be performed for this

Page 4
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
4
material family. In addition, neither STM nor STS experiments have been successful so
far on cleaved La2−xSrxCuO4, probably because both the CuO2 and La1− x
2 Srx
2 O surfaces
are polar, with a charge zCuO2 = −(2−x) and z(La,Sr)O = (1−x/2), and thus critically
unstable. To date, the most rich and comprehensive spectroscopic information on high-
Tc cuprates from both STS and ARPES have been obtained on the Bi compounds,
and in particular Bi2Sr2CuO6+δ and Bi2Sr2CaCu2O8+δ [2, 3, 4]. However, similar to
La2−xSrxCuO4, this family has greater problems with disorder than YBa2Cu3O7−δ [27],
leaving a disconnection between materials for which we have the most information. This
dilemma interferes with strict tests of the theory of high-Tc superconductivity, in which
one must connect the single particle excitations to the other physical properties and
dynamic susceptibilities of the cuprates [28].
A further problem in the study of HTSCs is that, due to the high Tc values, the
physical properties of the underlying normal state can be probed only at relatively high
temperatures. This often makes the detailed interpretation of the experimental results
more complex and less informative (e.g., charge and thermal conductivity, Hall number,
magnetoresistance, etc.), or can even make some experiments completely unfeasible
(as in the case of the de Haas-van Alphen effect). Alternatively, at least for certain
techniques, one can access the low-temperature normal-state properties by suppressing
superconductivity via the application of an external magnetic field. Once again however,
due to the very high Tc and in turn the extraordinarily large value of the second critical
field Hc2, this approach is precluded near optimal doping on the cleanest cuprates.
Extreme overdoping is a means of reducing both Tc and Hc2, so that one can study both
the superconducting and underlying normal states, but here most samples tend to be less
clean because of the need to dope with large concentrations of cation impurities, such
as Sr in La2CuO4 and Pb in Bi2Sr2CuO6+δ [27]. In this case it can be unclear whether
changes observed in the material’s properties are attributable to the increased doping,
or the increased scattering and/or other extrinsic effects. Due to these material and
experimental limitations, the underlying normal state of HTSCs, the heavily overdoped
side of the HTSC phase diagram, and the metallic state that emerges beyond the end
of the superconducting dome have remained so far largely unexplored (Fig.1).
The most promising compound for investigating the deeply overdoped regime is
the single-layer system Tl2Ba2CuO6+δ (Tl2201, see Fig. 2), whose natural doping range
varies from optimal to extreme overdoping as one increases the oxygen content δ.
Tl2201 has two extremely flat CuO2 planes located far apart from each other in the
body-centered unit cell; there are no complicating chain layers (as opposed to the
Y-based cuprates), or bilayer band-splitting effects (as opposed to Bi2Sr2CaCu2O8+δ
and YBa2Cu3O7−δ). Like La and Bi-based systems, Tl2201 can exhibit a cation non-
stoichiometry but this consists in excess Cu substituting Tl in the Tl2O2 double layer,
farther removed from the CuO2 planes than the Bi-Sr or La-Sr substitutions that have
been shown to have a large impact on the superconductivity of the Bi and La cuprates
[27]. Tl2201 may be reversibly doped through interstitial oxygen and, near optimal
doping, vacancies in the Tl2O2 bilayer [29]. The ability to overdope a crystal reversibly

Page 5
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
5
Figure 2. (a) Unit cell of Tl2201. (b,c) Electrical resistivity and thermal conductivity
of overdoped Tl2201 (Tc ≃ 15 K), plotted vs. temperature for different values of an
external magnetic field applied normal to the CuO2 planes (after Ref. [34]). From the
T =0 K intercepts of ρ(T) and κ(T)/T, the ratio ρ0κ0/T =0.99±0.01L0 was calculated
(L0 =2.44 × 10−8 WΩK2 is Sommerfeld’s value for the Lorenz ratio L≡κ/σT), which
indicates that the Wiedemann-Franz law is obeyed in overdoped cuprates although the
normal-state ρ(T) is dominated by a non Fermi-liquid like T-linear term (see inset).
from optimal doping to the full suppression of superconductivity, by changing its oxygen
content alone, is a fortuitous property of Tl2201, although this has been successfully
demonstrated over the whole doping range only on ceramic samples.
A variety of bulk properties have already been studied in Tl2201. At optimal
doping, the pure dx2−y2 symmetry of the superconducting gap was confirmed by phase-
sensitive measurements [30], and the so-called magnetic resonant mode at 47.5 meV was
detected below Tc in neutron scattering experiments [31], the first time for a single-layer
cuprate. The high sample quality has been recently demonstrated by measurements
of angular magnetoresistance oscillations (AMRO), a bulk transport technique that
requires the long mean free paths afforded by high purity and crystallinity [32]. The
AMRO study of Tl2201 resulted in the first estimate of the normal state Fermi surface
of a cuprate superconductor from a bulk sensitive probe. The combination of charge
[33, 34] and heat [34] transport measured at low temperature on very overdoped Tl2201
in the normal state, by applying up to 13T magnetic fields to suppress superconductivity
(see Fig.2), allowed the Lorenz ratio L ≡ κ/σT to be extracted. The value obtained,
0.99 ± 0.01 L0, allowed one to conclude that the Wiedemann-Franz law is obeyed in
overdoped cuprates: the electronic carriers of heat are fermionic excitations of charge
−e, suggesting that the low-energy excitations in overdoped cuprates might indeed be
described in terms of Fermi-liquid quasiparticles [34]. At variance with the hints of
Fermi-liquid-like behavior, however, the ab-plane resistivity is not purely described by
a T2 power law but is still dominated by a T-linear term (see inset of Fig 2). In
this context, the detailed study of Tl2201 with modern single particle spectroscopies
would be extremely desirable, as it might provide direct insights into the nature of the

Page 6
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
6
quasiparticles in both the normal and superconducting states [2, 3, 5].
Unfortunately, despite its great potential, progress on the Tl2201 system has been
slow because of difficulties in the growth of single crystals. The thallium oxide is volatile,
reactive, and highly toxic, a combination which places constraints on the crystal growth
technique. Standard growth procedures such as flux-growth in open crucibles or the use
of a floating zone image furnace would allow toxic vapors to escape into the lab and
deplete the thallium in the reaction vessel by allowing it to sublime out in a colder part of
the furnace. It is thus essential to contain the growth components within a non-reactive
enclosure so that thallium is not lost through either reaction or evaporation.
In Section 2, we report on our effort in the growth of high-purity single crystals of
Tl2201, by a copper-rich self-flux method, and in their careful annealing in a controlled
oxygen atmosphere. We will show that this material can be cleaved to expose a
clean and stable [001] surface suitable for single particle spectroscopy experiments, as
was demonstrated by the first successful study of the Fermi surface and quasiparticle
excitations of Tl2201 by ARPES [35]. The detailed results of our ARPES study of several
doping levels will be presented in Section 3, and a discussion linked to other theoretical
and experimental results will be given in Section 4. The quantitative agreement between
the ARPES and AMRO [32] determined Fermi surfaces indicates that Tl2201 may be the
ideal testing ground for finally joining together the modern single particle spectroscopies
and a host of well-established bulk probes of metals and superconductors. In particular,
Tl2201 might be the ideal cuprate for a broad-based study of the normal metal that
may underlie the superconductor in the overdoped regime, providing an understandable
anchor point in the phase diagram of the high-temperature superconductors.
2. Tl2201 Single-Crystal Preparation
Difficult to prepare even as a ceramic, Tl2201 is even more challenging to grow as a
crystal. Tl2201 melts incongruently and its crystals are grown from the high temperature
solution (melt) made of its components. Ideally, this melt would be decanted to reveal
freestanding flux-free crystals. While this technique is commonly used for growth of
complex oxide crystals, there are a few challenges particular to Tl2201 system. The
difficulties are due to the thallium oxide precursor. Around 800C, trivalent thallium
oxide decomposes into a monovalent oxide and oxygen gas [36]:
Tl2O3 ⇋ Tl2O+O2 ↑ .
While the oxygen gas evolved must be allowed to escape, this decomposition leads
to a more serious problem: at growth temperatures, the monovalent oxide’s vapor
pressure is of the order of 0.1atm, high enough to allow it to escape rapidly. If the
Tl2O gas leaves the crucible, however slowly, the composition of the melt will be at
best time-varying and at worst quickly devoid of thallium. The volatility of Tl2O can
also make the subsequent annealing of the crystals difficult, as the crystal surface may
slowly decompose at temperatures at which the oxygen atoms are mobile. There are

Page 7
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
7
several procedures that may be employed to mitigate these problems, including using
pre-reduced thallium precursors such as Tl2Ba2O5 [37], encapsulation [38], or very high
oxygen partial pressures. A further issue is that Tl2O is quite reactive, most notably
reacting with the quartz often used for encapsulation and furnace tubes.
To date, only a handful of groups have grown crystals of Tl2201. Mackenzie’s
group, using a gold lid on their crucible as a barrier to diffusion, found they could
only successfully obtain crystals within a 2C by 2 minute window [39, 40]. Hasegawa
et al. used reduced thallium precursors and an alumina high-pressure bomb to control
thallium loss [41, 38]. Kolesnikov et al. [42, 43, 44] do not report any special precautions
for controlling the loss of thallium. In our growth effort, as we will discuss in detail in
Section 2.1, we developed a flexible sealing scheme that permits the oxygen to escape
as it is evolved, while at the same time containing the Tl2O.
A last crucial issue with the preparation of Tl2201 single crystals, alluded to above,
is that the crystal surfaces can degrade during the oxygen post-annealing required to set
the charge carrier content, and in turn the temperature of the superconducting phase
transition. So far most work on annealing of Tl2201 has been carried out on ceramic
samples [45], but annealing of single crystals, where surface damage is more obvious,
has been less successful — only Mackenzie’s group reported success [39]. As a result,
many crystals studied to date have been unannealed, resulting in inhomogeneous oxygen
content and broad transitions (∆Tc =10 ∼20K), which can obscure many features and
add unnecessary disorder. As part of our growth effort, we have also carefully explored
the annealing of Tl2201 single crystals, succeeding in the preparation of samples ranging
from near optimal doping to very overdoped, while avoiding surface damage.
2.1. Single-Crystal Growth
For the crystal growth we employ a copper-rich self-flux technique using a barium and
copper precursor, BaCuO2 powder that was prepared from 99.999% pure BaCO3 and
99.995% pure CuO by repeated calcining under flowing oxygen. The use of carbonate-
free barium cuprate avoids both carbon poisoning from BaCO3 and the cation impurities
that may be encountered if one employs BaO2, which is only available at lower cation
purity. Carbon contamination is of particular concern in this system because there are
several known superconducting thallium-based oxycarbonates [46].
The barium cuprate powder was mixed with 99.99% pure Tl2O3 to a final cation
ratio of Tl:Ba:Cu = 1.05:1:1 [47], then loaded into a 10mL alumina crucible. The
crucible was sealed using a gold lid fixed in place with a 1kg weight. Employing this
sealing scheme, O2 gas may still escape as Tl2O3 decomposes, but the seal becomes
very effective once the contents of the crucible equilibrate at the growth temperature,
thus preventing significant loss of Tl2O (as monitored by mass-loss measurements). The
charge was heated rapidly (300C/h) to 935C, while flowing oxygen to flush out any
escaping thallium oxide, and held for 12h to equilibrate. It was cooled at −0.5C/h
to 890C, then allowed to cool freely to room temperature. The oxygen gas flowing

Page 8
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
8
Figure 3. (a) Photo of several resulting Tl2201 crystals on millimeter graph paper; the
orthorhombic a-axis of one crystal is marked. (b) Field-cooled magnetization curves for
representative Tl2201 crystals annealed to different hole-doping levels; the sharpness
of the superconducting transitions is an indication of the crystals’ homogeneity.
through the furnace was bubbled through sulphuric acid and water to remove any
thallium released during growth that did not condense in colder areas of the furnace.
Most resulting crystals were embedded in flux, but free-standing platelet single crystals
of 1∼2 mm2 were harvested from voids in the flux ingot. As shown in Fig.3(a), these
crystals exhibited mirror surfaces with fine curved growth steps. All characterization
and spectroscopy reported here was performed on flux-free, void-grown single crystals.
The as-grown samples typically have superconducting transitions of 5∼10 K (onset)
and several Kelvins wide, a width comparable to other groups’ as-grown samples. As-
grown Tcs are determined by the cation substitution level, crystal dimensions, cooling
rate, and the atmosphere around the crystals, the last two parameters being coupled
through the equation of state for oxygen gas. Since as-grown superconducting transitions
are typically quite broad and not homogeneous throughout the crystal, it is crucial that
the crystals be annealed before they are studied. The crystals were annealed under
controlled oxygen partial pressures at temperatures between 290C and 500C. This
produces samples whose oxygen contents place them on the overdoped side of the phase
diagram, with sharp superconducting transitions ranging from 5 to 85K. We have not
attempted to reach optimal doping, which in this system is believed to correspond to
a superconducting Tc of at least 93 K [29], because even higher annealing temperatures
would be required, increasing the risk of thallium loss.
2.2. Physical and Chemical Analysis
Since cation substitution can cause increased scattering, one objective of this work was
to grow crystals with the lowest possible cation substitution level. Of more immediate
importance than the reduction of cation defects, however, is reproducibility. Because
cation substitution also dopes the crystal, variation between crystals would lead to

Page 9
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
9
Figure 4.
(a) The (0 0 10) rocking curve for a typical Tl2201 single crystal
(Tc=67.7K); the FWHM is 0.034, indicating excellent crystallinity. (b) The (6 0 0)
and (0 6 0) peaks for a Tc =24K sample, showing the orthorhombicity of the crystal
structure. The doublets correspond to CuKα1 and CuKα2 radiation results.
different Tcs as well as different scattering rates. More problematic still would be a
variation within a crystal, which would arise from changes in the composition of the
melt over the course of the crystal’s growth. Although cation substitution is not desired
and does dope the crystals, it must be emphasized that this is not used as a control
parameter in doping these crystals. If the crystals can be grown with a reproducible
level of cation substitution, the cation disorder will be the same for all dopings. This is
in contrast to LSCO, where cation substitution is the primary control parameter, and
must be varied by roughly a factor of four to traverse the superconducting dome.
To determine the levels of cation substitution in our crystals, electron-probe micro-
analyses (EPMA) of several crystals were performed on a fully-automated CAMECA
SX-50 instrument in wavelength-dispersion mode.‡ The cation composition (normalized
to the barium content) was determined to be Tl1.884(6)Ba2Cu1.11(1)O6+δ and was not
observed to be position dependent on individual crystals, nor did it vary between
crystals. This result corresponds to a level of copper substitution at the thallium site
comparable to that reported previously [49, 39, 40, 44] for cation-substituted Tl2201.
Al contamination was checked separately by EPMA and found to be below the 50 ppm
detection limit, corroborating our observation that the crucibles were not corroded by
the melt.§ Despite the cation non-stoichiometry, the high quality of these crystals is
‡ The EPMA was performed under the following operating conditions: excitation voltage, 15kV;
beam current, 20nA; peak count time, 80s; background count time, 40s; and spot diameter, 10 µm.
Data reduction was performed using the “PAP” ϕ(ρZ) method [48]. For the elements studied, the
following standards, X-ray lines and monochromator crystals were used: elemental Tl, TlMα, PET;
YBa2Cu3O6.920, BaLα, PET; and YBa2Cu3O6.920, CuKα, LIF. Tight Pulse Height Analysis (PHA)
control was used to eliminate to the degree possible any interference from higher-order lines.
§ Operating conditions were the following: excitation voltage, 15kV; beam current, 40nA; peak count
time, 500 s; background count time, 250 s; and spot diameter, 10 µm. The matrix composition was fixed
at stoichiometric; AlKα was standardized on α-Al2O3 using a TAP crystal monochromator.

Page 10
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
10
Figure 5. Schematic depiction of the dependence of the orthorhombic/tetragonal
structural transition on y and δ in Tl2−yBa2Cu1+yO6+δ. This scenario is suggested by
the data from polycrystalline samples of Wagner et al. [29] and our own single crystal
diffraction results (see Fig. 6). The CuO2 plane and unit cell (in green) are shown with
and without the orthorhombic distortion, which is here exaggerated for clarity. The
gray bar indicates where our crystals fall on this plot: y=0.116.
evidenced by their narrow superconducting transitions, measured by magnetization in
a Quantum Design SQUID magnetometer (MPMS). Fig.3(b) shows the field-cooled
magnetization measured at 1 Oe (H c), for crystals with transition widths (10%-90%)
ranging from 4 to 0.7K. The observed 40∼60% Meissner fraction (i.e., fraction of flux
excluded when cooling through Tc) compares well with field-cooled results on other
superconducting cuprates. It should also be mentioned that Tl2201’s extremely low
first critical field (Hc1) makes the transitions appear broader in higher fields.
2.3. Structural Analysis: Crystallinity and Symmetry
To investigate the samples’ crystallinity, X-ray diffraction spectra and rocking curves
were taken on a BEDE model 200 double-crystal diffractometer with a Si(111)
monochromator; the entire sample was illuminated by the X-ray beam. Fig.4(a)
depicts the (0 0 10) X-ray rocking curve of a 0.5×1.2 mm2 crystal annealed to
Tc=67.7K. While the FWHM of 0.034is broader than the 0.006obtained for the
best YBa2Cu3O7−δ crystals grown in BaZrO3 crucibles, it is similar to the 0.02-0.03
achieved for YBa2Cu3O7−δ grown in YSZ crucibles [50], indicating a good degree of
crystalline perfection. A crystal annealed to achieve a relatively high oxygen content

Page 11
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
11
Figure 6. Comparison of the orthorhombicity between our crystals and Wagner’s [29]
and Shimakawa’s [51] ceramics: orthorhombicity and c-axis lattice constant vs. Tc.
and strong overdoping (Tc=24 K) was found to be orthorhombic with lattice parameters
a=5.4580(3) Å, b=5.4848(5)Å and c=23.2014(5) Å, consistent with the lattice constants
reported for ceramics with similar Tc and Cu substitution [29]. This orthorhombicity is
determined by the x-ray diffraction data presented in Fig.4(b): the two data sets were
recorded independently for a and b axes and would be identical to one another if the
crystals were tetragonal. Closer to optimal doping, and at a lower interstitial oxygen
content, the Tc=67.7 K sample was tetragonal to within our resolution
Tl2201’s crystal symmetry depends on the level of cation substitution and the
oxygen content. Stoichiometric and near-stoichiometric ceramics have been shown to
be orthorhombic for all oxygen contents, while heavily-substituted ones are strictly
tetragonal [51]. For intermediate levels of substitution, the orthorhombicity depends on
the oxygen content: samples with higher oxygen contents are more orthorhombic [29].
The orthorhombic distortion is thought to arise from a lattice mismatch between the
CuO2 plane layer and the naturally larger Tl2O2 double layer. Interstitial oxygens
increase the distortion, while cation disorder seems to suppress it. The distortion
and its dependence on the oxygen content δ and the copper substitution y are shown
schematically in Fig. 5. Powder X-ray diffraction shows that the orthorhombic distortion
is an elongation along one plaquette diagonal, with the Cu-O bonds remaining the same
in both directions. The distortion is quite minor: instead of being at right angles,
the Cu-O bonds meet at ≈89.7in orthorhombic samples. The plaquette diagonals,
identified with nodes in the superconducting gap, remain orthogonal, so no mixing of
order parameter symmetries is required. For these reasons and for ease of comparison
with other systems, all ARPES analysis reported here is based on the original tetragonal
unit cell, regardless of the actual symmetry of the crystals.
Our crystals’ orthorhombicity, defined as 2(b − a)/(b + a), is consistent with that
found by Wagner et al. [29] for similar levels of cation substitution, as shown in Fig. 6(a).
One should keep in mind, however, that this quantity is highly prone to uncertainties

Page 12
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
12
because of the smallness of the orthorhombic distortion, so that quantitative conclusions
based on this graph should be treated with caution. Fig.6(b) compares two of our
annealed crystals’ c-axis lattice parameters to the orthorhombic ceramic samples of
Wagner et al. [29] and Shimakawa et al. [51]. Our c-axis lattice parameters are again
similar to those of Wagner et al. [29], but we observe higher Tc values for the same
lattice parameters (or a shorter c axis for the same Tc). Possible explanations for this
include improvements in Tc from higher homogeneity, cleaner samples or larger grain
size (crystals vs. ceramics), more homogeneous annealing, or possibly a calibration
disagreement between the diffractometers. Wagner’s samples have transitions a few
Kelvins wide, with long tails towards zero Kelvin, and Tc was defined using the 50%
point, which may not be an accurate average of the bulk in such a situation.
3. The Low-Energy Electronic Structure of Tl2201
Angle-resolved photoemission spectroscopy (ARPES) is an important tool in the
study of the electron dynamics in correlated-electron systems. Within the sudden
approximation, it probes the energy and momentum dependence of the electronic
excitation spectrum of an N − 1 particle system, the so-called electron-removal portion
of the single-particle spectral function A(k,ω) [5]. In the non-interacting picture, this
spectral function consists of delta-function peaks located at the precise energy and
momentum given by the band structure, i.e. A(k,ω) = δ(ω−ǫk). When interactions
are considered, the single-particle spectral function is modified by the inclusion of the
electron proper self energy Σ(k,ω)=Σ(k,ω)+iΣ′′(k,ω), which captures all the of the
many-body correlation effects. One can then write A(k,ω) = −1
π
Σ′′(k,ω)
[ω−ǫk−Σ(k,ω)]2+[Σ′′(k,ω)]2 .
With respect to the non-interacting case, the peaks in the spectral function shift in
energy and gain a finite width, in a way dependent on the energy and momentum of the
excitations. At those ω and k for which the spectral function is still characterized by a
single pole, energy and lifetime renormalization are directly described by Σ(k,ω) and
Σ′′(k,ω), respectively. The ARPES lineshape thus gives direct access to the lifetime of
the excitation and can provide insights into the nature of the underlying interactions, for
example whether or not electron-electron interactions are Fermi liquid-like, as discussed
in relation to the Tl2201 transport data of Fig. 2. As it will be discussed in the following
sections, the first successful ARPES experiments on Tl2201 have arrived at an agreement
with bulk probes on key features such as the normal-state Fermi surface [32] and the
superconducting gap [52]. This success suggests that detailed ARPES studies of Tl2201
have the potential to reveal the nature and strength of many-body correlations, upon
approaching high-Tc superconductivity from the more conventional overdoped regime.
3.1. Band Structure Calculations
Since it is generally believed that the normal metal on the very overdoped side of the
HTSC phase diagram can be described as a rather conventional Fermi liquid system, in

Page 13
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
13
Figure 7. (a) Electronic dispersion obtained from band structure calculations within
the local-density approximation (LDA). (b,c) Fermi surface calculated for two different
doping levels enclosing a volume, counting holes, of 50% (orange) and 63% (blue) of the
two-dimensional projected Brillouin zone; the corresponding position of the chemical
potential with respect to the dispersive electronic bands is indicated in panel (a). Here,
as in the rest of the paper, k-space labels are expressed modulo the lattice constants.
contrast to the strongly correlated Mott insulator found at half filling (p=0 in Fig.1),
exploring the electronic structure and Fermi surface of heavily overdoped Tl2201 can
start with the results of non-interacting band structure calculations performed within
the local density approximation (LDA). As shown in Fig. 2 and especially Fig. 5, the most
important structural element is the CuO2 planes, which are well separated from each
other by the TlO and BaO layers. As in all other HTSC cuprates, the CuO2-derived
bands are expected to be the lowest energy electronic states and therefore directly
determine the macroscopic electronic properties.
The results of band structure calculations along the high symmetry directions
in the body-centered tetragonal Brillouin zone of Tl2201 are presented in Fig.7(a).
Coordinates for the atoms are taken from Ref.[53]. The essential low-energy feature of
the band structure is indeed the highly dispersive Cu(3dx2−y2 )-O(1)(2px,y) band; note
however that this band is highly two-dimensional, with little dispersion along the kz
direction (0, 0,kz =0)→(0, 0,kz =π). Its top and bottom are located at (π, π) and (0, 0),
respectively, within the two-dimensional projected Brillouin zone, while at (π, 0) we find

Page 14
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
14
the well-known van Hove singularity. The other two bands that sit at a comparable
energy, i.e 0 to −2eV in Fig.7(a), are the anti-bonding Tl(6s)-O(2),O(3)(pz) bands,
which do show significant dispersion in the kz direction. Based on the formal valences of
stoichiometric and undoped Tl2Ba2CuO6+δ (2:2:1:6 with Tl3+, Ba2+, Cu2+, O2−, δ=0),
the Fermi energy would be located at the red line in Fig. 7(a), and both the CuO and TlO
bands would cross Ef . This would result in a Fermi surface with hole pockets associated
with the CuO band, and a small, spheroidal TlO electron pocket at the Γ = (0, 0)
point, as in Fig.7(b). This small electron pocket was originally proposed as a possible
explanation for why undoped Tl2Ba2CuO6 does not show the Mott insulating behavior
expected for materials with a half-filled 3dx2−y2 CuO band, such as undoped La2CuO4
[53]. However, the non-stoichiometry of our samples (Tl1.884(6)Ba2Cu1.11(1)O6+δ, see
Section 2.2) provides additional hole doping (∼ 0.14 holes/formula unit) that would
push the CuO band further away from half-filling, driving the TlO band above Ef . This
shift would generate a Fermi surface consisting solely of a CuO hole pocket around
(π, π) similar to the results shown in blue in Fig. 7(a,c), which were calculated to match
the 63% volume observed by ARPES on our Tc =30 K overdoped Tl2201 crystal [35].
Finally, it should be mentioned that if the effect of hole doping through interstitial
oxygen were explicitly included in the calculations, the TlO bands would be pushed to
much higher energies, beyond the rigid band picture [54]. The analogous effects of Pb
substitution or excess oxygen in the Bi-O layers of Bi-cuprates has recently been used
to account for the lack of Bi-O electron pockets around (π,0) in those materials [55].
3.2. Electronic Dispersion and Fermi Surface by ARPES
Fig. 8(b) shows ARPES data from a Tc =30 K very overdoped Tl2201 sample (Tl2201-
OD30) taken close to the (0,0)-(π,π) direction in momentum space, the so-called “nodal”
direction where the d-wave superconducting gap is zero. Each line in Fig.8(b) is an
energy distribution curve (EDC) from a given position in momentum space along the
line marked as I in the Fermi surface plot of Fig.8(a). These EDCs show a strongly
dispersing quasiparticle peak, related to the 3dx2−y2
CuO band, which emerges from
the background at high binding energies and progressively sharpens as it disperses all
the way to the Fermi energy EF . Note that the term quasiparticle (QP) is used in a
loose sense to identify a reasonably sharp dispersive peak, without specifying in detail
how the peak should be interpreted in terms of specific models for correlation effects
The ARPES experiments were carried out at the Swiss Light Source (SLS) on the Surface and
Interface Spectroscopy Beamline [56, 57, 58] and at the Stanford Synchrotron Radiation Laboratory
(SSRL) on Beamline 5-4, in both cases using a Scienta SES-2002 photoelectron spectrometer. At SLS,
all measurements were performed with circularly polarized 59eV photons, with energy and angular
resolutions of approximately 24 meV and 0.2; the data were acquired on a strongly overdoped sample
with Tc=30K (Tl2201-OD30), and a lightly overdoped sample with Tc=63K (Tl2201-OD63). The
samples were cleaved in situ at 6×10−11 torr and kept at 10 K throughout the measurements. At SSRL,
all data were acquired at 28 eV with linearly polarized light and with energy and angular resolutions of
15 meV and 0.35. The SSRL data were from lightly overdoped samples with Tc=74 K (Tl2201-OD74),
which were cleaved at 3×10−11 torr and temperature cycled between 10 and 85K.

Page 15
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
15
Figure 8. (a) Fermi surface calculated for a 63% Brillouin zone volume, counting
holes, as in Fig.7(c). (b,c) ARPES spectra taken at T =10K on Tl2201-OD30 along
the directions marked by arrows I and II in (a). (d) Second derivative vs. energy of
the spectra from along arrow III in (a); the red line is our tight-binding fit (see text).
(more discussion will be given in Sections 3.4 and 4). From the results of band structure
calculations (Fig. 7), we expect the bottom of this band to be located at the Γ point, or
(0,0), at a binding energy of approximately 1 eV with respect to the chemical potential.
Although it is hard to track the band all the way down to high binding energies in the
raw data, it is possible to highlight the band dispersion by taking a second derivative
of the ARPES intensity with respect to energy. This is plotted in Fig.8d and indicates
a filled bandwidth of only ∼ 250meV, suggesting a renormalization factor of about 4
between single particle band structure calculations and measured ARPES spectra.
Near (π,0), the so-called “antinodal” direction where the d-wave superconducting
gap exhibits its largest value, we also detected well-defined dispersive QP peaks, which
define a shallow parabolic band. The bottom of this band corresponds to the extended
van Hove singularity observed near (π,0) in all superconducting cuprates [2, 3]. However,
at this high doping level, the van Hove singularity is located approximately 40 meV below
the chemical potential. In comparison to the band structure calculations, this also shows
a renormalization factor of about 5.
The momentum at which a dispersing QP peak is observed to reach EF and
disappear provides ARPES’s means of determining a Fermi wavevector kF . This
wavevector identifies one point along the normal state Fermi surface, which is the

Page 16
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
16
Figure 9. Fermi surfaces, in the projected two-dimensional Brillouin zone, obtained
from: (a) band structure calculations (imposing a volume of 63% counting holes, as
observed by ARPES on Tl2201-OD30); our (b) ARPES experiments on the Tc = 30 K
sample (Tl2201-OD30); and (c) the AMRO study on a Tc = 20K Tl2201 crystal by
Hussey et al. [32]. The red line in (b) is the result of our tight-binding fit of ARPES
Fermi surface and dispersion; the width of the Fermi surface contour in (c) reflects the
magnitude of the c-axis dispersion, here emphasized by a factor of 4 for clarity [32].
contour in momentum space that separates filled from empty electronic states and
whose existence and details are of fundamental importance for the understanding of the
macroscopic physical properties of a material. By integrating the ARPES spectra over
a ±5 meV energy window about EF for a large number of cuts in momentum space and
then plotting the results versus momentum in the two-dimensional projected tetragonal
Brillouin zone, one obtains an estimate of the normal-state Fermi surface. This has been
done for Tl2201-OD30, and is shown in Fig.9(b). The location of the Fermi surface
crossings has been determined over more than one quadrant across two different zones,
and then downfolded to the first Brillouin zone and four-fold symmetrized to clearly
show the detailed shape of the Fermi surface in the reduced zone scheme, with improved
signal-to-noise. As expected, there is no indication of a TlO electron pocket around (0,0).
The Fermi surface determined from this procedure takes the form of a large hole pocket
centered at (π,π) with an area that occupies 63±2% of the Brillouin zone, corresponding
to a carrier concentration of 1.26±0.04 hole/Cu atom, or p=0.26±0.04 greater hole
density than the half-filled Mott insulator with 1 hole/Cu (p=0 in Fig.1). This result
is in superb agreement with the recent study of the Fermi surface by angular dependence
of magnetoresistance oscillations (AMRO) experiments [32], which found a hole-pocket
volume of 62% of the Brillouin zone in a slightly more overdoped Tc = 20K sample
(Fig.9(a)). The ARPES determination is also in good agreement with the estimates
from low temperature measurements of the Hall coefficient, which gave a hole doping
of p=1.30 holes/Cu for a Tc ≲15K sample [33]. It is worth noting that the commonly
used empirical formula, Tc/Tmax
c
= 1−82.6(p−0.16)2, which purports to connect the
value of Tc to doping in a universal fashion [59], gives a much stronger dependence of Tc
on doping than is shown by ARPES, AMRO, and Hall coefficient data. For our Tl2201-
OD30 sample, from the above formula and the value p=0.26 given by the Fermi surface

Page 17
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
17
volume, one would obtain Tc ≃16 K; and for the Tc =15 K sample studied by Mackenzie
et al. [33], with the formula and the value p=0.30 determined from the Hall coefficient,
one would actually get Tc = 0K. According to the empirical formula and contrary to
what observed, Tl2201 should become non-superconducting already for p≃0.27.
As a quantitative measure of the shape of the Fermi surface and of the many-body
renormalized electronic dispersion, the Tl2201-OD30 ARPES data can be modelled
by the tight-binding formula ǫk = µ+ t1
2
(cos kx + cosky)+t2 coskx cosky + t3
2
(cos 2kx +
cos 2ky)+ t4
2
(cos 2kx cos ky +coskx cos 2ky)+t5 cos 2kx cos 2ky [60], setting a = 1 for the
lattice constant. With parameters µ = 0.2438, t1 = −0.725, t2 = 0.302, t3 = 0.0159,
t4 =−0.0805, t5 =0.0034, all expressed in eV, this dispersion reproduces both the Fermi
surface shape, solid red line in Fig.9(b), and the QP energy at (0,0) and (π,0), as
seen in Fig.8(c,d). It is worth noting that experimentally the band bottom at (π,0)
is extremely flat, a behavior that could not be reproduced by including only t1 and t2
hopping parameters in the model. Alternatively, a simple analytical formula for the
three-dimensional electronic dispersion of Tl2201 has recently been derived, within the
framework of the linear combination of atomic orbitals (LCAO) approximation, and was
used to fit both ARPES and AMRO results [61]. A basis set spanning the Cu 4s, Cu
3dx2−y2 , O 2px and O 2py states was used in order to take into account the effective
Cu-Cu hopping amplitude between Cu 4s orbitals in neighboring CuO2 layers. This
approach emphasizes the significance of the Cu 4s hopping amplitude in order to obtain
the correct three-dimensional Fermi surface shape, an issue that was originally raised
by Andersen et al. [62]. With regards to the detailed shape of the Fermi surface, LDA
band structure calculations predict a more square contour than observed by ARPES and
AMRO (Fig.9(a)). The inclusion of correlation effects, as well as Cu-Tl substitution
or interstitial O-doping beyond a rigid-band picture, might lead to a better agreement;
preliminary attempts in this direction have not yet led to qualitative improvements [54].
3.3. ARPES Study of the Superconducting Gap
In order for Tl2201 to be a model system to study the overdoped HTSCs by surface and
bulk sensitive probes, it is important to show that ARPES measures quantities that are
characteristic of the bulk, for both normal and superconducting states. Therefore, in
addition to the quantitative agreement on the normal state Fermi surface, the observance
of a superconducting gap in agreement with bulk measurements is also a necessary
requirement. In the following, we will discuss three different methods for the observation
of a gap by ARPES. The first two show a gap consistent with a dx2−y2 form. The third
method allows one to follow the temperature dependence of the gap, highlighting the
minimal surface degradation that occurs as the temperature is cycled from 10 to 85 K.
The detection of a dx2−y2
gap using ARPES can be most easily visualized by
the comparison of nodal and antinodal symmetrized spectra [64]. The spectra are
symmetrized in energy about Ef , by taking I(ω)+I(−ω), which minimizes the effects

Page 18
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
18
Figure 10. (a,b) Symmetrization of Tl2201-OD30 ARPES spectra from along cuts I
and II in Fig.8(b,c). ARPES spectra and their symmetrization along (c,d) the nodal
direction for overdoped Bi2Sr2CaCu2O8+δ (Tc =88K), and (e,f) perpendicular to the
antinodal direction for overdoped Bi2Sr2CuO6+δ (Tc =23K); after Ref.[63]. For all
samples, the approximate locations of the k-space cuts are indicated in the Brillouin
zone sketches. Bold lines in (a,c,d) mark the spectra that cross the Fermi energy,
identifying a Fermi wavevector kF ; no crossing is observed for the spectra in (b,e,f).
of the Fermi function.¶ While this procedure does not return a quantitative value for
the size of the superconducting gap, it provides a qualitative criterion for determining
whether or not there is a Fermi crossing, and hence whether or not a superconducting
gap has opened along the normal-state Fermi surface. In the symmetrized ARPES
data, the presence of a peak in the spectra at EF indicates the presence of a Fermi
surface crossing. This procedure has been used extensively for the Bi-based cuprates,
both single and bilayer compounds, in detailed investigations of the normal-state Fermi
surface, as shown for instance in Fig. 10(c-f) [63], and of the superconducting as well as
normal state pseudogap [64, 66]. In the case of our Tl2201-OD30 sample, this crossing
is clearly seen in the nodal direction (bold line in Fig.10(a)) but not in the antinodal
direction (Fig.10(b)), which is consistent with a d-wave functional form for the gap.
A more quantitative analysis of the gap can be performed by fitting ARPES spectra
along cuts that cross the underlying normal-state Fermi surface, as shown in Fig. 11. The
¶ The symmetrization procedure assumes that the ARPES spectra are described by I(ω)=f(ω)A(k,ω),
and that there is particle-hole symmetry for a small range of ω about EF such that A(−ǫk, −ω) =
A(ǫk,ω). With the identity f(−ω)=1−f(ω), it then follows that I(ω)+I(−ω)= A(k,ω). It is worth
pointing out that this procedure is not strictly valid in the case that energy and momentum resolutions
are included in the description of I(ω), via convolution with the resolution function R(∆k, ∆ω).

Page 19
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
19
Figure 11. (a) Enlarged view of the Fermi surface of Tl2201-OD30 near (π,0). (b)
Selected spectra from along cut II in (a); their k-space positions are indicated by circles
of corresponding color. (c,d) QP linewidth Γ and peak position from a Lorentzian fit of
the energy distribution curves along cut II in (a). (e) Similarly, QP peak position along
cut III in (a). Black lines in (a,d,e) are our tight-binding results (see text). (f) Normal
and (g) superconducting state single-particle spectral function, highlighting particle-
hole mixing and backward dispersion of Bogoliubov QPs below Tc (after Ref. [65]).
model used to reproduce the spectra is a Lorentzian peak plus a step-like background.
The latter is determined from the ARPES spectra with k ≫ kF and is used to help
phenomenologically isolate the coherent part of the spectral function. This function
is then multiplied by the Fermi function and convolved with the instrumental energy
resolution function, to obtain the functional form to be fit to the data [2]. This procedure
for determining the gap has been used previously for Bi2Sr2CaCu2O8+δ [67]. The inset
of Fig.11(b) shows good agreement between the raw data and the fit. Since heavily
overdoped cuprates have weaker, possibly Fermi liquid-like electron correlations, good
agreement between the measured QP peak and the Lorentzian is in principle expected.
Fig. 11(d,e) show the peak positions from the fits compared with our tight-binding
description of the normal-state electronic structure near the antinodal region. At higher
binding energies, there is good agreement between the fit peak positions and the normal-
state dispersion. At lower binding energies, however, the peak does not reach EF , but
instead reaches a minimum at ∆P ≃17meV and then disperses back to higher binding
energy. This behavior is a hallmark of Bogoliubov QPs in a superconductor, as shown in
the sketch of Fig. 11(f,g), for which a beautiful experimental demonstration was recently
obtained by ARPES on the trilayer Bi-cuprate Bi2Sr2Ca2Cu3O10+δ [65]. Simultaneously,

Page 20
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
20
Figure 12.
Symmetrized ARPES spectra measured at (a) 10K and (b) 85K on
overdoped Tl2201-OD74, along the (π, 0)−(π, π) direction. Bold lines indicate the
EDCs closest to kF . (c) The k integrated spectral weight at the antinode, for several
temperatures, together with a 10K Au Fermi edge and an 85K resolution-broadened
Fermi function; an enlarged view of the leading edge positions is given in the inset.
we observe a reduction of spectral weight, by about a factor of 2, when the QP peak
has dispersed from the van Hove singularity all the way to kf . This is consistent with
the spectral intensity of Bogoliubov QPs being determined by the coherence factor
v2
k
=1−u2
k
= 1
2
(1−ǫk/Ek), with Ek =√ǫ2
k +∆2
k
, which corresponds to 1/2 when ǫk kf =0
and 1 for |ǫk|≫∆k. Extending this analysis to other momenta along the Fermi surface,
one can study the k-dependence of the gap: at ∼(π/2, π/2) the peak does cross EF (not
shown), while at intermediate momenta a gap smaller than at (π, 0) is observed, as on
the right-hand side of Fig.11(d), consistent with d-wave symmetry.
The temperature dependence of the gap at the antinode has been measured for
Tl2201-OD74. Fig.12(a,b) present symmetrized ARPES data along the (π, 0)−(π, π)
direction, at temperatures above and below Tc. In exactly the same manner as the
k-dependent symmetrized data of Fig.10(a,b), the temperature-dependent data of
Fig.12(a,b) show a clear gap at k = kF at 10K, which is clearly smaller at 85K (with
this type of analysis, the broadness of the QP peak and noise level at 85 K make it hard
to conclude whether the gap has completely closed). Fig.12(c) shows averaged EDCs
taken over a narrow k-range around kf at 10, 45, 85, and again at 10K. Included in
the figure is a 10K Au reference spectrum, and the simulation of an energy-resolution
broadened Fermi function at 85K. The inset of Fig.12(c) shows a closer view of these
k-integrated EDCs around Ef . It is clear from this figure that, as the temperature
increases, the size of the gap decreases although it does not seem to completely close
at 85K. This is suggested by comparing the position of the leading edge of the 85K
ARPES data with the simulated Fermi function; they are close in energy but not quite
coincident. It is possible that there is still a pseudogap at temperatures slightly above

Page 21
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
21
Figure 13. (a) Superconducting gap magnitude estimated by tunneling and ARPES
from the superconducting peak position (SCP) and the leading-edge midpoint (LEM)
shift, plotted versus Tc for various optimally or overdoped cuprates (after Ref. [68]). (b)
Gap values from thermal conductivity, tunneling, and ARPES plotted vs. |1−Tc/Tmax
c
|
for many HTSCs across the phase diagram (after Ref. [52]); dashed lines are guides to
the eye. Both plots include our own Tl2201-OD74 and Tl2201-OD30 ARPES data.
Tc = 74K at this doping level. It is worth emphasizing that as the temperature is
lowered back to 10K, the gap reopens to approximately the same size. Although a
partial degradation of the ARPES features can be seen, the QP peaks and the gap are
still clearly observable. This is strong evidence that the surface is stable under these
experimental conditions and is thus suitable for detailed surface-sensitive experiments.
From these data, a superconducting peak (SCP) position of ∼33 meV and a leading-edge
midpoint (LEM) gap of ∼15 meV can be extracted.
As a direct comparison with other cuprates and different means of determining
the gap, Fig.13(a,b) present a compilation of the doping dependence of the gap
magnitude. Fig.13(a) refers only to spectroscopic studies of optimally and overdoped
HTSCs, for which the gap is proportional to Tc. Our results from Tl2201-OD74 and
Tl2201-OD30 are fully consistent with the behavior observed on the other cuprates,
as far as both SCP and LEM positions are concerned. Even more significant is
Fig.13(b), which demonstrates that ARPES measures gap values that follow the same
trend as those derived from thermal conductivity, a bulk transport property. A more
accurate comparison would require an analysis of the ARPES data beyond the mere
determination of SCP and LEM positions, for instance on the basis of a model spectral
function; this is however complicated by the non-trivial doping dependence of the
ARPES data, and in the present case was done for Tl2201-OD30 but not Tl2201-OD74.

Page 22
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
22
3.4. ARPES Lineshape Analysis
The electrical and thermal transport properties suggest that overdoped Tl2201 may be
regarded as a Fermi liquid [33, 34, 32]. Since the ARPES data for both the normal and
superconducting states indicate that single-particle surface-sensitive probes do provide
information representative of the bulk, overdoped Tl2201 might be an ideal material
for a quantitative study of the strength and nature of the many-body effects, which are
revealed by the energy and momentum evolution of a QP’s intrinsic ARPES lineshape
[5]. In general, however, the shape of an ARPES peak does not correspond directly to
an excitation’s intrinsic lineshape. This is primarily due to an extrinsic contribution
from momentum and energy resolution of the analyzer. Recent work on the model
two-dimensional Fermi liquid system Sr2RuO4 highlights the significant issues involved
in removing the effects of the analyzer resolution [69]; in particular, it clearly shows
that both the lineshape width and the peak position can be affected by resolution
effects. There are significant challenges to completely removing the analyzer resolution
from the data and stringent requirements on the data quality needed to attempt such
analysis; therefore, we will analyze the current Tl2201 spectra in a more qualitative
and phenomenological manner, concentrating primarily on the momentum dependence
of the QP peaks measured in the superconducting state for different doping levels.
Beyond the issue of the experimental resolution, matrix element effects and the
handling of the ARPES background are the two main complications in attempting
even a qualitative description of the momentum dependence of the QP lineshapes.
Experimentally, problems due to matrix element effects may be reduced by appropriate
choice of measurement conditions such as geometry, photon energy, and polarization.
For the experiments performed at the SLS and the data discussed in the following,
circularly polarized 59eV photons were used. Circular polarization was chosen
specifically to minimize matrix element effects, which, using symmetry arguments, can
be shown to be most extreme for linear polarization [2]. The photon energy was selected
based on our experience on other cuprates — specifically, La2−xSrxCuO4 (LSCO) where
similar photon energies were used to reveal structure that had previously been missed in
lower-energy ARPES data [70, 26]. Finally, in order to gain more reliable information,
the ARPES data were taken over multiple Brillouin zones; this way it was possible to
verify that, for the same point in the reduced zone scheme, the lineshapes from different
zones in the extended zone scheme were qualitatively the same.
The most commonly used method to isolate a QP peak from the experimental
background [2, 71], is to subtract a phenomenological ARPES background determined
from the data at k-values far removed from those at which the QP-like peaks are
detected. In our analysis we will use this approach, taking as background the weakly
k-dependent photoemission intensity observed in each momentum space cut for k ≫ kF .
The fundamental problem with this approach is that if some of the background is related
to the incoherent part of the spectral function, as one would expect for correlated
electron system, then by subtracting the background we are actually disregarding some

Page 23
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
23
Figure 14. (a) Tl2201-OD30 ARPES spectra at k slightly smaller than kF along
the Fermi surface contour (corresponding to a QP binding energy of ∼35meV). (b)
Selected spectra from (a) along with their corresponding k ≫ kF background. (c)
Spectral weight of the background subtracted spectra, integrated over different energy
ranges and normalized with respect to the α=0 antinodal value, plotted vs. the Fermi-
surface angle α. (d) Quasiparticle linewidth Γ plotted vs. the Fermi-surface angle α.
of the most crucial information contained in the ARPES spectra. In fact, the k- and
ω-dependence of the incoherent part of the spectral function is, in many respects, as
important as the behavior of the QP peak itself; unfortunately we do not yet have the
means to reliably extract information from it.
In Fig. 8(b), similar to what was previously observed in all photoemission studies of
the cuprates [2, 3], we saw that in the nodal region the width of the QP peaks increases
as a function of binding energy, as expected from simple phase-space arguments. This
is, however, in sharp contrast to what is observed in the antinodal region, where the
sharpest peak is found at the bottom of the band: in Fig.11(c), the linewidth Γ of the
QP peak is observed to grow from ∼30 to ∼55 meV as the QP peak disperses from ∼39 to
∼20 meV binding energy. Possible origins for this anomalous behavior will be discussed
later, but it should be pointed out that an analogous effect was recently reported for
very overdoped and nearly non-superconducting Bi1.74Pb0.38Sr1.88CuO6+δ [72]. Another
striking feature of the lineshape evolution is that the QP peaks are much broader in
the nodal region than in the antinodal region, as evidenced by the direct comparison of
Fig. 8(b) and 8(c). To elaborate on this, in Fig. 14(a) we present a compilation of spectra
from along the Fermi surface contour, with k slightly smaller than kF and corresponding
to a binding energy of ∼35 meV. The choice in favor of spectral peaks at a binding energy

Page 24
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
24
Figure 15.
Nodal and antinodal spectra from (a) Tl2201-OD63 and (b) Tl2201-
OD30, at k ≲ kF . Inset: similar data from overdoped LSCO-OD24 (p = 0.22), from
Ref. [73]. For both samples in (a), Tc 2
3 Tmax
c
with respect to the Tmax
c
of each family
[27].
lower than EF is dictated by the need to compare spectra not affected by either the
opening of the d-wave gap or the anomalous low-binding-energy broadening seen at
the antinode. The sharp peak near (π, 0) becomes progressively broader as (π/2, π/2)
is approached. In order to determine whether this apparent broadening represents an
increase in the QP linewidth Γ, or is merely a loss of spectral weight, the momentum-
independent background from k ≫ kF is subtracted from the ARPES spectra as shown
in Fig.14(b); then the data are integrated over varying energy ranges and plotted in
Fig. 14(c), normalized at the Fermi surface angle α=0. For narrow integration windows,
there is a drop as a function of α: the QPs show a loss of low-energy weight (and possibly
coherence) in the nodal region. However, once the integration window is expanded to
about 550meV, the spectral weight of the QP peaks becomes angle-independent. This
indicates that the k-dependent broadening of the ARPES spectra indeed reflects a loss
of coherence of the QP spectrum, rather than being due to matrix element effects.
The observed QP anisotropy of Tl2201-OD30, with peaks much broader in the
nodal than the antinodal region, is in sharp contrast to the behavior observed in
underdoped cuprates, where the QP peaks are sharp near (π/2, π/2) and very broad
around (π, 0), as summarized in Fig.17. As doping is increased, the antinodal QP
peaks sharpen but remain significantly broader than the nodal QPs up to optimal
doping, at least in the normal state. Even in the superconducting state, where antinodal
QPs sharpen considerably, the ARPES linewidths for underdoped and optimally-doped
materials are still highly anisotropic, with a minimum at (π/2, π/2) [2, 3]. Overall
this behavior leads to the expectation that increasing doping beyond the optimal value
will simply render the QP linewidths progressively more isotropic. Indeed, isotropic
lineshapes were observed on early data from very overdoped, non-superconducting

Page 25
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
25
Figure 16.
(a) ARPES intensity map for emission from the Fermi energy for
overdoped LSCO-OD24 (p=0.22). Red lines are the calculated LDA three-dimensional
Fermi surface for kz =0 and π/c; the enclosed area is the projection of the Fermi surface
onto the two-dimensional kx-ky plane, and denotes the region allowed for emission [74].
(b) ARPES and AMRO data from Tl2201-OD30. In (a,b) the yellow arrows identify
directions characterized by zero kz dispersion as indicated by the LDA calculations
and consistent with AMRO experiments, which also suggest much weaker residual
three-dimensionality effects in Tl2201 than in LSCO. (c) Resolution contributions to
the width of an ARPES EDC calculated, as a function of the Fermi velocity vF , from
the two-dimensional convolution of a Lorentzian quasiparticle peak (binding energy
ω = 30meV and intrinsic width Γ = 18meV) with Gaussian energy and momentum
resolution functions (results corresponding to our experimental parameters are in red).
Bi1.80Pb0.38Sr2.01CuO6−δ [75], although more recent work on the same material [72]
exhibits a behavior more consistent with what is reported here for Tl2201. Going back
to the QP anisotropy reversal observed for Tl2201-OD30 in Fig. 14, and presented again
in Fig. 15(b) for nodal and antinodal points only, it must be emphasized that the same,
albeit less pronounced, effect was observed in our Tl2201-OD63 sample, as shown in
Fig. 15(a). For comparison, in the inset of Fig. 15(a) we have plotted ARPES data from
overdoped (x=p=0.22) La2−xSrxCuO4 [73]. This sample has Tc 2
3
Tmax
c
, comparable
to the degree of overdoping of our sample Tl2201-OD63, which also has Tc 2
3
Tmax
c
for the much larger Tmax
c
of the Tl2201 family. The data from LSCO exhibit a QP
anisotropy similar to that in Tl2201-OD63, suggesting that the observed QP anisotropy
reversal might indeed be generic to the overdoped cuprates.
Before proceeding to the discussion of the broader significance of these findings, we
comment here on two extrinsic effects that could potentially contribute to an anomalous
k-dependent broadening of the QP lineshapes: residual kz electronic dispersion and
resolution broadening effects. First, the dispersion of the electronic structure along the
c-axis may indeed give rise to k||-dependent broadening of the ARPES features [74]

Page 26
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
26
(where k|| is the ab-plane momentum). This effect could be doping-dependent because
there is a 0.4% c-axis lattice constant reduction as doping is increased [29]. However, the
kz-dispersion vanishes in Tl2201 along the (0, 0)−(π, π) and (π, 0)−(π, π) directions, and
thus at all nodal and antinodal points. This is a direct consequence of the symmetry of
the body-centered tetragonal unit cell, and holds also for LSCO. Fig.16(a,b) illustrate
this point by showing ARPES data for both materials together with band structure
calculations for LSCO [74], and AMRO data for Tl2201 [32]. Along these high-symmetry
directions, no kz-dispersion is present either in the LSCO calculations or the Tl2201
AMRO data (in both cases the kz-dispersion is proportional to the width of the three to
two-dimension projected contours). This is consistent with the ARPES maps for the two
systems, although the broader EF intensity patches for LSCO indicate that the three-
dimensionality is much stronger in LSCO than in Tl2201. A residual kz dispersion might
contribute to the overall width of the ARPES features, but it cannot be responsible for
the reversal of the anisotropy observed on overdoped cuprates, since the anisotropy
exhibits a monotonic k||-dependence between (π/2, π/2) and ∼(π, 0).
The second extrinsic effect that influences the width in energy of an ARPES
spectrum, or EDC, is instrumental resolution. In a two-dimensional system, the width
in energy is determined by the intrinsic inverse lifetime of the QP excitation and by the
experimental resolution broadening, which reflects both the energy (∆ω) and angular
(∆k) resolution of the apparatus [5]. For a dispersionless feature, the instrumental
contribution to the total EDC width is determined solely by the energy resolution ∆ω.
However, for a dispersive feature, the angular resolution ∆k contributes to the energy
broadening in a manner directly proportional to the band velocity (vF =0 in Fig. 16(c)).
Since the Fermi velocity around the Fermi surface of the cuprates is highly anisotropic,
varying from approximately 0.5 to 1.8 eVÅ in going from antinodal to nodal region, this
should give rise to a momentum-dependent resolution broadening of the EDCs. This
broadening would be more severe at the nodes, where vF is at its largest, which raises
the concern that the larger nodal widths shown in Fig. 14 and 15 might be an artifact of
the instrumental resolution. We have carefully evaluated the contribution of resolution
broadening effects for our experimental conditions with a variety of different procedures
and the results are summarized in Fig.16(b). For our experimental parameters (red
data, ∆ω = 24meV and ∆k = 0.3) the k-dependent resolution broadening is of the
order of few meV, nowhere near the factor of 4 observed in Fig.14(d).
4. Discussion and Conclusions
Using a self-flux method and partial encapsulation, we have grown single crystals
of the single-layer overdoped cuprate superconductor Tl2201. The crystals were
annealed under controlled oxygen partial pressures to set a homogeneous doping level
whilst preventing decomposition, and typically exhibited sub-Kelvin superconducting
transition widths as measured in a SQUID magnetometer. Their high quality and
homogeneity were evidenced not only by these narrow transition widths but also by

Page 27
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
27
Figure 17.
Nodal-antinodal quasiparticle anisotropy reversal observed in the
cuprates. As a function of the Fermi surface angle α, sharp QP peaks are observed at
the nodes (N) and ill-defined QPs at the antinodes (A) in underdoped cuprates (e.g.,
data from Ca2−xNaxCuO2Cl2, after Ref.[10]); in overdoped Tl2201, this behavior is
reversed.
their extremely narrow rocking curve widths, comparable to YBa2Cu3O7−δ grown in
YSZ crucibles [50]. An EPMA study found the crystals to have the composition
Tl1.884(6)Ba2Cu1.11(1)O6+δ, which was homogeneous and was not observed to vary
between crystals. This level of cation substitution is similar to that reported earlier for
single crystals of Tl2201. For higher dopings (lower Tcs), the crystals were found to be
orthorhombic, a slight deviation from a square-planar lattice that is often associated with
the unsubstituted phase which has only been prepared in ceramics. The orthorhombic
distortion appears to be suppressed by cation disorder, but stabilized by the oxygen
interstitials that do increase the doping into the very overdoped regime.
While these crystals constitute a significant step forward, much optimization of
the growth technique remains, and there is still the challenge of growing single crystals
without the cation disorder in the Tl layer. Still, the high quality Tl2201 single crystals
grown as part of this effort have enabled the first successful ARPES experiments of
this compound. The results on the Fermi surface of Tl2201 mark an important new
starting point for understanding the cuprates, namely that there is a material where both
a surface-sensitive single-particle spectroscopic technique (ARPES) and comparable
bulk transport measurements (AMRO) have arrived at quantitative agreement on a
major feature of the normal state [32, 35]. This is a first for the copper-oxide high-Tc
superconductors and, within the more general class of 3d and 4d transition metal oxides,
it is second only to the case of Sr2RuO4 for which a similar quantitative agreement was
reached between de Haas-van Alphen [76, 77] and ARPES [69, 78, 79, 80, 81]. The
detailed agreement on the Fermi surface (Fig.9), together with the good quantitative
agreement achieved for the superconducting gap by thermal conductivity and ARPES
(Fig.13), establishes Tl2201 as an ideal system to study the overdoped regime of the

Page 28
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
28
cuprate phase diagram with a wide spectrum of techniques. Next it will be important
to study the single-particle excitation spectrum with both ARPES and STS in enough
detail to compare to other normal state properties, thus critically testing whether or not
a conventional Fermi-liquid description of the electronic properties captures the physics
of the normal metal in the heavily overdoped regime. In this regard, it is important
to elaborate on the most surprising result that emerged from this study of overdoped
Tl2201 by ARPES, one which is counterintuitive within the realm of Fermi liquid theory:
the quasiparticle anisotropy reversal observed across optimal doping. A summary of the
momentum anisotropy in under and overdoped cuprates is given in Fig.17, where data
from Ca2−xNaxCuO2Cl2 [10], for several doping levels, and Tl2201 are directly compared.
Early magnetotransport experiments on very overdoped Tl2201, for which a small
magnetoresistance and a weak T-dependence of the resistivity and cotangent of the Hall
angle have been observed, do not support the presence of strongly k-dependent low-
temperature scattering rates in the normal state [33]. More recently, a temperature,
doping, and magnetic field dependent AMRO study of Tl2201 suggested a two-
component scattering rate consisting of an isotropic T2 term as well as an anisotropic
T-linear term [82], which vanishes at the nodes and has a maximum near the antinodes.
The anisotropic term becomes weaker upon overdoping the material, which is consistent
with the view that as hole doping increases, the antinodal QPs should gain coherence
faster than the nodal ones, eventually leading to a relatively small and fully isotropic
scattering rate. Therefore, what is at variance between the ARPES and AMRO results
from Tl2201 is not the behavior of the antinodal QPs, which in both cases are indeed
gaining coherence upon overdoping, but rather that of nodal QPs which in ARPES
become progressively less coherent. In addition, the overall magnitude of the scattering
rate seen by ARPES is approximately a factor of 10 larger than estimated by AMRO.
What could be the origin for this discrepancy between ARPES and AMRO
determined scattering rates? First of all, one has to note that while the ARPES
experiments are performed in the superconducting state, the AMRO measurements
are carried out in the normal state; and the latter is achieved by the application of
external magnetic fields as high as 45T, which might potentially affect the low-energy
QP dynamics. As for the reliability of the ARPES data on this specific point of the loss of
coherence of nodal QPs at large dopings, it is important to mention that these findings
are also supported by a preliminary study of overdoped Bi2Sr2CaCu2O8+δ by STM
and STS [83]. From the Fourier analysis of the energy-dependent spatial modulations
observed in the tunneling conductance [84], it was concluded that for strongly overdoped
samples, the nodal (i.e., low energy) quasiparticle interference signal is no longer visible,
consistent with a decoherence of the nodal states [83].
A reconciliation of the results might require consideration of the specific
sensitivity of transport probes and single-particle spectroscopies to electronic scattering
phenomena. In particular, photoemission and scanning tunneling are very sensitive
to electronic scattering involving small momentum transfer, while transport is rather
insensitive to it. A possible source of such scattering could be extended impurities

Page 29
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
29
far from the CuO2 planes [85], such as cation substitution or interstitial oxygen which
increase with doping in most cuprates. Recent calculations show that the inclusion of
elastic forward scattering in a d-wave superconductor can lead to an overall increase of
the scattering rate in the normal state, but also to a strong enhancement of antinodal
QP coherence in the superconducting state [86]. A very different model, based on the
description of unitary-limit scattering beyond the Born approximation so as to account
for multiple scattering from a single impurity, would also lead, below Tc, to loss of
nodal coherence and anomalous enhancement of antinodal coherence [87]. However,
both models predict an approximately isotropic QP scattering above Tc, as well as
temperature dependent effects limited to the energy scale of the gap itself. To test the
applicability of these scenarios to the present case, a thorough temperature dependent
study is required. Finally, it should also be pointed out that sources of scattering relevant
to this discussion are not just limited to impurities; also inelastic (i.e., electronic)
scattering involving small momentum transfer might result in anomalous decoherence
of the nodal QPs. For instance, low-energy (i.e., small q) quantum-critical fluctuations
associated with proximity to a competing superconducting dx2−y2 +idxy phase [88], or
ferromagnetic phases, could also lead to a similar QP anisotropy reversal [89].
Acknowledgments
We gratefully acknowledge D.G. Hawthorn, K.M. Shen, N.E. Hussey, A.P. Mackenzie,
J.C. Davis, D.J. Scalapino, and G.A. Sawatzky for discussions and M. Platé, N.P.
Armitage, A.B. Kuzmenko, S. Chiuzbaian, M. Shi, M. Falub, and L. Patthey for
assistance during the ARPES experiments. We are also grateful to machine and
beamline groups at the Swiss Light Source and Stanford Synchrotron Radiation
Laboratory, whose outstanding efforts made the experiments possible. This work
was supported by the Canada Research Chairs Program, the Natural Sciences and
Engineering Research Council of Canada, the Canadian Institute for Advanced
Research, and the British Columbia Synchrotron Institute.
5. References
[1] March 2006, volume 2 no 3, special issue on high-temperature superconductors. Nat. Phys., 2:133–
210, 2006.
[2] Andrea Damascelli, Zahid Hussain, and Zhi-Xun Shen. Angle-resolved photoemission studies of
the cuprate superconductors. Rev. Mod. Phys., 75:473–541, 2003.
[3] J.C Campuzano, M.R. Norman, and M. Randeria. Photoemission in the High Tc Superconductors,
volume II, pages 167–265. Springer, Berlin, 2004.
[4] Øystein Fischer, Martin Kugler, Ivan Maggio-Aprile, Christophe Berthod, and Christoph Renner.
Rev. Mod. Phys., in press, 2006.
[5] A. Damascelli. Probing the electronic structure of complex systems by ARPES. Phys. Scr.,
T109:61–74, 2004.
[6] C. Kim, P.J. White, Z.-X. Shen, T. Tohyama, Y. Shibata, S. Maekawa, B. O. Wells, Y. J. Kim, R. J.
Birgeneau, and M. A. Kastner. Systematics of the photoemission spectral function of cuprates:
insulators and hole- and electron-doped superconductors. Phys. Rev. Lett., 80:4245–4248, 1998.

Page 30
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
30
[7] F. Ronning, C. Kim, K.M. Shen, N.P. Armitage, A. Damascelli, D.H. Lu, D.L. Feng, Z.-X. Shen,
L.L. Miller, Y.-J. Kim, F. Chou, and I. Terasaki. Universality of the electronic structure from
a half filled CuO2 plane. Phys. Rev. B, 67:035113, 2003.
[8] K.M. Shen, F. Ronning, D.H. Lu, W.S. Lee, N.J.C. Ingle, W. Meevasana, F. Baumberger, A.
Damascelli, N.P. Armitage, L.L. Miller, Y. Kohsaka, M. Azuma, M. Takano, H. Takagi, and
Z.-X. Shen. Missing quasiparticles and the chemical potential puzzle in the doping evolution of
the cuprate superconductors. Phys. Rev. Lett., 93:267002, 2004.
[9] T. Hanaguri, C. Lupien, Y. Kohsaka, D.H. Lee, M. Asuma, M. Takano, H. Takagi, and Z.-
X. Shen. A ‘checkerboard’ electronic crystal state in lightly hole-doped Ca2−xNaxCuO2Cl2.
Nature, 430:1001–1005, 2004.
[10] Kyle M. Shen, F. Ronning, D.H. Lu, F. Baumberger, N.J.C. Ingle, W.S. Lee, W. Meevasana, Y.
Kohsaka, M. Asuma, M. Takano, H. Takagi, and Z.-X. Shen. Nodal quasiparticles and antinodal
charge ordering in Ca2−xNaxCuO2Cl2. Science, 307:901–904, 2005.
[11] S. Smadici, P. Abbamonte, M. Taguchi, Y. Kohsaka, T. Sasagawa, M. Azuma, M. Takano,
and H. Takagi. Absence of long-ranged charge order in NaxCa2−xCuO2Cl2 at x=0.08.
cond-mat/0603671, 2006.
[12] W.N. Hardy, D.A. Bonn, D.C. Morgan, R. Liang, and Kuan Zhang. Precision measurements of the
temperature dependence of λ in YBa2Cu3O6.95: Strong evidence for nodes in the gap function.
Phys. Rev. Lett., 70:3999–4002, 1993.
[13] C.C. Tsuei and J.R. Kirtley. Pairing symmetry in cuprate superconductors. Rev. Mod. Phys.,
72:969–1016, October 2000.
[14] J.R. Kirtley, C.C. Tsuei, A. Ariando, C.J.M. Verwijs, S. Harkema, and H. Hilgenkamp. Angle-
resolved phase-sensitive determination of the in-plane gap symmetry in YBa2Cu3O7−δ. Nat.
Phys., 2:190–194, 2006.
[15] D.A. Bonn and W.N. Hardy. edited by D.M. Ginsberg, chapter 2, page 7. in Physical Properties
of High Temperature Superconductors. World Scientific, 1996.
[16] R.W. Hill, Christian Lupien, M. Sutherland, Etienne Boaknin, D.G. Hawthorn, Cyril Proust, F.
Ronning, Louis Taillefer, Ruixing Liang, D.A. Bonn, and W.N. Hardy. Transport in ultraclean
YBa2Cu3O7: Neither unitary nor Born impurity scattering. Phys. Rev. Lett., 92:027001, 2004.
[17] D.N. Basov and T. Timusk. Electrodynamics of high-Tc superconductors. Rev. Mod. Phys.,
77:721–779, 2005.
[18] T.P. Devereaux and R. Hackl. Inelastic light scattering from correlated electrons. Rev. Mod.
Phys. in press, cond-mat/0607554, 2006.
[19] M. Eschrig. The effect of collective spin-1 excitations on electronic spectra in high-Tc
superconductors. Advances in Physics, 55:47–183, 2006.
[20] J.M. Tranquada. Neutron scattering studies of antiferromagnetic correlations in cuprates. to be
published as a chapter in Treatise of High Temperature Superconductivity, edited by J. Robert
Schrieffer, cond-mat/0512115, 2005.
[21] D.J. Derro, E.W. Hudson, K.M. Lang, S.H. Pan, J.C. Davis, J.T. Markert, and A.L. de Lozanne.
Nanoscale one-dimensional scattering resonances in the CuO chains of YBa2Cu3O6+x. Phys.
Rev. Lett., 88:097002, 2002.
[22] M.C. Schabel, C.-H. Park, H. Matsuura, Z.-X. Shen, D.A. Bonn, Ruixing Liang, and W.N.
Hardy. Angle-resolved photoemission on untwinned YBa2Cu3O6.95. I. Electronic structure and
dispersion relations of surface and bulk bands. Phys. Rev. B, 57:6090, 1998.
[23] D.H. Lu, D.L. Feng, N.P. Armitage, K.M. Shen, A. Damascelli, C. Kim, F. Ronning, Z.-X. Shen,
D.A. Bonn, R. Liang, W.N. Hardy, A.I. Rykov, and S. Tajima. Superconducting gap and strong
in-plane anisotropy in untwinned YBa2Cu3O7−δ. Phys. Rev. Lett., 86:4370–4373, 2001.
[24] S.V. Borisenko, A.A. Kordyuk, V. Zabolotnyy, J. Geck, D. Inosov, A. Koitzsch, J. Fink,
M. Knupfer, B. Bchner, V. Hinkov, C.T. Lin, B. Keimer, T. Wolf, S.G. Chiuzbian, L. Patthey,
and R. Follath. Kinks, nodal bilayer splitting, and interband scattering in YBa2Cu3O6+x. Phys.
Rev. Lett., 96:117004, 2006.

Page 31
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
31
[25] I. Maggio-Aprile, Ch. Renner, A. Erb, E. Walker, and Ø. Fischer. Direct vortex lattice imaging
and tunneling spectroscopy of flux lines on YBa2Cu3O7−δ. Phys. Rev. Lett., 75:2754–2757,
1995.
[26] X.J. Zhou, T. Cuk, T. Devereaux, N. Nagaosa, and Z.-X. Shen. Angle-resolved photoemission
spectroscopy on electronic structure and electron-phonon coupling in cuprate superconductors.
to be published as a chapter in Treatise of High Temperature Superconductivity, edited by J.
Robert Schrieffer, cond-mat/0604284, 2006.
[27] H. Eisaki, N. Kaneko, D.L. Feng, A. Damascelli, P.K. Mang, K.M. Shen, Z.-X. Shen, and M. Geven.
Effect of chemical inhomogeneity in bismuth-based copper oxide superconductors. Phys. Rev.
B, 69:064512, 2004.
[28] M. Chaio, R.W. Hill, C. Lupien, L. Taillefer, P. Lambert, R. Gagnon, and P. Fournier. Low-energy
quasiparticles in cuprate superconductors: A quantitative analysis. Phys. Rev. B, 62:3554–3558,
2000.
[29] J.L. Wagner, O. Chmaissem, J.D. Jorgensen, D.G. Hinks, P.G. Radaelli, B.A. Hunter, and
W.R. Jensen. Multiple defects in overdoped Tl2Ba2CuO6+δ: Effects on structure and
superconductivity. Physica C, 277:170–182, 1997.
[30] C.C. Tsuei, J.R. Kirtley, Z.F. Ren, J.H. Wang, H. Raffy, and Z.Z. Li. Pure dx2−y2 order-parameter
symmetry in the tetragonal superconductor Tl2Ba2CuO6+δ. Nature, 387:481–483, 1997.
[31] H. He, P. Bourges, Y. Sidis, C. Ulrich, L.P. Regnault, S. Pailhes, N.S. Berzigiarova, N.N.
Kolesnikov, and B. Keimer. Magnetic resonant mode in the single-layer high-temperature
superconductor Tl2Ba2CuO6+δ. Science, 295:1045–1047, 2002.
[32] N.E. Hussey, M. Abdel-Jawad, A. Carrington, A.P. Mackenzie, and L. Ballcas. A coherent three-
dimensional Fermi surface in a high-transition-temperature superconductor. Nature, 425:814–
817, 2003.
[33] A.P. Mackenzie, S.R. Julian, D.C. Sinclair, and C.T. Lin. Normal-state magnetotransport in
superconducting Tl2Ba2CuO6+δ to millikelvin temperatures. Phys. Rev. B, 53:5848–5855, 1996.
[34] Cyril Proust, Etienne Boaknin, R.W. Hill, Louis Taillefer, and A.P. Mackenzie. Heat transport
in a strongly overdoped cuprate: Fermi liquid and a pure d-wave BCS superconductor. Phys.
Rev. Lett., 89:147003, 2002.
[35] M. Platé, J.D.F. Mottershead, I.S. Elfimov, D.C. Peets, R. Liang, D.A. Bonn, W.N. Hardy,
S. Chiuzbaian, M. Falub, M. Shi, L. Patthey, and A. Damascelli. Fermi surface and quasiparticle
excitations of overdoped Tl2201 by ARPES. Phys. Rev. Lett., 95:077001, 2005.
[36] M.P. Siegal, E.L. Venturini, B. Morosin, and T.L. Aselage. Synthesis and properties of Tl–Ba–
Ca–Cu–O superconductors. J. Mater. Res., 12:2825–2854, 1997.
[37] T.K. Jondo, R. Abraham, M.T. Cohen-Adad, and J.L. Jorda. The BaO—TlO1.5 system. J. Alloys
Compd., 186:347–359, 1992.
[38] Masashi Hasegawa, Humihiko Takei, K. Izawa, and Y. Matsuda. Crystal growth techniques for
Tl-based cuprate superconductors. J. Cryst. Growth, 229:401–404, 2001.
[39] André Tyler. An Investigation into the Magnetotransport Properties of Layered Superconducting
Perovskites. Ph.D. thesis, University of Cambridge, 1998.
[40] R.S. Liu, S.D. Hughes, R.J. Angel, T.P. Hackwell, A.P. Mackenzie, and P.P. Edwards. Crystal
structure and cation stoichiometry of superconducting Tl2Ba2CuO6+δ single crystals. Physica
C, 198:203–208, 1992.
[41] Masashi Hasegawa, Yoshitaka Matsushita, Yasuhiro Iye, and Humihiko Takei. Growth and
anisotropic superconducting properties of Tl2Ba2CuO6 and Tl2Ba2CaCu2O8 single crystals.
Physica C, 231:161–166, 1994.
[42] N.N. Kolesnikov, M.P. Kulakov, V.N. Molchanov, I.F. Schegolev, R.P. Shibaeva, V.I. Simonov,
R.A. Tamazyan, and O.M. Vyasilev. Comparative study of Tl-2201 single crystals with Tc =
30 and 110 K by means of X-ray structural analysis and NMR. Physica C, 242:385–392, 1995.
[43] O.M. Vyasilev, N.N. Kolesnikov, M.P. Kulakov, and I.F. Schegolev. Tl NMR study of Tl2Ba2CuOx
single crystals with various Tc. Physica C, 199:50–58, 1992.

Page 32
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
32
[44] N.N. Kolesnikov, V.E. Korotkov, M.P. Kulakov, V.N. Molchanov, R.A. Tamazyan, and
V.I. Simonov.
Structure of superconducting single crystals of 2201 thallium cuprate
(Tl1.85Cu0.15)Ba2CuO6, Tc = 110 K. Physica C, 195:219–224, 1992.
[45] C. Opagiste, G. Triscone, M. Couach, T.K. Jondo, J.-L. Jorda, A. Junod, A.F. Khoder, and
J. Muller. Phase diagram of the Tl2Ba2CuO6 compounds in the T, p(O2) plane. Physica C,
213:17–25, 1993.
[46] B. Raveau, C. Michel, B. Mercey, J.F. Hamet, and M. Hervieu. Crystal chemistry of
superconducting copper oxycarbonates. J. Alloys Compd., 229:134, 1995.
[47] J.L. Jorda, T.K. Jondo, R. Abraham, M.T. Cohen-Adad, C. Opagiste, M. Couach, A. Khoder,
and F. Sibieude. Preparation of pure Tl2Ba2CuO6±x: The contribution of phase equilibrium
studies. Physica C, 205:177–185, 1993.
[48] J.L. Pouchou and F. Pichoir. “PAP” ϕ(ρZ) Procedure for Improved Quantitative Microanalysis,
pages 104–106. San Francisco Press, 1985.
[49] Y. Shimakawa, Y. Kubo, T. Manako, and H. Igarashi. Neutron-diffraction study of Tl2Ba2CuO6+δ
with various Tc’s from 0 to 73 K. Phys. Rev. B, 42:10165–10171, 1990.
[50] Ruixing Liang, D.A. Bonn, and W.N. Hardy. Growth of high-quality YBCO single crystals using
BaZrO3 crucibles. Physica C, 304:105–111, 1998.
[51] Y. Shimakawa. Chemical and structural study of tetragonal and orthorhombic Tl2Ba2CuO6.
Physica C, 204:247–261, 1993.
[52] D.G. Hawthorn, S.Y. Li, M. Sutherland, E. Boaknin, R.W. Hill, C. Proust, F. Ronning,
M.A. Tanatar, J. Paglione, D. Peets, R. Liang, D.A. Bonn, W.N. Hardy, N.N. Kolesnikov,
and L. Taillefer. The quasiparticle gap in Tl2Ba2CuO6+δ from low-temperature thermal
conductivity. Phys. Rev. B in press, cond-mat/0502273, 2006.
[53] David J. Singh and Warren E. Pickett. Electronic characteristics of Tl2Ba2CuO6–Fermi surface,
positron wavefunction, electric field gradients, and transport parameters. Physica C, 203:193–
202, 1992.
[54] S. Sahrakorpi, H. Lin, R.S. Markiewicz, and A. Bansil. Effect of hole doping on the electronic
structure of Tl2201. cond-mat/0607132, 2006.
[55] Hsin Lin, S. Sahrakorpi, R.S. Markiewicz, and A. Bansil. Raising Bi-O bands above the Fermi
energy level of hole-doped Bi2Sr2CaCu2O8+δ and other cuprate superconductors. Phys. Rev.
Lett., 96:097001, 2006.
[56] T. Schmidt et al. Nucl. Instrum. Methods Phys. Res., Sect. A, 467-468:126–129, 2001.
[57] U. Fleshig et al. Nucl. Instrum. Methods Phys. Res., Sect. A, 467-468:497–481, 2001.
[58] See also http://sls.web.psi.ch/view.php/beamlines/sis/∼ .
[59] M.R. Presland, J.L. Tallon, R.G. Buckley, R.S. Liu, and N.E. Flower. General trends in oxygen
stoichiometry effects on Tc in Bi and Tl superconductors. Physica C, 176:95–105, 1991.
[60] M.R. Norman. Magnetic collective mode dispersion in high-temperature superconductors. Phys.
Rev. B, 63:092509, 2001.
[61] Todor Mishonov, Savina Savona, and Martin Stoev. LCAO model for 3D Fermi surface of high-Tc
cuprate Tl2Ba2CuO6+δ. cond-mat/0504290, 2005.
[62] O.K. Andersen, A.I. Liechtenstein, O. Jepsen, and F. Paulsen. LDA energy bands, low-energy
hamiltonians, t, t′′, t(k), and J. J. Phys. Chem. Solids, 56:1573–1591, 1995.
[63] J. Mesot, M. Randeria, M.R. Norman, A. Kaminski, H.M. Fretwell, J.C. Campuzano, H. Ding,
T. Takeuchi, T. Sato, T. Yokoya, T. Takahashi, I. Chong, T. Terashima, M. Takano, T. Mochiku,
and K. Kadowaki. Determination of the Fermi surface in high Tc superconductors by angle-
resolved photoemission spectroscopy. Phys. Rev. B, 63:224516, 2001.
[64] M.R. Norman, H. Ding, M. Randeria, J.C. Campuzano, T. Yokoya, T. Takeuchi, T. Takahashi,
T. Mochiku, K. Kadowaki, P. Guptasarma, and D.G. Hinks. Destruction of the Fermi surface
in underdoped high-Tc superconductors. Nature, 392:157–160, 1998.
[65] H. Matsui, T. Sato, S.-C. Wang, H.-B. Yang, H. Ding, T. Fujii, T. Watanabe, and A. Matsuda.
BCS-like Bogoliubov quasiparticles in high-Tc superconductors observed by angle-resolved

Page 33
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
33
photoemission spectroscopy. Phys. Rev. Lett., 90:217002, 2003.
[66] A. Kanigel, M.R. Norman, M. Randeria, U. Chatterjee, S. Souma, A. Kaminski, H.M. Fretwell,
S. Rosenkranz, M. Shi, T. Sato, T. Takahashi, Z.Z. Li, H. Raffy, K. Kadowaki, D. Hinks,
L. Ozyuzer, and J.C. Campuzano. Evolution of the pseudogap from Fermi arcs to the nodal
liquid. Nat. Phys., 2:447–451, 2006.
[67] H. Ding, M.R. Norman, J.C. Campuzano, M. Randeria, A.F. Bellman, T. Yokoya, T. Takahashi,
T. Mochiku, and K. Kadowaki. Angle-resolved photoemission spectroscopy study of the
superconducting gap anisotropy in Bi2Sr2CaCu2O8+x. Phys. Rev. B, 54:R9678–R9681, 1996.
[68] D.L. Feng, A. Damascelli, K.M. Shen, N. Motoyama, D.H. Lu, H. Eisaki, K. Shimizu, J.-I.
Shimoyama, K. Kishio, N. Kaneko, M. Greven, G.D. Gu, X.J. Zhou, C. Kim, F. Ronning,
N.P. Armitage, and Z.-X. Shen. Electronic structure of the trilayer cuprate superconductor
Bi2Sr2Ca2Cu3O10+δ. Phys. Rev. Lett., 88:107001, 2002.
[69] N.J.C. Ingle, K.M. Shen, F. Baumberger, W. Meevasana, D.H. Lu, Z.-X. Shen, A. Damascelli,
S. Nakatsuji, Z.Q. Mao, Y. Maeno, T. Kimura, and Y. Tokura. Quantitative analysis of Sr2RuO4
angle-resolved photoemission spectra: Many-body interactions in a model Fermi liquid. Phys.
Rev. B, 72:205114, 2005.
[70] A. Damascelli, D.H. Lu, and Z.-X. Shen. From mott insulator to overdoped superconductor:
Evolution of the electronic structure of cuprates studied by ARPES. J. Electron Spectrosc.
Relat. Phenom., 117-118:165–187, 2001.
[71] A. Kaminski, S. Rosenkranz, H.M. Fretwell, J. Mesot, M. Randeria, J.C. Campuzano, M.R.
Norman, Z.Z. Li, H. Raffy, T. Sato, T. Takahashi, and K. Kadowaki. Identifying the background
signal in angle-resolved photoemission spectra of high-tempertature cuprate superconductors.
Phys. Rev. B, 69:212509, 2004.
[72] K. Yang, B.P. Xie, D.W. Shen, J.F. Zhao, H.W. Ou, J. Wei, S. Wang, Y.H. Wang, D.H. Lu,
R.H. He, M. Arita, S. Qiao, A. Ino, H. Namatame, M. Taniguchi, F.Q. Xu, N. Kaneko,
H. Eisaki, and D.L. Feng. Normal state electronic structure in the heavily overdoped regime of
Bi1.74Pb0.38Sr1.88CuO6+δ single-layer cuprate superconductors. Phys. Rev. B, 73:144507, 2006.
[73] X.J. Zhou, T. Yoshida, D.-H. Lee, W.L. Yang, V. Brouet, F. Zhou, W.X. Ti, J.W. Xiong, Z.X.
Zhao, T. Sasagawa, T. Kakeshita, H. Eisaki, S. Uchida, A. Fujimori, Z. Hussain, and Z.-X.
Shen. Dichotomy between nodal and antinodal quasiparticles in underdoped (La2−xSrx)CuO4
superconductors. Phys. Rev. Lett., 92:187001, 2004.
[74] S. Sahrakorpi, M. Lindroos, R.S. Markiewicz, and A. Bansil. Evolution of midgap states and
residual three dimensionality in La2−xSrxCuO4. Phys. Rev. Lett., 95:157601, 2005.
[75] A. Kaminski, H.M. Fretwell, M.R. Norman, M. Randeria, S. Rosenkranz, U. Chatterjee, J.C.
Campuzano, J. Mesot, T. Sato, T. Takahashi, T. Terashima, M. Takano, K. Kadowaki, Z.Z. Li,
and H. Raffy. Momentum anisotropy of the scattering rate in cuprate superconductors. Phys.
Rev. B, 71:014517, 2005.
[76] A.P. Mackenzie, S.R. Julian, A.J. Diver, G.J. McMullan, M.P. Ray, G.G. Lonzarich, Y. Maeno,
S. Nishizaki, and T. Fujita. Quantum oscillations in the layered perovskite superconductor
Sr2RuO4. Phys. Rev. Lett., 76:3786–3789, 1996.
[77] C. Bergemann, S.R. Julian, A.P. Mackenzie, S. NishiZaki, and Y. Maeno. Detailed topography of
the Fermi surface of Sr2RuO4. Phys. Rev. Lett., 84:2662–2665, 2000.
[78] A. Damascelli, D.H. Lu, K.M. Shen, N.P. Armitage, R. Ronning, D.L. Feng, C. Kim, Z.-X. Shen,
T. Kimura, Y. Tokura, Z.Q. Mao, and Y. Maeno. Fermi surface, surface states and surface
reconstruction in Sr2RuO4. Phys. Rev. Lett., 85:5194–5197, 2000.
[79] A. Liebsch. Fermi surface, surface states and surface reconstruction in Sr2RuO4. Phys. Rev. Lett.,
87:239701, 2001.
[80] A. Damascelli, K.M. Shen, D.H. Lu, and Z.-X. Shen. Fermi surface, surface states and surface
reconstruction in Sr2RuO4. Phys. Rev. Lett., 87:239702, 2001.
[81] K.M. Shen, A. Damascelli, D.H. Lu, N.P. Armitage, R. Ronning, D.L. Feng, C. Kim, Z.-X. Shen,
D.J. Singh, I.I Mazin, S. Nakatsuji, Z.Q. Mao, Y. Maeno, T. Kimura, and Y. Tokura. Surface

Page 34
Tl2201 Brings Spectroscopic Probes Deep into the Overdoped Regime
34
electronic structure of Sr2RuO4. Phys. Rev. B (Rapid Comm.), 64:180502(R), 2001.
[82] M. Abdel-Jawad, M.P. Kennett, L. Balicas, A. Carrington, A.P. Mackenzie, R.H. McKenzie,
and N.E. Hussey. Anisotropic scattering and anomalous transport in a high temperature
superconductor. submitted, 2006.
[83] J.A. Slezak and J.C. Davis. APS March Meeting 2006, invited talk within the focused session ’The
electronic properties of overdoped cuprates: the clean gateway to high-Tc superconductivity’, 2006.
[84] J.E. Hoffman, K. McElroy, D.-H. Lee, K.M. Lang, H. Eisaki, S. Uchida, and J.C. Davis. Imaging
quasiparticle interference in Bi2Sr2CaCu2O8+δ. Science, 297:1148–1151, 2002.
[85] E. Abrahams and C.M. Varma. What angle-resolved photemission experiments tell about the
microscopic theory for high-temperature superconductors. Proc. Natl. Acad. Sci. U. S. A.,
97:5714–5716, 2000.
[86] L. Zhu, P.J. Hirschfeld, and D.J. Scalapino. Elastic forward scattering in the cuprate
superconducting state. Phys. Rev. B, 70:214503, 2004.
[87] K. Wakabayashi, T.M. Rice, and M. Sigrist. Enhanced coherence of antinodal quasiparticles in a
dirty d-wave superconductor. Phys. Rev. B, 72:214517, 2005.
[88] Matthias Vojta, Ying Zhang, and Subir Sachdev. Quantum phase transitions in d-wave
superconductors. Phys. Rev. Lett., 85:4940–4943, 2000.
[89] A. Kopp, A. Ghosal, and S. Chakravarty. Competing ferromagnetism in high temperature copper
oxide superconductors. cond-mat/0606431, 2006.