You are on page 1of 8

PHYSICAL REVIEW B 69, 064512 共2004兲

Effect of chemical inhomogeneity in bismuth-based copper oxide superconductors


H. Eisaki,* N. Kaneko,† D. L. Feng,‡ A. Damascelli,§ P. K. Mang, K. M. Shen, Z.-X. Shen, and M. Greven
Department of Applied Physics, Physics, and Stanford Synchrotron Radiation Laboratory, Stanford University,
Stanford, California 94305, USA
共Received 8 October 2003; revised manuscript received 4 November 2003; published 27 February 2004兲
We examine the effect on the superconducting transition temperature (T c) of chemical inhomogeneities in
Bi2 Sr2 CuO6⫹ ␦ and Bi2 Sr2 CaCu2 O8⫹ ␦ single crystals. Cation disorder at the Sr crystallographic site is inherent
in these materials and strongly affects the value of T c . Partial substitution of Sr by Ln 共Ln ⫽ La, Pr, Nd, Sm,
Eu, Gd, and Bi兲 in Bi2 Sr1.6Ln0.4CuO6⫹ ␦ results in a monotonic decrease of T c with increasing ionic radius
mismatch. By minimizing Sr site disorder at the expense of Ca site disorder, we demonstrate that the T c of
Bi2 Sr2 CaCu2 O8⫹ ␦ can be increased to 96 K. Based on these results we discuss the effects of chemical
inhomogeneity in other bulk high-temperature superconductors.

DOI: 10.1103/PhysRevB.69.064512 PACS number共s兲: 74.62.Bf, 74.62.Dh, 74.72.Hs

I. INTRODUCTION Motivated by this line of reasoning, we have examined


the effects of chemical inhomogeneity in single-layer
The possible existence of nanoscale electronic Bi2 Sr2 CuO6⫹ ␦ 共Bi2201兲 and double-layer Bi2212. Although
inhomogeneity—the propensity of charge carriers doped into widely used for surface sensitive measurements such as STM
the CuO2 plane to form nanoscale structures—has drawn 共Ref. 2兲 and angle-resolved photoemission spectroscopy
much attention in the field of high-T c superconductivity. 共ARPES兲,6 a detailed understanding of their materials prop-
Neutron-scattering studies on Nd codoped La2⫺x Srx CuO4 erties is very limited when compared to other materials such
共Nd-LSCO兲 共Ref. 1兲 and scanning tunneling microscopy/ as LSCO or YBa2 Cu3 O7⫺ ␦ .
scanning tunneling spectroscopy 共STM/STS兲 studies on The Bi-based cuprates contain excess oxygens in BiO
Bi2 Sr2 CaCu2 O8⫹ ␦ 共Bi2212兲 共Ref. 2兲 have led to suggestions planes and one can change their carrier concentration by
that such self-organization may manifest itself as one- changing the amount, ␦ . The excess oxygen would engender
dimensional ‘‘stripes’’ in Nd-LSCO, or two-dimensional a random Coulomb potential in the CuO2 planes. Besides the
‘‘patches’’ in Bi2212. In the former case, the interstripe spac- oxygen nonstoichiometry in BiO planes, there exists another
ing in the superconducting regime is reported to be approxi- source of chemical inhomogeneity which inherently exists in
mately four times the in-plane lattice constant, about 1.5 nm, typical samples. Although referred to as Bi2201 and Bi2212,
and in the latter case the patches are estimated to be 1–3 nm it is empirically known that stoichiometric Bi2.0Sr2.0CuO6⫹ ␦
across. Many theoretical studies suggest that the spatial elec- and Bi2.0Sr2.0CaCu2 O8⫹ ␦ are very difficult to synthesize,7,8
tronic inhomogeneity in the hole-doped CuO2 planes is an even in a polycrystalline form. In order to more easily
essential part of high-T c physics.3 However, at present, the form the crystal structure, one usually replaces Sr2⫹ ions
importance of, or even the existence of, generic nanoscale by trivalent ions, such as excess Bi3⫹ ions or La3⫹ ions,
electronic inhomogeneity remains controversial.4,5 forming Bi2⫹x Sr2⫺x CuO6⫹ ␦ , Bi2 Sr2⫺x Lax CuO6⫹ ␦ , and
If nanoscale electronic inhomogeneity exists in the super- Bi2⫹x Sr2⫺x CaCu2 O8⫹ ␦ . As listed up in Ref. 9, a typical
conducting cuprates, the doped holes will distribute them- Bi:Sr nonstoichiometry x for Bi2212 is around 0.1,
selves in the CuO2 planes so as to minimize their total en- which yields a T c⫽89–91 K. To our knowledge, the
ergy. In real materials, the CuO2 planes are usually highest T c reported in the literature is 95 K 关Refs. 9共g兲,
inhomogeneous due to local lattice distortions and/or the ran- 9共i兲兴. For Bi2201, T c of Bi2⫹x Sr2⫺x CuO6⫹ ␦ is around 10 K
dom Coulomb potential resulting from chemical disorder, for x(Bi)⫽0.1, whereas La-substituted Bi2201
which differs from system to system. Therefore, even if elec- (Bi2 Sr2⫺x Lax CuO6⫹ ␦ ) has a higher T c⬎30 K for x(La)
tronic inhomogeneity may itself be a genuine property of ⬇0.4. 10 Since the Sr atom is located next to the apical oxy-
doped CuO2 planes, the spatial variation of doped holes will gen which is just above the Cu atoms, the effect of Sr site
likely depend on the details of each material. For example, in 共also referred to as the A site兲 cation inhomogeneity is ex-
the framework of the stripe model,1 incommensurate spin pected to be stronger than that of the excess oxygens in BiO
and charge correlations are stabilized in Nd-LSCO by the planes. Note that BiO planes are located relatively far away
long-range distortion of the CuO6 octahedra in the low- from CuO2 planes, with SrO planes in between.
temperature tetragonal phase, which creates one-dimensional In this study, we evaluate the effect of chemical inhomog-
potential wells. For Bi2212, it is argued that the random eniety in the Bi-based cuprates. For Bi2201, we have grown
Coulomb potential caused by excess oxygen atoms in the a series of Bi2 Sr1.6Ln0.4CuO6⫹ ␦ crystals with various triva-
BiO planes pins the doped holes, thus creating patch-shaped lent rare-earth 共Ln兲 ions. In this series, the magnitude of the
inhomogeneities.2 These observations suggest that electronic local lattice distortion can be changed systematically by
and chemical inhomogeneity are inseparable from each making use of the different ionic radii of the substituted Ln
other, and that the understanding of the latter is imperative ions. We find that T c monotonically decreases with increas-
for an understanding of the former. ing ionic radius mismatch. For Bi2212, a series of

0163-1829/2004/69共6兲/064512共8兲/$22.50 69 064512-1 ©2004 The American Physical Society


H. EISAKI et al. PHYSICAL REVIEW B 69, 064512 共2004兲

TABLE I. Sample preparation conditions and crystal compositions derived from the ICP analysis for Bi2212 single crystals. The ICP
analysis on the third sample in the table was done on cleaved, single-phase samples from an ingot containing a small amount of SrCuO2
secondary phase.

Annealing condition
Nominal composition ICP results Bi:Sr ratio Growth atmosphere UD OP OD

Bi2.2Sr1.8CaCu2 O8⫹ ␦ Bi2.19Sr1.86Ca1.07Cu2 O8⫹ ␦ 1.178 1 atm 700 °C 750 °C 400 °C


O2 Air O2 O2
Bi2.04Sr1.96CaCu2 O8⫹ ␦ Bi2.06Sr1.93Ca0.96Cu2 O8⫹ ␦ 1.069 3 atm 650 °C 500 °C as-grown
O2 Ar Ar
Bi2.00Sr2.00CaCu2 O8⫹ ␦ Bi2.06Sr2.04Ca0.87Cu2 O8⫹ ␦ 1.008 1 atm 650 °C 500 °C 650 °C
O2 :Ar⫽7:93 Ar Ar O2
Bi2.00Sr2.00Ca0.92Y0.08Cu2 O8⫹ ␦ Bi2.02Sr2.01Ca0.85Y0.08Cu2 O8⫹ ␦ 1.004 1 atm 650 °C 400 °C 500 °C
O2 :Ar⫽20:80 Ar Ar O2

Bi2⫹x Sr2⫺x CaCu2 O8⫹ ␦ crystals with varying values of x growth speed of 0.3–0.4 mm/h for Bi2201 and 0.15–0.2
were grown in order to evaluate the effect of Bi:Sr nonsto- mm/h for Bi2212. The growth atmospheres adopted for the
ichiometry. In addition, we also have grown Bi2212 growth are listed in Table I. Bi2201 single crystals
Bi2 Sr2 Ca1⫺y Yy Cu2 O8⫹ ␦ , and found that substitution of Y were grown in 1 atm of flowing O2 .
for Ca site helps to enforce Bi:Sr stoichiometry and to raise The growth condition for the x⫽0 Bi2212 sample was
T c to 96 K for y⫽0.08. more stringent than for the other samples. In order to obtain
Our results demonstrate that the cation disorder, in par- homogeneous polycrystalline feed rods, a mixture of starting
ticular that located at the Sr site, significantly affects the powders with the stoichiometric ratio Bi:Sr:Ca:Cu
maximum attainable T c (T c,max) in the Bi-based supercon- ⫽2:2:1:2 was calcinated in stages, at temperatures increas-
ductors. In order to explain our results we use a conceptual ing from 770 °C to 870 °C in 10 °C increments, with inter-
hierarchy that classifies and ranks the principal kinds of mediate grindings. The final calcination temperature
chemical disorder possible in these systems. We then extend (870 °C) was set to be just below the composition’s melting
our arguments to other cuprates to examine whether a gen- temperature (875 °C). The duration of each calcination was
eral trend exists in the hole-doped high-T c superconductors. about 20 h. To avoid possible compositional fluctuation in
This paper is organized as follows: Section II contains the feed rod, instead of premelting, the feed rod was sintered
detailed information about sample preparation and character- four times in the floating-zone furnace at a speed of 50
ization. The experimental results are presented in Sec. III and mm/h. This process allowed us to obtain dense feed rods,
discussed in Sec. IV, while the effects of disorder in other ⬇95% of the ideal density. The atmosphere required for
cuprates are addressed in Sec. V. stable crystal growth was (7⫾3)% O2 , a range much nar-
rower than for nonstoichiometric or Y-doped Bi2212. The
II. SAMPLE PREPARATION grown crystal rod contained small amounts of a single-
crystalline SrCuO2 secondary phase, indicating that the
Single crystals of Bi2201 and Bi2212 were grown using sample still suffers from Bi:Sr nonstoichiometry. Single-
the traveling-solvent floating-zone technique, which is now phase Bi2212 single crystals could be cleaved from the
the preferred method for synthesizing high-purity single grown rod. No traces of impurity phases were found in other
crystals of many transition-metal oxides. This technique al- compositions.
lows for greater control of the growth conditions than is pos- Inductively coupled plasma 共ICP兲 spectroscopy was car-
sible either by standard solid-state reactions or by the flux ried out to determine the chemical compositions of the crys-
method. tals. Additional electron-probe microanalysis was also car-
Powders of Bi2 O3 , SrCO3 , CaCO3 , Ln2 O3 (Ln⫽La, Pr, ried out on the Bi2201 crystals. The results confirm that the
Nd, Sm, Eu, Gd, Y兲, and CuO 共all of 99.99% or higher actual compositions follow the nominal compositions, as
purity兲 were well dried and mixed in the desired cation ratio listed in Table I for Bi2212. Hereafter, we basically denote
(Bi:Sr:Ln:Cu⫽2:2⫺x:x:1 for Bi2201 and Bi:Sr:Ca:Cu⫽2 the samples by nominal composition to avoid confusion. In
⫹x:2⫺x:1:2 for Bi2212兲, and then repeatedly calcinated at order to determine the maximum T c for each cation compo-
about 800 °C with intermediate grinding. Eventually, the sition 共i.e., to achieve optimal hole concentration兲, the
powder was finely ground and formed into a 100 mm long Bi2212 samples were annealed at various temperatures and
rod with a diameter of 5 mm. The crystal growth was per- oxygen partial pressures using a tube furnace equipped with
formed using a Crystal Systems Inc. infrared radiation fur- an oxygen monitor and a sample transfer arm, which allowed
nace equipped with four 150 W halogen lamps. Except for us to rapidly quench the annealed samples from high tem-
nearly stoichiometric (x⫽0) Bi2212, the rods were pre- peratures within a closed environment. This procedure en-
melted at 18 mm/h to form dense feed rods. The crystal sures that the variation of T c among the samples is primarily
growth was carried out without the use of a solvent and at a due to cation nonstoichiometry and not due to differing hole

064512-2
EFFECT OF CHEMICAL INHOMOGENEITY IN . . . PHYSICAL REVIEW B 69, 064512 共2004兲

series of Ln3⫹ -substituted samples grown under the same


conditions. The T c of the La-doped sample was 33 K with a
transition width of less than 1 K, a reasonable result for an
as-grown crystal with this composition.10 As shown in Fig.
1共b兲, T c decreases monotonically with increasing ⌬R, yield-
ing 29 K (Pr3⫹ ), 27.5 K (Nd3⫹ ), 23 K (Sm3⫹ ), 18 K
(Eu3⫹ ), and 12 K (Gd3⫹ ), respectively. There is no trace of
superconductivity down to 1.8 K for the Bi-substituted
sample. This is reasonable, since the ionic radius of Bi3⫹ is
the smallest of the Ln ions, and the shape of the Bi3⫹ ion
tends to be asymmetric due to the presence of a (6s) 2 lone
pair,8 which might cause additional local lattice distortions.
These observations demonstrate the strong sensitivity of T c
FIG. 1. 共a兲 Bi2 Sr1.6Ln0.4CuO6⫹ ␦ susceptibility curves, normal- to Sr site substitution in Bi2201.
ized to ⫺1 at the lowest temperature. 共b兲 T c values as a function of Although one can minimize the magnitude of the local
⌬R.
lattice distortion through the choice of the La3⫹ ion, which
has a radius similar to that of Sr2⫹ , there remains another
concentrations. Annealing conditions for obtaining optimal type of chemical disorder, that is, the random Coulomb po-
共OP兲, and typical underdoped 共UD兲 and overdoped 共OD兲 tential caused by heterovalent ion substitution. For LSCO,
samples are listed in Table I. The results for Bi2201 are on 63
Cu nuclear quadrupole resonance spin-lattice relaxation
the as-grown crystals and accordingly may not exactly reflect rate experiments have revealed signatures for an electronic
T c,max . However, by carrying out a series of annealing stud- inhomogeniety and the results have been discussed in con-
ies, we have confirmed that the systematic change of T c nection with the random distribution of Sr2⫹ ions.12 Consid-
among our samples is not due to different hole concentra- ering that both LSCO and Bi2201 contain A-site chemical
tions, but due to the different Ln ions. disorder, it seems likely that the disorder would affect the
Superconducting transition temperatures were determined electronic properties of Bi2201 to a degree comparable to
by ac susceptibility measurements using a Quantum Design that observed in LSCO, and much more strongly than for the
Physical Properties Measurement System. The transition structurally similar material Tl2 Ba2 CuO6⫹ ␦ 共Tl2201兲 which
temperatures reported here correspond to the onset of a dia- is believed to be essentially free of A-site disorder. We will
magnetic signal. We note that the different definition of T c discuss this issue in more detail below.
共such as the intercept between the superconducting transition
slope and the ␹ ⫽0 axis兲 does not affect our conclusions due
to the sharp superconducting transition 共less than 2 K for B. Results for Bi2212
most samples兲, as shown below. Although it is hard to deter- Since cation substitution at the Sr site has such a dramatic
mine the exact superconducting fraction due to the demag- effect on T c in Bi2201, in particular for Ln⫽Bi, one might
netization factor of the plate-shaped crystals, the magnitude expect to find similar results for Bi2212. To verify this, we
of the superconducting signal suggests the bulk superconduc- have grown single crystals of Bi2⫹x Sr2⫺x CaCu2 O8⫹ ␦ . By
tivity of the grown samples. No appreciable differences were adopting the methods described in Sec. II, we have managed
observed between different samples from the same growth, to grow single crystals over the range 0.0⬍x⭐0.2. In Figs.
or between different crystals prepared under identical condi- 2共a兲–2共c兲, we present magnetic susceptibility data for three
tions, an indication of the macroscopic homogeneity of different crystals with compositions x⫽0.2, 0.04, and ⯝0,
the crystals and the reproducibility of our sample growth respectively „⑀ in the chemical formula for the nominal x
process. ⫽0 sample 关Fig. 2共c兲兴 implies the presence of residual non-
stoichiometry in our sample, as discussed in the preceding
III. EXPERIMENTAL RESULTS section…. In the figures, OP indicates optimally doped
samples, which possess T c,max for a given cation composi-
A. Results for Bi2201
tion. Representative data for UD and OD samples, which
We have grown a series of Ln3⫹ -substituted were obtained by reducing and oxidizing OP samples, are
Bi2 Sr1.6Ln0.4CuO6⫹ ␦ single crystals with Ln⫽La, Pr, Nd, also plotted to demonstrate successful control of the hole
Sm, Eu, Gd, Bi. The ionic radius of Ln3⫹ ions, R Ln , concentration over a wide range.
decreases monotonically with increasing atomic number, As the Bi:Sr ratio approaches 1:1, T c,max increases from
1.14 Å(La3⫹ ), 1.06 Å(Pr3⫹ ), 1.04 Å(Nd3⫹ ), 82.4 K for x⫽0.2 to 91.4 K for x⫽0.07 共not shown兲, 92.6 K
3⫹ 3⫹
1.00 Å(Sm ), 0.98 Å(Eu ), 0.97 Å(Gd3⫹ ), and for x⫽0.04, and eventually to 94.0 K for the sample closest
3⫹ 11
0.96 Å(Bi ). Accordingly, by introducing different Ln to the stoichiometric composition that we could grow. We
ions, we can systematically change the ionic radius mismatch note that most of the samples studied in the literature contain
between Sr2⫹ (1.12 Å) and Ln3⫹ . We use this mismatch, a nonstoichiometry of x⬃0.1 with T c⫽89–91 K, 9 consistent
defined as ⌬R⬅ 兩 R Sr⫺R Ln兩 , to quantify the magnitude of the with the present results.
local lattice distortion around Ln3⫹ ions. Although T c of Bi2212 can be raised by trying to enforce
In Fig. 1共a兲, we show magnetic susceptibility data for a Bi:Sr stoichiometry, the preparation of nearly stoichiometric

064512-3
H. EISAKI et al. PHYSICAL REVIEW B 69, 064512 共2004兲

IV. DISCUSSION

The effect on T c of structural distortions associated with


cation substitution has been extensively studied in LSCO-
based materials,13 and it is established that T c strongly de-
pends on the A-site 共La site兲 ionic radius mismatch. For in-
stance, Attfield et al. used simultaneous cosubstitution of
several alkaline earth and Ln ions to hold the average A-site
ionic radius constant while systematically controlling the
variance of the A-site ionic radius, and found that T c is af-
fected not just by the average radius, but also by the degree
of disorder 共the variance兲 at that crystallographic site. Our
study of single-layer Bi2201 continues this line of inquiry to
a different superconducting material and demonstrates a
similar sensitivity of T c to A-site disorder. We note that our
results qualitatively agree with those of a previous study on
polycrystalline Bi2201 samples.14
One can see the same trend in Bi2212 crystals with vary-
ing degrees of chemical inhomogeniety. As expected, we find
that T c is strongly dependent on the A-site disorder intro-
duced by the Bi:Sr nonstoichiometry. Furthermore, we also
demonstrate that by the seemingly counterintuitive method
of introducing additional Y3⫹ ions, and hence a new type of
disorder, we can raise T c,max to 96 K while minimizing A-site
disorder. This suggests that, although the minimization of
FIG. 2. Bi2⫹x Sr2⫺x Ca1⫺y Yy Cu2 O8⫹ ␦ susceptibility curves, nor- chemical disorder is important for raising T c , different types
malized to ⫺1 at the lowest temperature. Data for optimally doped of disorder are not equally harmful. This is consistent with
共OP兲, underdoped 共UD兲, and overdoped 共OD兲 samples are indicated the observation15 that by carefully controlling disorder in the
for each cation composition by closed circles, open squares, and triple-layer material TlBa2 Ca2 Cu3 O9⫹ ␦ 共Tl1223兲 T c can be
open triangles, respectively.
raised from ⬃120 K to 133.5 K, a new record for that sys-
tem, and that Ba site 共the A site in this system兲 cation disor-
samples becomes much more difficult than when nonstoichi- der 共deficiency兲 has the strongest effect on T c .
ometry is allowed. This could be due to a greater stability of Numerous experiments on Bi2212 have suggested non-
the crystal structure when it contains additional positive uniformity in its electronic properties. These include broad
charges, which are usually introduced by allowing the linewidths seen in inelastic neutron-scattering experiments,16
Bi3⫹ :Sr2⫹ ratio to be larger than 1, as discussed in Ref. 8. If residual low-energy excitations in the superconducting state
this is indeed the case, one might expect to be able to syn- observed in penetration depth measurements,17 finite spin
thesize higher-T c,max samples more easily by introducing ex- susceptibility at low temperatures observed in NMR
tra positive charges via cation substitution that causes disor- studies,18 and short quasiparticle lifetimes detected by com-
der less severe than substitution of Bi3⫹ ions at the Sr site. plex conductivity experiments.19 The most recent of these are
In the case of Bi2201 we observed that the substitution of STM/STS measurements2 that purport to directly image
additional Ln atoms can eliminate excess Bi atoms from the patch-shaped, electronically inhomogeneous regions. The
unfavorable Sr site position, effectively lowering the magni- Bi:Sr nonstoichiometry which inherently exists in most
tude of disorder and raising T c . In the double-layer material samples may be partially responsible for these experimental
Bi2212, there is an additional crystallographic site, the Ca observations.
site located between the CuO2 planes, which can also accept Although the present results do not directly prove the
trivalent dopant ions. One might expect Ln3⫹ ions at the Ca presence of nanoscale electronic inhomogeneity, they can be
site to be a weaker type of disorder than Bi3⫹ ions at the Sr taken as a circumstantial supporting evidence, since they
site, since there are no apical oxygens in the Ca planes that successfully prove the existence of nanoscale chemical inho-
could couple to Cu atoms in the CuO2 planes. mogeneity which potentially pins down electronic inhomo-
To test this idea we have also grown Y-substituted geneity. To explain their STM/STS results, Pan et al.2 at-
Bi2 Sr2 Ca1⫺y Yy Cu2 O8⫹ ␦ crystals. We find that this com- tribute the source of pinning centers to excess oxygen in the
pound is as easy to prepare as ordinary 共nonstoichiometric兲 BiO planes. Although the overall framework addressed by
Bi2212, and that for Y-Bi2212 the Bi:Sr ratio indeed tends to Pan et al. should still hold, we consider that the Bi ions on
be stoichiometric. Furthermore, as shown in Fig. 2共d兲, T c,max the Sr site are more effective as pinning centers since they
for the y⫽0.08 sample was increased to 96.0 K, a value are closer to the CuO2 planes and affect T c more directly.
higher than for any other Bi2212 sample reported in the Indeed, assuming a random distribution of Bi ions on the Sr
literature.9 We also grew y⫽0.10 and y⫽0.12 samples and site and a nonstoichiometry of x⫽0.10, the average separa-
confirmed that T c,max⬎95 K in both cases. tion between Bi ions is ⬃1 –2 nm, comparable with the

064512-4
EFFECT OF CHEMICAL INHOMOGENEITY IN . . . PHYSICAL REVIEW B 69, 064512 共2004兲

FIG. 3. 共Color兲 Classification of bulk high-T c cuprates in terms of the disorder site and the number of CuO2 layers. Materials belonging
to the same family are indicated by the same color. Halogen family denotes (Ca, Sr) 2 CuO2 A2 (A⫽Cl, F兲 based materials. Bi family denotes
Bi22 (n⫺1)n(n⫽1,2,3). Pb family denotes Pb2 Sr2 Can⫺1 Cun⫹1 Oz . 1L Tl family denotes one-Tl-layer cuprates, TlBa2 Can⫺1 Cun O3⫹2n⫹ ␦ .
2L Tl family denotes two-Tl-layer cuprates, Tl2 Ba2 Can⫺1 Cun O4⫹2n⫹ ␦ . La family denotes La2 Can⫺1 Cun O2⫹2n . YBCO family denotes
LnBa2 Cu3 O6⫹ ␦ . Hg family denotes HgBa2 Can⫺1 Cun O2⫹2n⫹ ␦ . The transition temperatures are compiled from works listed in Ref. 21.

length scale observed in the STM/STS studies. V. DISORDER EFFECTS IN THE CUPRATES
Recent 89Y NMR experiments on yttrium barium copper
The two main lessons to be learned from our Bi2201 and
oxide 共YBCO兲 indicate that the spatial inhomogeniety in this Bi2212 case studies are that 共1兲 chemical inhomogeneity af-
system is much less severe than in LSCO or Bi2212.5 This is fects T c,max and that 共2兲 the effect of disorder differs depend-
reasonable since the latter two systems exhibit a much higher ing on its location. In the following, we attempt to classify
degree of disorder located at the A site 关La site 共LSCO兲 and the various sites at which chemical disorder is possible and
Sr site 共Bi2212兲兴, whereas YBCO 共Ba site兲 is thought to be categorize other superconducting families on the basis of
free from such cation disorder. Indeed, recent penetration which kind of disorder is prevalent in each system.
depth measurements on the YBCO variant In Fig. 3, we classify 25 cuprate superconductors based
Nd1⫹x Ba2⫺x Cu3 O7⫺ ␦ , with cationic disorder at the Ba site, on the pattern of the chemical disorder and the number of
demonstrate that the superconducting properties of this sys- CuO2 planes in the unit cell.21 In the first row, we illustrate
tem change quite sensitively with the degree of Nd/Ba three possible locations of chemical disorder relative to the
nonstoichiometry.20 CuO5 pyramids in multilayer materials, or to the CuO6 octa-

064512-5
H. EISAKI et al. PHYSICAL REVIEW B 69, 064512 共2004兲

hedra in single-layer materials. Pattern 共a兲 corresponds to the This trend is closely obeyed when one concentrates on the
Bi:Sr nonstoichiometry in Bi2201 and Bi2212, or Sr2⫹ ions variation within a single family of materials, each denoted by
doped into the La site in LSCO, referred to as A-site disorder a different color in the chart. For example, across the first
so far. The disorder is located next to the apical oxygen. row, oxygen-intercalated La2 CuO4⫹ ␦ located in 共c-1兲 has
Pattern 共b兲 corresponds to Y3⫹ substitution for Ca2⫹ in higher T c,max than Sr-substituted La2 CuO4 共a-1兲. Down the
Bi2212 and represents disorder located next to the CuO2 column, T c,max of the bilayer La-based system
plane, but at a position where there are no apical oxygen La2⫺x Srx CaCu2 O6 共a-2兲 is higher than that of its single-layer
atoms with which to bond. There is no corresponding 共b兲 site cousin. The classification suggests a negative correlation be-
in single-layer materials. Pattern 共c兲 disorder is further away tween the effective magnitude of chemical disorder and
from the CuO2 plane. We include excess oxygen ␦ in Bi- and T c,max . Additional remarks are made in Refs. 22–24.
Tl-based cuprates, oxygen defects in CuO chains in We note that one may also have to consider other factors
YBa2 Cu3 O7⫺ ␦ , and Hg deficiency y as well as excess oxy- which characterize the global material properties and are
gen ␦ in Hg1⫺y Ba2 Can⫺1 Cun O2n⫹2⫹ ␦ in this category. We likely to play a significant role as well in determining T c,max ,
note that the materials are cataloged based on the primary such as Madelung potential,26 bond valence sum,25 band
form of disorder that they are believed to exhibit. structure,27 block layer,28 multilayer,29 etc. Although the
As demonstrated in the present case study, the effect of present scheme is somewhat oversimplified and does not
the chemical disorder is expected to be stronger for pattern take account of these parameters, we believe it serves as a
共a兲 than for pattern 共b兲, which is reasonable considering the useful framework within which to consider the chemical dis-
role of the apical oxygen atom in passing on the effect of order effects prevalent in these materials, at least some of
disorder to nearby Cu atoms. First, the random Coulomb which, if ignored, have the potential to lead to the misinter-
potential caused by type 共a兲 disorder changes the energy lev- pretation of experimental data. Finally, similar to previous
els of the apical oxygen orbitals, which can be transmitted to work on Tl1223 共Ref. 15兲 and to the present work on
CuO2 planes through the hybridization between the apical Bi2212, it might be possible to raise T c,max of certain other
O(2p z ) orbital and the Cu(3d r 2 ⫺3z 2 ) orbital. Second, the cuprates by minimizing the effects of chemical disorder.
displacement of the apical oxygen caused by type 共a兲 disor-
der brings about a local lattice distortion to the CuO2 planes. VI. SUMMARY
The effect of pattern 共c兲 disorder is expected to be weakest
since the disorder is located relatively far away from the In summary, we present case studies of the effects of
CuO2 plane. chemical disorder on the superconducting transition tempera-
The number of CuO2 planes per unit cell may be regarded ture of the single-layer and double-layer Bi-based cuprate
as another parameter that, in effect, determines the magni- superconductors. We find that the superconducting transition
tude of the chemical disorder. As demonstrated in Bi2212, temperature of Bi2212 can be increased up to 96 K by low-
multilayer materials can accommodate heterovalent ions by ering the impact of Sr site disorder, the primary type of dis-
making use of type 共b兲 substitution whose effect on T c was order inherent to the bismuth family of materials, at the ex-
seen to be weaker than type 共a兲 substitution. Furthermore, the pense of Ca site disorder. Based on these experimental
space between the CuO2 planes forming a multilayer may results, we present a qualitative hierarchy of possible disor-
buffer the impact of the disorder. For instance, in single-layer der sites, and then proceed to categorize the hole-doped
materials, any displacement of the ‘‘upper’’ apical oxygen in high-temperature superconductors on that basis.
a CuO6 octahedron creates stress in the CuO2 plane because
ACKNOWLEDGMENTS
the motion of the octahedron is constrained by the ‘‘lower’’
apical oxygen. Double-layer materials contain CuO5 pyra- We thank G. Blumberg, J. Burgy, E. Dagotto, J. C. Davis,
mids rather than CuO6 octahedra, and the separation between D. S. Dessau, T. H. Geballe, E. W. Hudson, A. Iyo, A. Ka-
CuO2 planes relieves this stress, reducing the effects of type pitulnik, G. Kinoda, S. A. Kivelson, R. B. Laughlin, B. Y.
共a兲 disorder. This buffer zone between the outer CuO2 planes Moyzhes, D. J. Scalapino, T. Timusk, and J. M. Tranquada
is further increased in triple-layer materials, with the addi- for helpful discussions. This work was supported by the
tional benefit that the middle layer is somewhat ‘‘protected’’ NEDO grant ‘‘Nanoscale phenomena of self-organized elec-
from the direct effects of pattern 共a兲 关and 共c兲兴 disorder. trons in complex oxides-new frontiers in physics and de-
Cursory examination of Fig. 3 reveals that T c,max gener- vices,’’ by the U.S. DOE under Contract Nos. DE-FG03-
ally increases both across the rows and down the columns of 99ER45773 and DE-AC-03-76SF00515, by NSF Grants
the chart. Indeed, there is no material in 共a-1兲 which pos- Nos. DMR9985067, DMR0071897, ONR N00014-98-0195,
sesses a T c,max higher than 50 K. Furthermore, Bi2201 in and by the DOE’s Office of Basic Energy Science, Divisions
column 共a兲 has a lower T c,max than Tl2201 in column 共c兲, of Materials Sciences and Chemical Sciences, through the
despite their similar crystal structures. Similarly, T c,max of Stanford Synchrotron Radiation Laboratory 共SSRL兲. H.E.
TlBa1⫹x La1⫺x CuO5 共Tl1201兲 is lower than that of was supported by the Department of Applied Physics, Stan-
HgBa2 CuO4⫹ ␦ 共Hg1201兲. ford University.

064512-6
EFFECT OF CHEMICAL INHOMOGENEITY IN . . . PHYSICAL REVIEW B 69, 064512 共2004兲

*Present address: Nanoelectronic Research Institute, AIST, T. Fujii, and A. Matsuda, Phys. Rev. Lett. 79, 2113 共1997兲; 共g兲
Tsukuba 305-8568, Japan. ‘‘nearly perfect stoichiometric,’’ T c⫽95 K, M. Rübhausen, P.

Present address: National Metrology Institute of Japan, AIST, Guptasarma, D.G. Hinks, and M.V. Klein, Phys. Rev. B 58, 3462
Tsukuba 305-8568, Japan. 共1998兲; 共h兲 Bi2.15Sr1.85CaCu2 Oy , T c⫽90 K, J.W. Loram, J.L.

Present address: Department of Physics & Astronomy, The Uni- Luo, J.R. Cooper, W.Y. Liang, and J.L. Tallon, Physica C 341–
versity of British Columbia, 334-6224 Agricultural Rd. Vancou- 348, 831 共2000兲; 共i兲 Bi2.10Sr1.90Ca1.0Cu2.0Oy , T c⫽95 K, A. Mal-
ver, B. C. V6T 1Z1, Canada and Department of Physics, Fudan juk, B. Liang, C.T. Lin, and G.A. Emelchenko, ibid. 355, 140
University, Shanghai, China. 共2001兲.
§
Present address: Department of Physics & Astronomy, The Uni- 10
Y. Ando and T. Murayama, Phys. Rev. B 60, R6991 共1999兲; S.
versity of British Columbia, 334-6224 Agricultural Rd. Vancou- Ono, Y. Ando, T. Murayama, F.F. Balakirev, J.B. Betts, and G.S.
ver, B. C. V6T 1Z1, Canada. Boebinger, Phys. Rev. Lett. 85, 638 共2000兲; Y. Wang and N.P.
1
J.M. Tranquada, B.J. Sternlieb, J.D. Axe, Y. Nakamura, and S. Ong, Proc. Natl. Acad. Sci. U.S.A. 98, 11 091 共2001兲.
Uchida, Nature 共London兲 375, 561 共1995兲; J.M. Tranquada, J.D. 11
In this paper we use the ionic radii shown in L.H. Ahrens,
Axe, N. Ichikawa, A.R. Moodenbaugh, Y. Nakamura, and S. Geochim. Cosmochim. Acta 2, 155 共1952兲.
Uchida, Phys. Rev. Lett. 78, 338 共1997兲. 12
S. Fujiyama, Y. Ito, H. Yasuoka, and Y. Ueda, J. Phys. Soc. Jpn.
2
S.H. Pan, J.P. O’Neal, R.L. Badzey, C. Chamon, H. Ding, J.R. 66, 2864 共1997兲; P.M. Singer, A.W. Hunt, and T. Imai, Phys.
Engelbrecht, Z. Wang, H. Eisaki, S. Uchida, A.K. Guptak, K.W. Rev. Lett. 88, 047602 共2002兲; P.M. Singer, A.W. Hunt, and T.
Ng, E.W. Hudson, K.M. Lang, and J.C. Davis, Nature 共London兲 Imai, cond-mat/0302077 共unpublished兲.
413, 282 共2001兲; C. Howald, P. Fournier, and A. Kapitulnik, 13
J.D. Axe, A.H. Moudden, D. Hohlwein, D.E. Cox, K.M. Mo-
Phys. Rev. B 64, 100504 共2001兲; K.M. Lang, V. Madhavan, J.E. hanty, A.R. Moodenbaugh, and Y. Xu, Phys. Rev. Lett. 62, 2751
Hoffman, E.W. Hudson, H. Eisaki, S. Uchida, and J.C. Davis, 共1989兲; M.K. Crawford, R.L. Harlow, E.M. McCarron, W.E.
Nature 共London兲 415, 412 共2002兲; G. Kinoda, T. Hasegawa, S. Farneth, J.D. Axe, H. Chou, and Q. Huang, Phys. Rev. B 44,
Nakao, T. Hanaguri, K. Kitazawa, K. Shimizu, J. Shimoyama, 7749 共1991兲; K. Yoshida, F. Nakamura, Y. Tanaka, Y. Maeno,
and K. Kishio, Phys. Rev. B 67, 224509 共2003兲; A. Matsuda, T. and T. Fujita, Physica C 230, 371 共1994兲; B. Büchner, M.
Fujii, and T. Watababe, Physica C 388–389, 207 共2003兲; K. Breuer, A. Freimuth, and A.P. Kampf, Phys. Rev. Lett. 73, 1841
Matsuba, H. Sakata, T. Mochiku, K. Hirata, and N. Nishida, 共1994兲; B. Dabrowski, Z. Wang, K. Rogacki, J.D. Jorgensen,
ibid. 388–389, 281 共2003兲. R.L. Hitterman, J.L. Wagner, B.A. Hunter, P.G. Radaelli, and
3
J. Zaanen and O. Gunnarsson, Phys. Rev. B 40, 7391 共1989兲; S.A. D.G. Hinks, ibid. 76, 1348 共1996兲; J.P. Attfield, A.L. Kharlanov,
Kivelson, E. Fradkin, and V.J. Emery, Nature 共London兲 393, 550 and J.A. McAllister, Nature 共London兲 394, 157 共1998兲.
共1998兲; S.R. White and D.J. Scalapino, Phys. Rev. B 60, R753 14
H. Nameki, M. Kikuchi, and Y. Syono, Physica C 234, 255
共1999兲; J. Burgy, M. Mayr, V. Martin-Mayor, A. Moreo, and E. 共1994兲.
Dagotto, Phys. Rev. Lett. 87, 277202 共2001兲; I. Martin and V. 15
A. Iyo, Y. Tanaka, Y. Ishiura, M. Tokumoto, K. Tokiwa, T. Wa-
Balatsky, Physica C 357–360, 46 共2001兲. tanabe, and H. Ihara, Supercond. Sci. Technol. 14, 504 共2001兲;
4
B.G. Levi, Phys. Today 51 共6兲, 19 共1998兲; J. Orenstein and A.J. H. Kotegawa, Y. Tokunaga, K. Ishida, G.-q. Zheng, Y. Kitaoka,
Millis, Science 288, 468 共2000兲; J.W. Loram, J.L. Tallon, and A. Iyo, Y. Tanaka, and H. Ihara, Phys. Rev. B 65, 184504 共2002兲.
W.Y. Liang, cond-mat/0212461 共unpublished兲. 16
H.F. Fong, P. Bourges, Y. Sidis, L.P. Regnault, A. Ivanov, G.D.
5
J. Bobroff, H. Alloul, S. Ouazi, P. Mendels, A. Mahajan, N. Blan- Gu, N. Koshizuka, and B. Keimer, Nature 共London兲 398, 588
chard, G. Collin, V. Guillen, and J.-F. Marucco, Phys. Rev. Lett. 共1999兲.
89, 157002 共2002兲. 17
A. Maeda, T. Shibauchi, N. Kondo, K. Uchinokura, and M. Koba-
6
A. Damascelli, Z.-X. Shen, and Z. Hussain, Rev. Mod. Phys. 75, yashi, Phys. Rev. B 46, 14 234 共1992兲.
473 共2003兲. 18
M. Takigawa and D.B. Mitzi, Phys. Rev. Lett. 73, 1287 共1994兲.
7 19
Y. Ikeda, H. Ito, S. Shimomura, Y. Oue, K. Inaba, Z. Hiroi, and J. Corson, J. Orenstein, S. Oh, J. O’Donnell, and J.N. Eckstein,
M. Takano, Physica C 159, 93 共1989兲; A. Maeda, M. Hase, I. Phys. Rev. Lett. 85, 2569 共2000兲.
20
Tsukada, K. Noda, S. Takebayashi, and K. Uchinokura, Phys. M. Salluzzo, F. Palomba, G. Pica, A. Andreone, I. Miggio-Aprile,
Rev. B 41, 6418 共1990兲. O. Fischer, C. Cantoni, and D.P. Norton, Phys. Rev. Lett. 85,
8
H.E. Zandbergen, W.A. Groen, A. Smit, and G. Vantendeloo, 1116 共2000兲.
Physica C 168, 426 共1990兲. 21
共a-1兲 Ca2⫺x Nax CuO2 Cl2 , Z. Hiroi, N. Kobayashi, and M. Takano,
9
共a兲 Bi2.10Sr1.94Ca0.88Cu2.07Oy , T c⫽90 K, D.B. Mitzi, L.W. Lom- Nature 共London兲 371, 139 共1994兲; 共a-1兲 Pb2 Sr2⫺x Lax Cu2 Oz , H.
bardo, A. Kapitulnik, S.S. Laderman, and R.D. Jacowitz, Phys. Sasakura, K. Nakahigashi, S. Minamigawa, S. Nakanishi, M.
Rev. B 41, 6564 共1990兲; 共b兲 Bi2.04Sr1.93Ca0.92Cu2.02Oy , T c Kogachi, N. Fukuoka, M. Yoshikawa, S. Noguchi, K. Okuda,
⫽88 K, C. Kendziora, M.C. Martin, J. Hartge, L. Mihaly, and L. and A. Yanase, Jpn. J. Appl. Phys., Part 2 29, L583 共1990兲; 共a-1兲
Forro, ibid. 48, 3531 共1993兲; 共c兲 Bi2.03Sr1.74Ca1.08Cu2.05Oy , T c La2⫺x Srx CuO4 , H. Takagi, R.J. Cava, M. Marezio, B. Batlogg,
⫽92 K, T. Mochiku and K. Kadowaki, Physica C 235–240, 523 J.J. Krajewski, W.F. Peck, Jr., P. Bordet, and D.E. Cox, Phys.
共1994兲; 共d兲 Bi2.14Sr1.77Ca0.92Cu2.00Oy , T c⫽93 K, Y. Kotaka, T. Rev. Lett. 68, 3777 共1992兲; 共a-1兲 Bi2 Sr1⫺x Lnx CuO6⫹ ␦ , S. Ono,
Kimura, H. Ikuta, J. Shimoyama, K. Kitazawa, K. Yamafuji, K. Y. Ando, T. Murayama, F.F. Balakirev, J.B. Betts, and G.S. Boe-
Kishio, and D. Pooke, ibid. 235–240, 1529 共1994兲; 共e兲 binger, ibid. 85, 638 共2000兲; 共a-1兲 TlBa1⫹x La1⫺x CuO5 , T.
Bi2.05Sr1.79Ca0.93Cu2.00Ox , T c⫽92 K, G.D. Gu, T. Egi, N. Ko- Manako and Y. Kubo, Phys. Rev. B 50, 6402 共1994兲; 共c-1兲
shizuka, P.A. Miles, G.J. Russell, and S.J. Kennedy, ibid. 263, Sr2 CuO2 F2⫹x , M. Al-Mamouri, P.P. Edwards, C. Greaves, and
180 共1996兲; 共f兲 Bi2.1Sr1.9CaCu2 Oy , T c⫽89–90 K, T. Watanabe, M. Slaski, Nature 共London兲 369, 382 共1994兲; 共c-1兲 La2 CuO4⫹ ␦ ,

064512-7
H. EISAKI et al. PHYSICAL REVIEW B 69, 064512 共2004兲

F.C. Chou, J.H. Cho, and D.C. Johnston, Physica C 197, 303 Matsui, and S. Horiuchi, J. Phys. Soc. Jpn. 58, 2252 共1989兲兴;
共1992兲; 共c-1兲 Tl2 Ba2 CuO6⫹ ␦ , J.L. Wagner, O. Chmaissem, J.D. Bi2 Sr2 (Ln1⫺x Cex ) 2 Cu2 O10⫹y (T c,max⫽30 K for Ln⫽Gd) 关Y.
Jorgensen, D.G. Hinks, P.G. Radaelli, B.A. Hunter, and W.R. Tokura, T. Arima, H. Takagi, S. Uchida, T. Ishigaki, H. Asano,
Jensen, ibid. 277, 170 共1997兲; 共c-1兲 HgBa2 CuO4⫹ ␦ , A. Yama- R. Beyers, A.I. Nazzal, P. Lacorre, and J.B. Torrance, Nature
moto, K. Minami, W.-Z. Hu, A. Miyakita, M. Izumi, and S. 共London兲 342, 890 共1989兲兴. These materials possess CuO5 units
Tajima, Phys. Rev. B 65, 104505 共2002兲; 共a-2兲 which are sandwiched by planes containing cation disorders 关for
La2⫺x Srx CaCu2 O6 , R.J. Cava, B. Batlogg, R.B. van Dover, J.J. example, Nd 共Ce兲 O and Sr 共Nd兲 O planes for
Krajewski, J.V. Waszczak, R.M. Fleming, W.F. Peck, L.W. Nd2⫺x⫺y Cey Srx CuO4 ], and can be classified as 共a-1兲 materials.
22
Rupp, P. Marsh, A.C.W.P. James, and L.F. Schneemeyer, Nature YBa2 Cu4 O8 has a relatively low T c of 81 K, although this mate-
rial is essentially free from chemical disorder 关J. Karpinski, E.
共London兲 345, 602 共1990兲; 共a-2兲 (La1⫺x Cax ) (Ba1.75⫺x La0.25⫹x )
Kaldis, E. Jilek, S. Rusiecki, and B. Bucher, Nature 共London兲
Cu3 Oy , O. Chmaissen, J.D. Jorgensen, S. Short, A. Knizhnik, Y.
336, 660 共1988兲兴; It is known that pure YBa2 Cu4 O8 is under-
Eckstein, and H. Shaked, ibid. 397, 45 共1999兲; 共a-2兲
doped and that its T c increases to 90 K when Ca is substituted on
Bi2⫹x Sr2⫺x CaCu2 O8⫹ ␦ , Ref. 9; 共b-2兲
the Y site 关T. Miyatake, S. Gotoh, N. Koshizuka, and S. Tanaka,
Pb2 Sr2 Y1⫺x Cax Cu3 O8⫹ ␦ , Y. Koike, M. Masuzawa, T. Noji, H. ibid. 341, 41 共1989兲; T. Machi, I. Tomeno, K. Tai, N. Koshizuka,
Sunagawa, H. Kawabe, N. Kobayashi, and Y. Saito, Physica C and H. Yasuoka, Physica C 226, 227 共1994兲兴, which provides
170, 130 共1990兲; 共b-2兲 Y1⫺x Cax Ba2 Cu3 O7⫺ ␦ , J.L. Tallon, C. additional holes and makes the system optimally doped, but also
Bernhard, H. Shaked, R.L. Hitterman, and J.D. Jorgensen, Phys. increases the amount of chemical disorder. Similarly, stoichio-
Rev. B 51, 12 911 共1995兲; 共b-2兲 Bi2 Sr2 Ca1⫺x Yx Cu2 O8⫹ ␦ , this metric YBa2 Cu3 O7⫺ ␦ ( ␦ ⫽0.01, T c⫽88.5 K) is slightly over-
work; 共c-2兲 YBa2 Cu3 O7⫺ ␦ , R. Liang, D.A. Bonn, and W.N. doped, and T c,max⫽93.7 K is found for ␦ ⬇0.07 共Ref. 21兲.
Hardy, Physica C 304, 105 共1998兲; 共c-2兲 TlBa2 CaCu2 O7⫹ ␦ , S. 23
It is empirically known that materials with longer Cu-apical oxy-
Nakajima, M. Kikuchi, Y. Syono, K. Nagase, T. Oku, N. Koba- gen bonds tend to have higher values of T c. Indeed, in Fig. 3,
yashi, D. Shindo, and K. Hiraga, ibid. 170, 443 共1990兲; 共c-2兲 materials possessing type 共a兲 disorder tend to contain Sr2⫹ ions
Tl2 Ba2 CaCu2 O8⫹ ␦ , Y. Shimakawa, Y. Kubo, T. Manako, and H. 共1.12 Å兲 or La3⫹ ions 共1.14 Å兲, whereas those with type 共c兲
Igarashi, Phys. Rev. B 40, 11 400 共1989兲; 共c-2兲 disorder possess Ba2⫹ ions 共1.34 Å兲 on the A site. In the present
HgBa2 CaCu2 O6⫹ ␦ , R.L. Meng, L. Beauvais, X.N. Zhang, Z.J. framework, this tendency can be explained as follows. First,
Huang, Y.Y. Sun, Y.Y. Xue, and C.W. Chu, Physica C 216, 21 long Cu-apical oxygen bond lengths help to isolate CuO2 planes
共1993兲; 共a-3兲 Bi2⫹x Sr2⫺x Ca2 Cu3 O10⫹ ␦ , J.M. Tarascon, W.R. from the outer layers which may contain disorder. Second, due
McKinnon, P. Barboux, D.M. Hwang, B.G. Bagley, L.H. to the large difference in the ionic radii of Ba2⫹ 共1.34 Å兲 and
Greene, G.W. Hull, Y. LePage, N. Stoffel, and M. Giroud, Phys. other constituent cations such as Ln3⫹ , Bi3⫹ , or Tl3⫹ 共0.95 Å兲,
Rev. B 38, 8885 共1988兲; 共a-3兲 TlBa2⫺ ⑀ Ca2 Cu3 O9⫹ ␦ , A. Iyo, Y. their intersubstitution is usually more difficult than the Bi:Sr
Tanaka, Y. Ishiura, M. Tokumoto, K. Tokiwa, T. Watanabe, and intersubstitution seen in the Bi-based compounds, thus making
H. Ihara, Supercond. Sci. Technol. 14, 504 共2001兲; 共b-3兲 Ba-based materials resistant to type 共a兲 disorder. A good ex-
TlBa2 Ca2⫺ ⑀ Cu3 O9⫹ ␦ , same as above; 共c-3兲 ample of this concept is LnBa2 Cu3 O7 . Due to the large ionic
TlBa2 Ca2 Cu3 O9⫹ ␦ , same as above; 共c-3兲 Tl2 Ba2 Ca2 Cu3 O10⫹ ␦ , radius mismatch, Ln and Ba ions usually do not mix with each
Y. Shimakawa, Y. Kubo, T. Manako, and H. Igarashi, Phys. Rev. other. However, for Ln⫽La and Nd, Ln:Ba intersubstitution is
B 40, 11 400 共1989兲; 共c-3兲 HgBa2 Ca2 Cu3 O8⫹ ␦ , R.L. Meng, L. allowed since the mismatch is less severe. T c,max of the nonsto-
Beauvais, X.N. Zhang, Z.J. Huang, Y.Y. Sun, Y.Y. Xue, and ichiometric Ln1⫹x Ba2⫺x Cu3 O7 is lower than for other members
C.W. Chu, Physica C 216, 21 共1993兲; Besides those listed in Fig. of the series 关T. Wada, N. Suzuki, A. Maeda, T. Yabe, K. Uchi-
3, many materials may be classified within the present frame- nokura, S. Uchida, and S. Tanaka, Phys. Rev. B 39, 9126 共1989兲,
work. Examples are, 共I兲 Infinite layered material Ref. 20兴, although its average Cu-apical oxygen bond length is
(Sr1⫺x Cax ) 1⫺y CuO2 with T c,max⫽110 K 关M. Azuma, Z. Hiroi, longer.
M. Takano, Y. Bando, and Y. Takeda, Nature 共London兲 356, 775 24
For multilayer materials, T c,max typically increases up to n⫽3 and
共1992兲兴, which can be described as 共b-⬁). 共II兲 Electron doped then decreases for n⬎3 关S. Adachi et al., in Structures of High-
superconductor, Ln2⫺x Cex CuO4⫺ ␦ (Ln⫽Pr, Nd, Sm, T c,max Temperature Superconductors, edited by A. V. Narlikar 共Nova
⫽24 K for Nd兲 关Y. Tokura, H. Takagi, and S. Uchida, Nature Science, New York, 1997兲, Vol. 23, pp. 163–191兴. This may be
共London兲 337, 345 共1989兲兴 is similar to 共a-1兲 since the CuO2 because the benefit of the buffering effect due to the existence of
planes are sandwiched by two LnO planes containing Ce4⫹ ions multilayer is saturated in n⫽3.
at the Ln3⫹ site. 共III兲 T * compounds and its derivatives, 25
M.H. Whangbo and C.C. Torardi, Science 249, 1143 共1990兲; J.
Nd2⫺x⫺y Cey Srx CuO4 (T c,max⫽22 K) 关H. Sawa, S. Suzuki, M. Tallon, Physica C 168, 85 共1990兲.
26
Watanabe, J. Akimitsu, H. Matsubara, H. Watabe, S. Uchida, K. Y. Ohta, T. Tohyama, and S. Maekawa, Phys. Rev. B 43, 2968
Kokusho, H. Asano, F. Izumi, and E. Takayama-Muromachi, 共1991兲.
ibid. 337, 347 共1989兲兴; La1⫺x⫺z Ln1⫺x Srz CuO4 (T c,max⫽20 K 27
E. Pavarini, I. Dasgupta, T. Saha-Dasgupta, O. Jepsen, and O.K.
for Ln⫽Sm) 关Y. Tokura, H. Takagi, H. Watabe, H. Matsubara, S. Andersen, Phys. Rev. Lett. 87, 047003 共2001兲.
28
Uchida, K. Hiraga, T. Oku, T. Mochiku, and H. Asano, Phys. T.H. Geballe and B.Y. Moyzhes, Physica C 341–348, 1821
Rev. B 40, 2568 共1989兲兴; (Ln1⫺x Cex ) 2 (Ba1⫺y Lny ) 2 Cu3 O10⫺ ␦ 共2000兲.
(T c,max⫽33 K for Ln⫽Eu) 关H. Sawa, K. Obara, J. Akimitsu, Y. 29
S.A. Kivelson, Physica B 318, 61 共2002兲.

064512-8

You might also like