Next Article in Journal
Nanoscience and Nanotechnology, Proceedings of the INFN-LNF 2018 Conference
Next Article in Special Issue
Multi-Band Superconductivity and the Steep Band/Flat Band Scenario
Previous Article in Journal
Mössbauer Spectroscopy of Triphylite (LiFePO4) at Low Temperatures
Previous Article in Special Issue
Dzyaloshinskii–Moriya Coupling in 3d Insulators
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Probing Phase Separation and Local Lattice Distortions in Cuprates by Raman Spectroscopy

by
Efthymios Liarokapis
Department of Physics, National Technical University of Athens, 15780 Athens, Greece
Condens. Matter 2019, 4(4), 87; https://doi.org/10.3390/condmat4040087
Submission received: 25 September 2019 / Revised: 21 October 2019 / Accepted: 27 October 2019 / Published: 1 November 2019
(This article belongs to the Special Issue From cuprates to Room Temperature Superconductors)

Abstract

:
It is generally accepted that high temperature superconductors emerge when extra carriers are introduced in the parent state, which looks like a Mott insulator. Competition of the order parameters drives the system into a poorly defined pseudogap state before acquiring the normal Fermi liquid behavior with further doping. Within the low doping level, the system has the tendency for mesoscopic phase separation, which seems to be a general characteristic in all high Tc compounds, but also in the materials of colossal magnetoresistance or the relaxor ferroelectrics. In all these systems, metastable phases can be created by tuning physical variables, such as doping or pressure, and the competing order parameters can drive the compound to various states. Structural instabilities are expected at critical points and Raman spectroscopy is ideal for detecting them, since it is a very sensitive technique for detecting small lattice modifications and instabilities. In this article, phase separation and lattice distortions are examined on the most characteristic family of high temperature superconductors, the cuprates. The effect of doping or atomic substitutions on cuprates is examined concerning the induced phase separation and hydrostatic pressure for activating small local lattice distortions at the edge of lattice instability.

1. Introduction

Among the most fascinating discoveries over the years is the identification of mesoscopic phase separation (MePhS) in a number of materials, ferroelectrics [1,2], magnetoresistive manganites [3,4,5,6,7,8,9,10,11], nikelates [12,13], A15 alloys [14], and cuprates [15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52] in which case a spectacular diverse range of exotic magnetic, electronic, and crystal structures can coexist at different locations in the same crystal. In an apparently homogeneous material, it turns out to be surprisingly easy to achieve phase coexistence using a wide range of parameters by playing with the chemical composition or microstructure, increasing an electric or magnetic field, tuning the elastic strain field [53], or by photoinduced phase separation [54,55]. The coexisting phases may form robust magnetic, electronic and crystallographic textures of “mesoscopic” length scales, that is, over tens or hundreds of nanometers. By controlling an array of textured phases, analogous to those in liquid crystals, it is possible to modify locally the electronic structure and properties without atomic-scale fabrication. In manganites for example, a simple domain wall in the ferromagnetic metallic phase could spontaneously develop an insulating barrier of the charge order phase creating the ultimate spin-tunnel junction. Alternatively, in the cuprates, the superconducting domains are separated from an insulating matrix where the thread and the matrix are made of identical material. With these and numerous other ideas for potential applications, one can enter a realm of inorganic material science that is reminiscent of the complex self-organized structures seen in soft condensed matter physics or polymer science.
Structural instability appears in several systems of condensed matter and can play a key role in the characteristic properties of several other compounds (e.g., ferroelectrics), manganites, and high temperature superconductors (HTSC). The role played in the case of HTSC was revealed soon after their discovery [15] by J.C. Philips who correctly pointed out the structural instabilities that accompany either intermetallic or oxide superconductors [56]. For the old superconductors, it had already been found that, in A-15 compounds, structural instability and superconductivity are related [14]. The structural instability actually creates a softening of the lattice and is reflected in several physical properties including the associated phonon modes. In superconductors, the underlying idea of MePhS is that, in the insulator-superconductor transition, the holes introduced in the insulator matrix by the dopant are rejected, forming mesoscopic clusters of the superconducting phase (physical phase separation). The magnetic fluctuations are suppressed in the small dimensions of the randomly oriented clusters and superconductivity is stabilized. The whole process is self-organized, but the metastable phases of the system can be reached by tuning physical variables such as pressure or charge density. Although the mechanism that produces high transition temperature is still unknown, there is general agreement that understanding the normal state where several degrees of freedom (spin, charge, orbital, and lattice) compete, could clarify the situation. It is now well established that, between the Mott insulator and heavy doping charge, ordered states (stripes) are created [12,18] from the competition of the kinetic energy and Coulomb repulsion, but the exact connection of the stripes with superconductivity is not clear yet. It is, therefore, of interest to investigate the responses of the normal state to variations of external factors (temperature or pressure) or internal parameters (doping).
Interest in the role of lattice in the pairing mechanism of HTSC has increased since the discovery of superconductivity with high transition temperature (Tc) [15] in MgB2 [57,58,59,60,61,62,63], the alkali metal doped C60 [64,65,66], and, more recently, by the very high transition temperature achieved under hydrostatic pressure in sulfur hydrides [67,68,69,70]. In the case of cuprates, several experiments (e.g., [71,72,73,74,75,76,77,78,79]) suggest that phonons play a substantial role in the high Tc mechanism.
Phase separation has been observed using several methods in both pnictides and chalcogenides iron-based superconductors [80,81,82,83,84,85,86] and at the insulator to metal transition in VO2 [87,88,89,90,91,92]. It has been pointed out that this is generic feature in the proximity of topological electronic Lifshitz transitions [40,51,82] where lattice and electronic effects are both essential and key terms of the physics of these complex quantum phenomena. Therefore, the study of lattice anomalies, such as soft modes or structural instabilities, is of central interest in the investigation of the HTSC.
Raman spectroscopy is one of the tools used for investigating the characteristics of the new compounds and for providing important information on the carriers through their coupling with the phonons [93], the appearance of new phases [94,95], the anharmonic coupling of the phonons [96,97,98], the existence and the size of a gap [99], its symmetry [100], and on crystal field excitations [101], etc. In this review, cumulative results from our systematic micro-Raman studies on two families of the cuprates (yttrium and lanthanum families) are presented. Our studies on other high Tc compounds, such as the diborides [60,62,63,102] and the pnictides [103,104,105] not presented here, have shown similar results with the cuprates. The role of doping, atomic substitutions, and hydrostatic pressure are examined in detail for creating phase separated domains and structural instabilities. It should be noted that our spontaneous Raman studies cannot discriminate between static and dynamic phase separated effects, and it detects them as statistically distributed in space domains.

2. The Yttrium Family Cuprates

Detailed analysis of the lattice vibration symmetry of YBa2Cu3Ox can be found in numerous publications (see e.g., [106]). The compounds have an approximate tetragonal symmetry with a small orthorhombic distortion in the superconducting phase [107]. Because of the center of inversion, the zone-center optical phonons are either Raman or IR active. The atoms of the chains (Och, Cuch) and Y have full point group site symmetry being IR-active (B2u,3u and B1u), Ba, and apical oxygen (Oap). The CuO2 plane atoms (Cupl, Opl) have lower symmetry (C2) and they participate in three phonons of even symmetry Ag, B2g,3g and three phonons of odd symmetry B1u, B2u,3u (along the c, a, and b axis, correspondingly). The main difference between the orthorhombic and the tetragonal phase are the out-of-phase vibrations of Opl atoms, which in the former case have the Ag symmetry and in the latter case, B1g symmetry, and can be easily identified from the distinct polarization selection rules [106]. Characteristic Raman spectra are shown in Figure 1a,b for the y(cc)y and y(xx)y scattering geometries and the strong Raman active phonons of Ag symmetry along the c-axis for YBa2Cu3O7−δ are located at ≈116 cm−1 (Ba), ≈150 cm−1 (Cupl), ≈435 cm−1 (in-phase vibrations of Opl atoms), and ≈502 cm−1 (Oap).

2.1. Phase Separation from Oxygen Doping

The Ag symmetry strong modes due to the Ba and Cupl atoms vary considerably in intensity with oxygen concentration in the y(xx)y scattering configuration (Figure 1b) and Ba phonon has a strong asymmetric shape that gradually disappears in the tetragonal phase (x ≤ 6.35). The Cupl phonon shows the opposite behavior, being very weak in the fully oxygenized compounds (x ≈ 7) and gaining intensity by the orthorhombic to tetragonal (O-T) phase transition (Figure 1). In the y(cc)y scattering geometry, these phonons appear in all oxygen concentrations without considerable intensity variations (Figure 1a). In this scattering geometry, the Ba phonon is symmetric, which indicates that the asymmetry observed in the other geometry is due to its coupling with the carriers. Its intensity is modified by the amount of oxygen which reflects the carrier redistribution with doping between the BaO and the CuO2 planes. At low doping, the carriers are distributed outside the CuO2 planes screening the Ba phonon, which appears with very low intensity (or it is completely absent) in the xx scattering polarization. Exactly the opposite happens to the Cupl phonon, which is strong at low carrier concentrations and disappears with increasing doping. As carriers are transferred to the CuO2 planes they induce strong screening of the Cupl phonon, while the Ba mode gains intensity (Figure 1b).
Oxygen nonstoichiometry is the crucial parameter controlling such a diverse behavior for the compound as one follows the transition from YBa2Cu3O6 to YBa2Cu3O7. It is, therefore, natural to concentrate on the oxygen Raman active modes. The dependence of the Oap mode energy on the oxygen concentration is nonlinear [107]. Roughly described, the energy remains almost constant in the range of 6.75 < x < 7.00, is reduced in the range of 6.15 < x < 6.75, and, finally, for x < 6.15, there is a sudden drop in frequency (Figure 2a). The mode linewidth exhibits the following three minima: for maximum x values, for x ≈ 6.47, and in the range of x < 6.12. Contrary to early Raman spectra, which have shown a linear dependence of its energy with the amount of oxygen [108], the results in Figure 2 show that it remains practically constant in the range of 6.8 < x < 7.0 and, then, decreases with decreasing x (Figure 2a). More pronounced changes appear in the width of this phonon with a maximum close to x = 6.7 and then a decrease towards the two ends (Figure 2a), which is a result of the appearance of new phases, as discussed below.
In the YBa2Cu3Ox (Y123x) compounds, Cu atoms also exist outside the CuO2 planes, forming linear chains along the b-axis. As the rest of the HTSC does not have such chains, it is believed that they do not play a crucial role in superconductivity, although they influence the properties of the Y123x system. In the fully occupied chain (Y1237), all oxygen atoms are, in principle, located along the b-axis, and, as the number of oxygen atoms is decreased, there is always a tendency to form thermodynamically stable structures [109]. Two such phases with Tc at 60 K (ortho-II, for x ≈ 6.5), which correspond to alternating full and empty chains, and at 92 K (ortho-I, fully oxygenated samples) are well established appearing as characteristic plateaus in the phase diagram [110]. For lower concentrations (x ≤ 6.35), a transition takes place to an antiferromagnetic, non-superconducting phase [107]. The exact stoichiometries, where the phases appear, depend on the preparation conditions [107]. This is due to the existence of many other intermediate phases, which correspond to different combinations of full and empty chains that are examined below. Such ideal ordered microstructures can only be stabilized after a careful annealing of the materials; otherwise the disorder will make the compound appear macroscopically tetragonal [111]. The changes induced in the chains from the removal of oxygen, have a direct effect on the position of the apical oxygen, which moves away from the superconducting CuO2 planes [110]. In terms of the valences, this is an indication of redistribution of charges between the chains and the planes as oxygen is added in the system [100]. Another way to explain the phase transformation from the antiferomagnetic to the superconducting state was proposed within the phase separation model [112]. As oxygen atoms are added, and holes are transferred to the CuO2 planes, conducting microdomains are formed, which, when there are enough, are connected and the system becomes a superconductor. A further increase of holes in the planes induces another phase transition with higher Tc (92 K). The appearance of these phases is connected with the distribution of the oxygen atoms in the chains along the b-axis forming two- or three-dimensional superlattices [109]. Some of these structures have been observed (tetragonal, ortho-II, and ortho-III phases) and they correspond to x ≈ 6, x ≈ 6.5, and x ≈ 6.67 [109,110,111]. The effect of the removal of the chain oxygen on the Raman spectra is shown in Figure 1. As discussed above, on the one hand, the relative intensity of the low frequency modes (Ba and Cupl) depends strongly on the amount of oxygen (mainly in the y(xx)y scattering geometry), but their energy is not modified appreciably. On the other hand, the oxygen content affects very strongly both the energy and width of the apical oxygen and the in-phase phonons (Figure 2). This is due to the appearance of new phases, as resonance examination of these oxygen modes in the excitation energy range of 1.6–2.9 eV, for the tetragonal and orthorhombic Y123x system, has revealed [113]. Concerning the Ag symmetry apical oxygen phonon, it has been found to show a strong resonance in the IR excitation for the tetragonal case, with additional modes appearing, which are absent when the excitation is in the visible region of spectrum [114]. Work carried out by Iliev et al. [94] has shown a very characteristic variation of the apical Ag phonon when the excitation wavelength varied from blue (482.5 nm) to red (647.1 nm) for the YBa2Cu3O6.4 compound. Our investigations confirmed this by showing that, in the fully oxygenated ortho-I phase, the Oap mode does not shift with varying excitation wavelength (Figure 3a) and it is modified considerably for the underdoped compound (Figure 3b) due to the different cross sections of the coexisting phases, with the lower Tc phases being in resonance with lower energy excitations [113,114].
Other measurements also suggest the coexistence of phases [115]. Since the apical oxygen mode is very sensitive to the chain ordering [93,116] and its width shows minimum values around x ≈ 6, x ≈ 6.47, and x ≈ 7 (Figure 1a), one could assume that at least three phases are present. But the resonance measurements have revealed the existence of four peaks located at 502 cm−1, 493 cm−1, 486 cm−1, and 476 cm−1 that correspond to four coexisting phases. The higher and the lower energy peaks certainly correspond to ortho-I and the tetragonal x ≈ 6 phases. From our measurements and other measurements on x ≈ 6.5, it is clear beyond any doubt that the 493 cm−1 is a superstructure peak due to the ortho-II phase (doubling of unit cell with half occupation of the chains). In order to delineate the origin of the 486 cm−1 peak, we consistently carried out a deconvolution of all Raman spectra from 30 samples of varying oxygen concentrations with 16O and 23 samples with 18O, and the results are presented in Figure 4 (similar results exist for 18O in [98]). It appears that the 486 cm−1 peak corresponds to a phase in the tetragonal doping region. From the similarity in energy, the superstructure (S) peak at 486 cm−1 could correspond to the breathing (0.5, 0.5, 0) mode [117], but then it should peak around x ≈ 6.25, contrary to the deconvolution results (Figure 4). A better guess is the 2 2 ao × 2 2 ao × c diagonal superlattice phase based on the amount of oxygen [118,119,120]. In order to extract more information from the Raman spectra, as shown in Figure 5, we plotted a histogram of the peak positions for the apex and in-phase modes. A preference appears for certain wavenumbers, which correspond (from Figure 2) to the phases indicated in Figure 5. The ortho-I and ortho-II phases are clear, while the superstructure S peak seems to correspond to x ≈ 1/8 (2 2 ao × 2 2 ao × c) phase. Furthermore, there are indications of other phases for wavenumbers pointing to an ortho-III (x = 6.67) or even higher phases [121]. In a systematic imaging of the spatial ordering of the oxygen chains in YBa2Cu3O6+y, by micro X-ray diffraction measurements, the phase separated regions have been detected [122,123,124]. For y = 0.33, the ortho-II phase was clearly detected embedded in more disordered regions [124]. For y = 0.67, nano-puddles corresponding to y = 0.5 + 1/8 were found corresponding to the ortho-VIII puddles of alternating five filled and three empty oxygen wires every eight rows [122]. Other networks were created by thermal manipulation [123]. Unfortunately, no Raman data exist on the same compounds for comparison, but our studies agree with these results and at least the one-eighth oxygen ordering seems to be a preferential oxygen ordering for both the YBa2Cu3O6 and the YBa2Cu3O6.5 stable phases. Furthermore, the regions assigned to OIII phase in Figure 5 possibly consist of several other phases similar to the ones discovered by X-ray imaging [122,123]. This can explain the less pronounced nature of the OIII peaks in Figure 5 for both the apex and the in-phase modes. Finally, there is another peak in the overdoped region, as strongly indicated from the in-phase mode and discussed below.
As seen in Figure 2b, the in-phase Opl mode energy and linewidth values remain constant in the range of 6.00 < x < 6.40. At higher doping levels, the energy monotonically decreases, while the linewidth increases almost linearly and, finally, it softens by ~8 cm−1 close to optimal doping (Figure 2b). An analysis of the Raman measurements shows that microcrystallites of the two different phases could be found within the same compound with the in-phase phonon energy either corresponding to the optimally or the overdoped concentration. This is the reason for the increased error bars (the statistical distribution from measurements from several microcrystallites) for the overdoped data of in-phase phonon energy in Figure 2b. The EXAFS [125] and neutron [126] scattering measurements have shown that, at exactly this oxygen concentration, there is a sudden modification in the buckling of the CuO2 planes with a reduction in Tc [127]. Apparently, this softening is related to the change in the buckling. Therefore, we conclude that two phases coexist for x ≥ 6.94 and only the relative amount of each phase varies with oxygen content [127]. It should be stressed that the miscibility gap, seen as a split-off peak below 435 cm−1 in Figure 5 (in-phase mode), is observed even between microcrystals of the same sample, i.e., x = 6.940. This fact underlines the martensitic character of the phase transition. A similar softening has been observed for the 18O samples [98], and in samples where the La has been substituted in half by another rare earth (La0.5R0.5Ba2Cu3Ox) [128], but not in the homologous series Pr0.5R0.5Ba2Cu3Ox [129]. It has been shown that the softening is not related to the amount of doping, as the alternative way of doping by Ca substitution (YxCa1−xBa2Cu3Oy) does not induce such a softening ([130] and following session). As La has the tendency of attracting more oxygen atoms, the softening must be attributed to structural modifications induced by the excess of oxygen atoms.

2.2. Phase Separation from Atomic Substitutions

The B1g mode is sensitive to atomic substitutions for yttrium and its broadening was used to measure the superconducting gap by tuning the phonon energy across the gap energy for carefully chosen rare earth substitutions [131]. By studying a series of rare earth substitutions, La0.5R0.5Ba2Cu3O7−δ [128], R0.5Pr0.5Ba2Cu3O7−δ [129], Y1−xLaxBa2Cu3O7−δ [132], and Y1−xPrxBa2Cu3O7−δ [133], it has also been found that the B1g mode splits into two or three bands, depending on the difference of the atomic radii of substituting species and this is related to the superconducting property. On the one hand, in all these cases, the XRD studies could not discriminate more than one component, although modifications in the interatomic distances at critical differences of the ion size have been found [128]. This indicates that any phase separation in those systems occurred at distances smaller than the coherence length of XRD. On the other hand, Raman measurements with smaller coherent length could clearly detect the multiphase structure of the compounds [128,129,132,133]. Characteristic spectra for the evolution of the strong modes with rare earth substitution for Y1−xLaxBa2Cu3O7−δ are shown in Figure 6 [132]. Similar Raman spectra have been found for the other substitutions [128,129,133].
Concerning the other modes, the apex oxygen mode in Figure 6 shifts continuously depending on the average ionic radius of the substituting rare earth elements, in agreement with the results from the system RBa2Cu3O7−δ [134]. The in-phase mode is more or less insensitive to rare earth (Figure 6), although there might be variations, as much as 7 to 8 cm−1, if the oxygen content varies slightly below or above the optimal doping value (Figure 6). In addition, the Cupl phonon is not modified by the substitution. The Ba phonon also does not vary in energy, but its intensity has been strongly quenched in the xx polarization by Pr substitution [128,129], which is due to the partial substitution of Pr for Ba, explaining the unexpected strong reduction in Tc in that case [133]. The most sensitive is the B1g mode that follows one- or two-mode behavior [135] depending on the radius difference of the two atoms [128,129]. As seen in Figure 6b, this phonon splits into two or three bands, which apparently correspond to the end members and an intermediate one. In the case of Y1−xPrxBa2CuO7−δ, the transition temperature decreases as Pr substitutes for yttrium up to 60%, and above that, the compound does not superconduct (Figure 7). Plotting the relative amount of the various phases as induced by the B1g phonon mode, it is observed that Tc does not follow the evolution of the relative amount of the intermediate mixed state, which is the leading phase up to 80% Pr content, where the compound is not superconducting (Figure 7). This is proof that both the pure Pr123 (Y1−xPrxBa2CuO7−δ) and the intermediate mixed phases are not superconducting and the superconductivity in these compounds is only due to the percolation via the Y123 (YBa2Cu3O7−δ) phase.
Because, in YBa2CuOx, there is no way to add more oxygen beyond x = 7, overdoping can be achieved by Ca substitutes for yttrium. This substitution has two effects; an increase of holes creating overdoping, and creation of strains from the size difference between the Ca and Y ions. Furthermore, the local electric field in the CuO2 planes from the Y/Ca and Ba sites is modified, which could possibly affect the phase separation observed in the rare earth substitutions for yttrium [128,129]. The results of the Ca substitution on the strong phonon modes for the y(xx)y scattering polarization are shown in Figure 8. The Ag-symmetry modes, due to Ba and Cu2, are not modified considerably by the Ca substitution (Figure 8) implying that Ca does not substitute for isovalent Ba. The relative intensity of the Cu2 to Ba in the xx polarization is also not modified because the excess doping continues screening the Cu2 atoms [130]. In the Ca substitution for the Y overdoped region, the energy of the Oap phonon decreases (Figure 2a) following the bell-shaped trend of Tc. The width of the phonon remains close to its value for the ortho-I phase (Figure 2a), verifying that the substantial increase in the width with oxygen doping in pure YBa2Cu3Ox is due to the ordering of the chain oxygen atoms.
Figure 9 shows that the decrease in energy of the apical oxygen mode is due to the increase in the distance of the Oap from the CuO2 planes. Apparently, the extra carriers introduced in the CuO2 planes from the Ca substitution induce redistribution of charges between CuO2 planes and CuO chains repelling the apical oxygen from CuO2 planes. Although the data are limited, it seems that there is a stepwise increase in the CuO2-Oap distance with Ca overdoping. Since the accurately measured amount of oxygen does not vary appreciably with Ca substitution (Figure 10) and the width of the apex mode is almost the same as for the ortho-I phase (Figure 2), this effect could not be due to coexisting phases from chain ordering. Possibly, it indicates the tendency of Oap to stay at discrete positions along the c-axis as in a double well potential, or equivalently discrete charge transfers from chains to the superconducting planes. Such an important effect requires further investigation using several Ca concentrations and fully oxygenated samples.
The in-phase vibrations of Opl atoms are strongly dampened by oxygen doping, with their energy reducing and the width increasing with x in pure Y123x compound (Figure 2b). The overdoping with Ca further increases the width and to a lesser degree decreases the energy (Figure 2b), although the presence of the phonon softening at oxygen overdoping complicates the effect. The relative amount of the softened phase located at 428 cm−1 in the Ca substituted compounds to the total counts of in-phase mode, correlates very well with the amount of oxygen and not with the concentration in Ca (Figure 10). The EXAFS measurements have indicated that the carriers introduced by Ca do not reside at the CuO2 (or the BaO) planes, but the Ca atoms act as pinning centers for these excess carriers [136]. In such a case, the continuous reduction in Tc with the amount of Ca could be a result of percolation from the separation into at least two phases (Figure 8), a pure YBa2Cu3O7−δ and a mixed one not superconducting. In the studies of the R11−xR2xBa2Cu3O7−δ systems, we have observed that the B1g-mode splits into two or three modes that correspond to two end compounds and a mixed phase. In the Ca case, in addition to the B1g-symmetry mode of YBa2Cu3O7−δ which is gradually reduced in intensity with the amount of Ca, another mode of Ag-symmetry at ≈322 cm−1 appears to gain intensity with the substitution, but its energy remains unaffected by the amount of Ca. This extra mode must correspond to the mixed phase. On the basis of the B1g phonon energy dependence on the ion size [128,129,132], we can deduce that it could roughly correspond to a Y0.5Ca0.5Ba3O7−δ phase.
Furthermore, there is a very weak mode at ≈312 cm−1 activated by the Ca substitution (Figure 8). In the pure YBa2Cu3O7 system, there is a mode of B2g,3g symmetry due to vibrations of the yttrium ions. The substitution of heavier yttrium atoms by Ca should induce a shift of ≈5% in frequency as compared with the 10% difference observed. It is, therefore, unlikely that the two modes are related and most likely the ≈312 cm−1 mode corresponds to an ideal CaBa2Cu3Ox phase, which is unstable as a free compound.

2.3. Pressure Induced Lattice Effects

As detected by the softening of the in-phase mode (Figure 2b) and confirmed by other measurements [125,126], YΒ2Cu3O7−δ at optimal doping with maximum Tc is at the edge of a lattice instability undergoing a small lattice distortion. Systematic Raman studies on various high Tc cuprates have revealed nonlinear lattice distortions and lattice instabilities induced by the application of hydrostatic pressure [137,138,139,140,141], which correlate with the non-monotonic pressure dependence of Tc. This is not surprising because it is well accepted that structural and electronic inhomogeneities constitute intrinsic properties of cuprate superconductors [10,142,143]. The Raman study of hydrostatically compressed YBa2Cu3Oy (Y123x) and YBa2Cu4O8 (Y1248) cuprates revealed deviations from an expected linear behavior in the phonon characteristics indicating lattice instabilities [139,144] at those pressures where Tc dependence shows saturation or nonlinear behavior [145,146]. Synchrotron angle-dispersive powder diffraction measurements with dense sampling on optimally doped Y1237 and Y1248 [147] superconductors verified the Raman results showing clear deviation of the lattice constants from the expected equation of states (EOS) followed by a strong hysteresis effect in the same pressure range, which disappear in the non-superconducting PrBa2Cu3O7 (Pr1237) [147].
Figure 11 presents the comparative results on the hydrostatic pressure dependence of the lattice parameters a, b, and c for the three compounds (superconducting Y1237, Y1248, and non-superconducting Pr1237). At the bottom of the figure, the corresponding relative strain is depicted for the three cases. Full symbols correspond to the measurements to increase the pressure and open symbols for those obtained on pressure release. It was found, as expected, that the compressibility along the a-, b-, and c-axis is anisotropic, with the c-axis exhibiting the strongest reduction upon pressure [147]. The main result is the strong deviation of the pressure dependence of the c-axis that starts around 3.7 GPa and extends to 8 to 10 GPa (Figure 11). The deviation also exists in the same pressure range for the other two axes, but it is much smaller. It is interesting that there is a substantial hysteresis effect with the data following the decompression the expected EOS. The nonlinear evolution of the lattice constants, with the applied pressure, originates from similar modifications in the interatomic distances such as the Ba distance from the basal plane, the Cu2-Opl bond length, and the Cu2-Cu1 distance along the c-axis [147]. Concerning the relative strain calculated from the (113) line, it shows three pressure regions (Figure 1, bottom). Up to ~3 GPa, the data among the three compounds agree following the expected slight increase in the line width and, in the region ~3 to 8 GPa, there is a change in the slope only in the two superconducting compounds (Y1237 and Y1248) to be followed by a further increase in slope at higher pressures, which is common in all three cases. It seems that at the critical pressure range of ~3–8 GPa, something happens to the lattice which affects the carriers related to superconductivity. Above ~8 GPa, there must be another effect which is common to all compounds.
Figure 12 presents the comparative results of Raman measurements among superconducting Y1237 and non-superconducting Pr1237 for the strong Ag-symmetry phonons that involve the Oap, Oin-ph, Ba, and Cu2 atoms. The clear anomaly and the plateau around 3 GPa observed in Y1237 is absent for Pr1237 and the results agree with the X-ray diffraction measurements (Figure 11) and the modifications in the interatomic distances [147]. Raman data have shown that, even for the ortho-II phase, a similar anomaly exists in the Ag-symmetry phonon modes [147]. Although no hydrostatic Raman pressure data exist for the Y1236 compound, the absence of this anomaly for Pr1237 supports the hypothesis that this effect is related with the carriers.
The c-axis modification with pressure is a result of interatomic bond length changes resulting from pressure-induced charge redistribution or internal strains in the structure. Charge redistribution among the planes has been found to occur in the case of chemical pressure [148] affecting Tc and must certainly play some role in the observed effects. The correlation of the structural characteristics with Raman spectral modifications of the in-phase Opl and Oap phonon modes (Figure 12) and corresponding changes of Tc [145,146] imply that the trigger of the lattice instabilities lies among the CuO2 and BaO planes. The average Cu2-Opl bond distance is almost constant in the low pressure region, up to ~1.5 GPa, it decreases for increasing pressure up to 3.7 GPa, remains almost constant in the range of 3.7 GPa < p < 8 GPa, and decreases further for p > 8 GPa [149]. The fractional coordinate z of the Ba atom along the c-axis also shows similar modifications at the same pressures (Figure 13, bottom). In the low-pressure region, up to ~1.5 GPa, the fractional coordinate z of Ba decreases almost linearly with pressure. With increasing pressure, z increases and levels off for 3.7 GPa < p < 8 GPa, indicating that the Ba atom moves away from the basal plane. For p > 8 GPa, z is decreasing indicating that Ba atom is again pushed towards the basal plane. Upon pressure release, the movement of the Ba atom clearly shows a hysteresis effect (Figure 13). The Cu2-Cu1 distance in the perovskite block hosting the Ba atom shows similar deviations from linear behavior at the same pressures (3.7 GPa and ~8 GPa, top of Figure 13). Since this distance is related to the transfer of carriers from the CuO chains to the CuO2 superconducting planes, its abnormal modification with pressure is certainly connected with an irregular transfer of carriers to the CuO2 planes. Therefore, the data point to lattice instability at a critical pressure, which affects the carrier distribution and Tc.
The new phase and the lattice anomalies could not be related with a non-hydrostatic environment of the transmitting medium. Although the line width of the luminescence line of ruby and the XRD line width cannot provide a definite proof of a hydrostatic medium, no broadening was observed before the 12.7 GPa pressure. Furthermore, the two-dimensional (2D) patterns in the synchrotron detector do not bear any sign of a non-hydrostatic pressure for the lines of the main phase. This is in complete agreement with all measurements performed up to now that do not provide any evidence for a non-hydrostatic environment in the methanol-ethanol mixture for pressures up to 10 GPa [150]. It was also not related to any oxygen deficiency, as the apex mode of the Raman spectra indicate fully oxygenated compounds (Figure 4). In order to investigate the effect of doping and the transmitting medium, the La2−xSrxCuO4 system has been examined at hydrostatic pressures by Raman spectroscopy for selected doping levels using the same methanol-ethanol transmitting medium and synchrotron XRD with two different media, i.e., a mixture of methanol-ethanol and neon [151].
Concerning the effect of the transmitting medium, it was found that Tc gradually increases when Bi2Sr2CaCu2O8 is exposed to methanol at room temperature, reaching the maximum Tc after a month [152]. The whole process could be accelerated substantially under hydrostatic pressures. But in the same work no change in Tc was detected for the La2−xSrxCuO4 and YBa2Cu3O7 compounds. In another systematic investigation of the effect of methanol on La2−xSrxCuO4, it was discovered that the adsorption of methanol depends on the amount of Sr via the smaller Cu+3 ions in the compound, which facilitate intercalation and the whole effect peaks around x = 0.2 doping [153]. Again, hydrostatic pressure should accelerate and possibly modify the process. Synchrotron XRD data collected on the same La2CuO4 compound with a mixture of methanol-ethanol and neon have really detected differences that begin around 7 to 8 GPa [151]. This pressure region almost coincides with the anomalies observed in both Y1237 and Pr1237 (Figure 11), strongly suggesting that the origin of this effect is related to the presence of methanol. For some reason in Y1248 no such effect has been observed (Figure 11) and this requires further high-pressure investigation using another transmitting medium. In any case, the effect observed both Y1237 and Y1248 at lower pressures (~3 to 4 GPa, Figure 11) must be related to an intrinsic instability of the lattice in these compounds, independent of the transmitting medium.
Concerning the La2−xSrxCuO4 compound, synchrotron XRD and Raman data indicate modifications with pressure that depend on the amount of Sr doping [151]. The top section of Figure 14 presents the results for the c-axis of La1.85Sr0.15CuO4 and pure La2CuO4 systems. For the x = 0.15 case, the XRD measurements show a deviation around 3 GPa, which disappears for the case of pure La2CuO4. As in the cases of yttrium systems, it also shows a hysteresis effect (Figure 14). The bottom portion of Figure 14 presents the pressure dependence of the La/Sr mode for the case of x = 0.15. Something similar, but a smaller effect, happens for the apex phonon [151]. It is clear that Raman data indicate a deviation from linear pressure dependence at ~3 GPa, i.e., the same pressure where the c-axis deviates from normal EOS. Preliminary high-pressure Raman measurements show that the effect depends strongly on the amount of doping [151]. Similar conclusions are obtained from the FWHM of the 004, 103, and 110 diffraction lines [151].
The effect originates from an anomalous pressure dependence of the apical oxygen position along the c-axis. The La/Sr-O(2) and Cu-O(2) bond length modifications along the c-axis are indications of charge redistribution with hydrostatic pressure [154]. It is also interesting that the tolerance factor, t, calculated from the average distance of the La-O bond lengths, as compared with those of the Cu-O ones [t = dLaO/ 2 dCuO] and associated with charge redistribution [155], also shows an anomaly at ~2 GPa for the optimally doped compound. Furthermore, the tolerance factor appears to jump among different almost equidistant values discontinuously [151]. Although the statistics is not good enough to verify this effect and draw final conclusions, this could be related to the abrupt changes observed for the Cupl-Oap in the system Y1−yCayBa2Cu3O7 (Figure 9), and the plateaus observed in the Raman measurements under hydrostatic pressure (Figure 12). Such effects could be described as a signature of a stepwise modification of the carrier concentration in the superconducting planes, as in topologically distinct states.

3. Conclusions

Raman studies on a series of cuprates were presented showing lattice distortions induced either by chemical doping or by applying hydrostatic pressure. From the spectral modifications (frequency and width) of the apex and in-phase phonons of the yttrium compounds it was found that the oxygen non-stoichiometry is related with the ordering of the chain oxygen atoms creating separation into submicron phases with varying transition temperature. Our Raman and more recent measurements [122,123,124] prove that these phases coexist in the YBa2Cu3Ox compounds and only the relative amount of the phases varies with doping. Just above the optimum doping, another lattice effect was discovered from the softening of the in-phase mode, which is related to a slight reduction of the transition temperature and originates from a sudden change in the CuO2 buckling, as verified by other techniques [125,126].
From the study of the Raman active B1g mode it was found that rare earth substitution for yttrium in the same series of superconductors induces a separation into superconducting and non-superconducting phases. Depending on the relative size and the kind of the substituting rare earth element this affects the ability of the compound to superconduct and the transition temperature. Overdoping by Ca substitution for Y shows a very similar phase separation behavior.
Hydrostatic pressure was found to create small lattice distortions in YBa2Cu3O7 and YBa2Cu4O8 compounds, which are absent for the non-superconducting PrBa2Cu3O7 system. The same happens for the La2−xSrxCuO4 series, which again disappear in the pure non-superconducting La2CuO4 system. Raman studies under hydrostatic pressure in the Bi-cuprates [141] also indicate pressure-induced lattice anomalies at low hydrostatic pressures. In all cases, there are indications for charge redistribution, which apparently affects the carrier concentration in the superconducting planes and modifies Tc.
Raman studies on other high Tc compounds, such as the diborides [60,62,63,102] and the pnictides [103,104,105], have revealed phase separation and local lattice distortions. This is an indication that the observed effects possibly characterize most of the high temperature superconductors. In all cases, (cuprates, pnictides, and diborides) the lattice anomalies are accompanied by modifications in the transition temperature, indicating some connection between the two effects. It is unclear whether the observed effects are related to the pairing mechanism, although the lattice modifications or the phase separation phenomena apparently induce charge redistribution between the superconducting and the insulating planes in the studied layered structured superconductors. They certainly show that the lattice degrees of freedom participate in the related phenomena to an extent that needs to be clarified.

Funding

This research received no external funding.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Yukalov, V.I. Heterophase fluctuations in ferroelectrics. Ferroelectrics 1988, 82, 11–24. [Google Scholar] [CrossRef]
  2. Yukalov, V.I. Phase transitions and heterophase fluctuations. Phys. Rep. 1991, 208, 395–489. [Google Scholar] [CrossRef]
  3. Nagaev, E.L. Phase-separation mechanism for giant magnetoresistance of lanthanum manganites. Phys. Lett. A 1996, 218, 367–372. [Google Scholar] [CrossRef]
  4. Lanzara, A.; Saini, N.L.; Brunelli, M.; Natali, F.; Bianconi, A.; Radaelli, P.G.; Cheong, S.W. Crossover from large to small polarons across the metal-insulator transition in manganites. Phys. Rev. Lett. 1998, 81, 878–881. [Google Scholar] [CrossRef]
  5. Yunoki, S.; Hu, J.; Malvezzi, A.L.; Moreo, A.; Furukawa, N.; Dagotto, E. Phase separation in electronic models for manganites. Phys. Rev. Lett. 1998, 80, 845–848. [Google Scholar] [CrossRef]
  6. Uehara, M.; Mori, S.; Chen, C.H.; Cheong, S.W. Percolative phase separation underlies colossal magnetoresistance in mixed-valent manganites. Nature 1999, 399, 560–563. [Google Scholar] [CrossRef]
  7. Kagan, M.Y.; Kugel, K.I. Inhomogeneous charge distributions and phase separation in manganites. Phys. Uspekhi 2001, 44, 553–570. [Google Scholar] [CrossRef]
  8. Kagan, M.Y.; Klaptsov, A.V.; Brodsky, I.V.; Kugel, K.I.; Sboychakov, A.O.; Rakhmanov, A.L. Nanoscale phase separation in manganites. J. Phys. A Math. Gen. 2003, 36, 9155–9164. [Google Scholar] [CrossRef]
  9. Dagotto, E.; Burgy, J.; Moreo, A. Nanoscale phase separation in colossal magnetoresistance materials: Lessons for the cuprates? Solid State Commum. 2003, 126, 9–22. [Google Scholar] [CrossRef]
  10. Dagotto, E. Complexity in strongly correlated electronic systems. Science 2005, 309, 257–262. [Google Scholar] [CrossRef]
  11. Tokura, Y. Critical features of colossal magnetoresistive manganites. Rep. Prog. Phys. 2006, 69, 797–851. [Google Scholar] [CrossRef]
  12. Hayden, S.M.; Lander, G.H.; Zaretsky, J.; Brown, P.J.; Stassis, C.; Metcalf, P.; Honig, J.M. Incommensurate magnetic correlations in La1.8Sr0.2NiO4. Phys. Rev. Lett. 1992, 68, 1061. [Google Scholar] [CrossRef] [PubMed]
  13. Campi, G.; Poccia, N.; Joseph, B.; Bianconi, A.; Mishra, S.; Lee, J.; Roy, S.; Nugroho, A.A.; Buchholz, M.; Braden, M.; et al. Direct Visualization of Spatial Inhomogeneity of Spin Stripes Order in La1.72Sr0.28NiO4. Condens. Matter 2019, 4, 77. [Google Scholar] [CrossRef]
  14. Testardi, L.R. Structural instability and superconductivity in A-15 compounds. Rev. Mod. Phys. 1975, 47, 637–648. [Google Scholar] [CrossRef]
  15. Bednorz, J.G.; Müller, K.A. Possible high Tc superconductivity in the Ba-La-Cu-O system. Z. Phys. B Condens. Matter 1986, 64, 189–193. [Google Scholar] [CrossRef]
  16. Jorgensen, A.J.; Dabrowski, B.; Pei, S.; Hinks, D.G.; Soderholm, L.; Morosin, B.; Schirber, E.L.; Venturini, E.L.; Ginley, D.S. Superconducting phase of La2CuO4+δ: A superconducting composition resulting from phase separation. Phys. Rev. B 1988, 38, 11337–11345. [Google Scholar] [CrossRef]
  17. Benedek, G.; Müller, K.A. Phase Separation in Cuprate Superconductors; World Scientific: Singapore, 1992. [Google Scholar]
  18. Bianconi, A. Phase Separation in Cuprate Superconductors; Müller, K.A., Benedek, G., Eds.; World Scientific: Singapore, 1993; pp. 125–138. [Google Scholar]
  19. Kremer, R.K.; Sigmund, E.; Hizhnyakov, V.; Hentsch, F.; Simon, A.; Müller, K.A.; Mehring, M. Percolative phase separation in La2CuO4+δ and La2−xSrxCuO4. Z. Phys. B Condens. Matter 1992, 86, 319–324. [Google Scholar] [CrossRef]
  20. Benedek, G.; Müller, K.A. Phase Separation in Cuprate Superconductors; Springer: Berlin/Herdelberg, Germany, 1994. [Google Scholar]
  21. Bianconi, A.; Missori, M. The Coupling of a Wigner Polaronic Charge Density Wave with a Fermi Liquid Arising from the Instability of a Wigner Polaron Crystal: A Possible Pairing Mechanism in High Tc Superconductors. In Phase Separation in Cuprate Superconductors; Springer: Berlin/Heidelberg, Germany, 1994; pp. 272–289. [Google Scholar]
  22. Bianconi, A. On the Fermi liquid coupled with a generalized Wigner polaronic CDW giving high Tc superconductivity. Solid State Commun. 1994, 91, 1–5. [Google Scholar] [CrossRef]
  23. Bianconi, A.; Missori, M. High Tc superconductivity by quantum confinement. J. Phys. 1994, 4, 361–365. [Google Scholar] [Green Version]
  24. Bianconi, A. On the possibility of new high Tc superconductors by producing metal heterostructures as in the cuprate perovskites. Solid State Commum. 1994, 89, 933–936. [Google Scholar] [CrossRef] [Green Version]
  25. Hizhnyakov, V.; Sigmund, E.; Seibold, G. Polaron formation and percolative phase separation in HTSC. In Phase Separation in Cuprate Superconductors; Springer: Berlin/Heidelberg, Germany, 1994; pp. 50–65. [Google Scholar]
  26. Bianconi, A.; Missori, M.; Oyanagi, H.; Yamaguchi, H.; Ha, D.H.; Nishiara, Y.; Della Longa, S. The measurement of the polaron size in the metallic phase of cuprate superconductors. Eur. Lett. 1995, 31, 411. [Google Scholar] [CrossRef]
  27. Nagaev, E.L. Phase separation in high-temperature superconductors and related magnetic materials. Uspekhi Fizicheskikh Nauk 1995, 165, 529–555. [Google Scholar] [CrossRef]
  28. Zech, D.; Conder, K.; Keller, H.; Kaldis, E.; Liarokapis, E.; Poulakis, N.; Müller, K.A. Anharmonic Properties of High-Tc Cuprates; Mihailovic, D., Ruani, G., Kaldis, E., Müller, K.A., Eds.; World Scientific: Singapore, 1995; pp. 18–29. [Google Scholar]
  29. Kusmartsev, F.V.; Di Castro, D.; Bianconi, G.; Bianconi, A. Transformation of strings into an inhomogeneous phase of stripes and itinerant carriers. Phys. Lett. A 2000, 275, 118–123. [Google Scholar] [CrossRef]
  30. Gor’kov, L.P. Phase separation in a two-component model for cuprates. J. Supercond. 2000, 13, 765–769. [Google Scholar] [CrossRef]
  31. Bianconi, A.; Di Castro, D.; Bianconi, G.; Pifferi, A.; Saini, N.L.; Chou, F.C.; Johnston, D.C.; Colapietro, M. Coexistence of stripes and superconductivity: Tc amplification in a superlattice of superconducting stripes. Phys. C 2000, 341, 1719–1722. [Google Scholar] [CrossRef]
  32. Coleman, A.J. Phase Transitions and Self-Organization in Electronic and Molecular Networks; Phillips, J.C., Thorpe, M.F., Eds.; Kluwer: New York, NY, USA, 2001; pp. 23–35. [Google Scholar]
  33. Phillips, J.C. Percolative model of nanoscale phase separation in high-temperature superconductors. Philos. Mag. B 2002, 82, 783–790. [Google Scholar] [CrossRef]
  34. Bishop, A.R.; Lookman, T.; Saxena, A.; Shenoy, S.R. Elasticity-driven nanoscale texturing in complex electronic materials. Eur. Phys. Lett. 2003, 63, 289–295. [Google Scholar] [CrossRef]
  35. Phillips, J.C.; Saxena, A.; Bishop, A.R. Pseudogaps, dopants, and strong disorder in cuprate high-temperature superconductors. Rep. Prog. Phys. 2003, 66, 2111–2182. [Google Scholar] [CrossRef]
  36. De Mello, E.V.L.; Caixeiro, E.S. Effects of phase separation in the cuprate superconductors. Phys. Rev. B 2004, 70, 224517. [Google Scholar] [CrossRef] [Green Version]
  37. Kugel, K.I.; Rakhmanov, A.L.; Sboychakov, A.O. Phase separation in Jahn-Teller systems with localized and itinerant electrons. Phys. Rev. Lett. 2005, 95, 267210. [Google Scholar] [CrossRef]
  38. De Mello, E.V.L.; Dias, D.H.N. Phase separation and the phase diagram of cuprate superconductors. J. Phys. Condens. Matter 2007, 19, 086218. [Google Scholar] [CrossRef]
  39. Fine, B.V.; Egami, T. Phase separation in the vicinity of quantum-critical doping concentration: Implications for high-temperature superconductors. Phys. Rev. B 2008, 77, 014519. [Google Scholar] [CrossRef] [Green Version]
  40. Kugel, K.I.; Rakhmanov, A.L.; Sboychakov, A.O.; Poccia, N.; Bianconi, A. Model for phase separation controlled by doping and the internal chemical pressure in different cuprate superconductors. Phys. Rev. B 2008, 78, 165124. [Google Scholar] [CrossRef] [Green Version]
  41. Fratini, M.; Poccia, N.; Bianconi, A. The Feshbach resonance and nanoscale phase separation in a polaron liquid near the quantum critical point for a polaron Wigner crystal. J. Phys. 2008, 108, 012036. [Google Scholar] [CrossRef]
  42. Innocenti, D.; Ricci, A.; Poccia, N.; Campi, G.; Fratini, M.; Bianconi, A. A model for liquid-striped liquid phase separation in liquids of anisotropic polarons. J. Supercond. Nov. Magn. 2009, 22, 529–533. [Google Scholar] [CrossRef]
  43. De Mello, E.V.L.; Kasal, R.B.; Passos, C.A.C. Electronic phase separation transition as the origin of the superconductivity and pseudogap phase of cuprates. J. Phys. Condens. Matter 2009, 21, 235701. [Google Scholar] [CrossRef] [Green Version]
  44. Fratini, M.; Poccia, N.; Ricci, A.; Campi, G.; Burghammer, M.; Aeppli, G.; Bianconi, A. Scale-free structural organization of oxygen interstitials in La2CuO4+y. Nature 2010, 466, 841–844. [Google Scholar] [CrossRef]
  45. Poccia, N.; Fratini, M.; Ricci, A.; Campi, G.; Barba, L.; Vittorini-Orgeas, A.; Bianconi, G.; Aeppli, G.; Bianconi, A. Evolution and control of oxygen order in a cuprate superconductor. Nat. Mater. 2011, 10, 733–736. [Google Scholar] [CrossRef]
  46. Poccia, N.; Ricci, A.; Campi, G.; Fratini, M.; Puri, A.; Di Gioacchino, D.; Marcelli, A.; Reynolds, M.; Burghammer, M.; Saini, N.L.; et al. Optimum inhomogeneity of local lattice distortions in La2CuO4+y. Proc. Natl. Acad. Sic. USA 2012, 109, 15685–15690. [Google Scholar] [CrossRef]
  47. Pinheiro, C.F.S.; De Mello, E.V.L. Random resistivity network calculations for cuprate superconductors with an electronic phase separation transition. Physica A 2012, 391, 1532–1539. [Google Scholar] [CrossRef]
  48. De Mello, E.V.L. Describing how the superconducting transition in La2CuO4+y is related to the iO phase separation. J. Supercond. Nov. Magn. 2012, 25, 1347–1350. [Google Scholar] [CrossRef]
  49. De Mello, E.V.L. Description and connection between the oxygen order evolution and the superconducting transition in La2CuO4+y. Eur. Phys. Lett. 2012, 98, 57008. [Google Scholar] [CrossRef]
  50. Phillips, J.C. Ineluctable complexity of high temperature superconductivity elucidated. J. Supercond. Nov. Magn. 2014, 27, 345–347. [Google Scholar] [CrossRef]
  51. Bianconi, A.; Poccia, N.; Sboychakov, A.O.; Rakhmanov, A.L.; Kugel, K.I. Intrinsic arrested nanoscale phase separation near a topological Lifshitz transition in strongly correlated two-band metals. Supercond. Sci. Technol. 2015, 28, 024005. [Google Scholar] [CrossRef] [Green Version]
  52. Campi, G.; Bianconi, A.; Poccia, N.; Bianconi, G.; Barba, L.; Arrighetti, G.; Innocenti, D.; Karpinski, J.; Zhigadlo, N.D.; Kazakov, S.M.; et al. Inhomogeneity of charge-density-wave order and quenched disorder in a high-Tc superconductor. Nature 2015, 525, 359–362. [Google Scholar] [CrossRef]
  53. Agrestini, S.; Saini, N.L.; Bianconi, G.; Bianconi, A. The strain of CuO2 lattice: The second variable for the phase diagram of cuprate perovskites. J. Phys. A Math. Gen. 2003, 36, 9133–9142. [Google Scholar] [CrossRef]
  54. Campi, G.; Ricci, A.; Poccia, N.; Fratini, M.; Bianconi, A. X-rays Writing/Reading of charge density waves in the CuO2 plane of a simple cuprate superconductor. Condens. Matter 2017, 2, 26. [Google Scholar] [CrossRef]
  55. Campi, G.; Bianconi, A. Evolution of complexity in out-of-equilibrium systems by time-resolved or space-resolved synchrotron radiation techniques. Condens. Matter 2019, 4, 32. [Google Scholar] [CrossRef]
  56. Phillips, J.C. Physics of High-Tc Superconductors; Academic: Boston, MA, USA, 1989. [Google Scholar]
  57. Nagamatsu, J.; Nakagawa, N.; Muranaka, T.; Zenitani, Y.; Akimitsu, J. Superconductivity at 39 K in magnesium diboride. Nature 2001, 410, 63–64. [Google Scholar] [CrossRef]
  58. Bauer, E.; Paul, C.; Berger, S.; Majumdar, S.; Michor, H.; Giovannini, M.; Saccone, A.; Bianconi, A. Thermal conductivity of superconducting MgB2. J. Phys. Condens. Mat. 2001, 13, L487–L493. [Google Scholar] [CrossRef]
  59. Agrestini, S.; Di Castro, D.; Sansone, M.; Saini, N.L.; Saccone, A.; De Negri, S.; Giovannini, M.; Colapietro, M.; Bianconi, A. High Tc superconductivity in a critical range of micro-strain and charge density in diborides. J. Phys. Condens. Matter 2001, 13, 11689–11695. [Google Scholar] [CrossRef]
  60. Agrestini, S.; Metallo, C.; Filippi, M.; Siminelli, L.; Campi, G.; Sanipoli, C.; Liarokapis, E.; De Negri, S.; Giovannini, M.; Saccone, A.; et al. Substitution of Sc for Mg in MgB2: Effects on transition temperature and Kohn anomaly. Phys. Rev. B 2004, 70, 134514. [Google Scholar] [CrossRef]
  61. Palmisano, V.; Simonelli, L.; Puri, A.; Fratini, M.; Busby, Y.; Parisiades, P.; Liarokapis, E.; Brunelli, M.; Fitch, A.N.; Bianconi, A. Controlling mesoscopic phase separation near electronic topological transitions via quenched disorder in ternary diborides. J. Phys. Condens. Matter 2008, 20, 434222. [Google Scholar] [CrossRef]
  62. Simonelli, L.; Palmisano, V.; Fratini, M.; Paridiades, P.; Lampakis, D.; Liarokapis, E.; Bianconi, A. Isotope effect on the E2g phonon and mesoscopic phase separation near the electronic topological transition in Mg1−xAlxB2. Phys. Rev. B 2009, 80, 014520. [Google Scholar] [CrossRef]
  63. Parisiades, P.; Liarokapis, E. Lattice effects in diborides. J. Supercond. Nov. Magn. 2011, 24, 49–56. [Google Scholar] [CrossRef]
  64. Hebard, A.F.; Rosseinsky, M.J.; Haddon, R.C.; Murphy, D.W.; Glarum, S.H.; Palstra, T.T.M.; Ramirez, A.P.; Kortan, A.R. Superconductivity at 18 K in potassium doped C60. Nature 1991, 350, 600–601. [Google Scholar] [CrossRef]
  65. Zhou, O.; Zhu, Q.; Fischer, J.E.; Coustel, N.; Vaughan, G.B.M.; Heiney, P.A.; McCauley, J.P., Jr.; Smith, A.B. Compressibility of M3C60 fullerene superconductors: Relation between Tc and lattice parameter. Science 1992, 255, 833–835. [Google Scholar] [CrossRef]
  66. Palstra, T.T.M.; Zhou, O.; Iwasa, Y.; Sulewski, P.E.; Fleming, R.M.; Zegarski, B.R. Superconductivity at 40K in cesium doped C60. Solid State Commum. 1995, 93, 327–330. [Google Scholar] [CrossRef]
  67. Drozdov, A.P.; Eremets, M.I.; Troyan, I.A.; Ksenofontov, V.; Shylin, S.I. Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system. Nature 2015, 525, 73–76. [Google Scholar] [CrossRef]
  68. Bianconi, A.; Jarlborg, T. Superconductivity above the lowest Earth temperature in pressurized sulfur hydride. Eur. Phys. Lett. 2015, 112, 37001. [Google Scholar] [CrossRef] [Green Version]
  69. Jarlborg, T.; Bianconi, A. Breakdown of the Migdal approximation at Lifshitz transitions with giant zero-point motion in the H3S superconductor. Sci. Rep. 2016, 6, 24816. [Google Scholar] [CrossRef] [Green Version]
  70. Bianco, R.; Errea, I.; Calandra, M.; Mauri, F. High-pressure phase diagram of hydrogen and deuterium sulfides from first principles: Structural and vibrational properties including quantum and anharmonic effects. Phys. Rev. B 2018, 97, 214101. [Google Scholar] [CrossRef] [Green Version]
  71. Bianconi, A.; Saini, N.L.; Lanzara, A.; Missori, M.; Rossetti, T.; Oyanagi, H.; Yamaguchi, H.; Oka, K.; Ito, T. Determination of the lattice Distortions in the CuO2 Plane of La1.85Sr0.15CuO4. Phys. Rev. Lett. 1996, 76, 3412–3415. [Google Scholar] [CrossRef] [PubMed]
  72. Lanzara, A.; Zhao, G.M.; Saini, N.L.; Bianconi, A.; Conder, K.; Keller, H.; Müller, K.A. Oxygen-isotope shift of the charge-stripe ordering temperature in La2−xSrxCuO4 from x-ray absorption spectroscopy. J. Phys. Condens. Matter 1999, 11, L541–L546. [Google Scholar] [CrossRef]
  73. Lanzara, A.; Bogdanov, P.V.; Zhou, X.J.; Kellar, S.A.; Feng, D.L.; Lu, E.D.; Yoshida, T.; Eisaki, H.; Fujimori, A.; Kishio, K.; et al. Evidence for ubiquitous strong electron-phonon coupling in high-temperature superconductors. Nature 2001, 412, 510–514. [Google Scholar] [CrossRef] [PubMed]
  74. Shen, Z.X.; Lanzara, A.; Ishihara, S.; Nagaosa, N. Role of the electron-phonon interaction in the strongly correlated cuprate superconductors. Philos. Mag. B 2002, 82, 1349–1368. [Google Scholar] [CrossRef]
  75. Gweon, G.H.; Sasagawa, T.; Zhou, S.Y.; Graf, J.; Takagi, H.; Lee, D.H.; Lanzara, A. An unusual isotope effect in a high-transition-temperature superconductor. Nature 2004, 430, 187–190. [Google Scholar] [CrossRef] [Green Version]
  76. Graf, J.; Jozwiak, C.; Smallwood, C.L.; Eisaki, H.; Kaindl, R.A.; Lee, D.H.; Lanzara, A. Nodal quasiparticle meltdown in ultrahigh-resolution pump-probe angle-resolved photoemission. Nat. Phys. 2011, 7, 805–809. [Google Scholar] [CrossRef]
  77. Perali, A.; Innocenti, D.; Valletta, A.; Bianconi, A. Anomalous isotope effect near a 2.5 Lifshitz transition in a multi-band multi-condensate superconductor made of a superlattice of stripes. Supercond. Sci. Technol. 2012, 25, 124002. [Google Scholar] [CrossRef]
  78. Guguchia, Z.; Khasanov, R.; Bendele, M.; Pomjakushina, E.; Conder, K.; Shengelaya, A.; Keller, H. Negative Oxygen Isotope Effect on the Static Spin Stripe Order in Superconducting La2−xBaxCuO4 (x = 1/8) Observed by Muon-Spin Rotation. Phys. Rev. Lett. 2014, 113, 057002. [Google Scholar] [CrossRef]
  79. Bendele, M.; von Rohr, F.; Guguchia, Z.; Pomjakushina, E.; Conder, K.; Bianconi, A.; Simon, A.; Bussmann-Holder, A.; Keller, H. Evidence for strong lattice effects as revealed from huge unconventional oxygen isotope effects on the pseudogap temperature in La2−xSrxCuO4. Phys. Rev. B 2017, 95, 014514. [Google Scholar] [CrossRef]
  80. Caivano, R.; Fratini, M.; Poccia, N.; Ricci, A.; Puri, A.; Ren, Z.A.; Dong, X.L.; Yang, J.; Lu, W.; Zhao, Z.X.; et al. Feshbach resonance and mesoscopic phase separation near a quantum critical point in multiband FeAs-based superconductors. Supercond. Sci. Technol. 2008, 22, 014004. [Google Scholar] [CrossRef]
  81. Ricci, A.; Poccia, N.; Campi, G.; Joseph, B.; Arrighetti, G.; Barba, L.; Reynolds, M.; Burghammer, M.; Takeya, H.; Mizuguchi, Y.; et al. Nanoscale phase separation in the iron chalcogenide superconductor K0.8Fe1.6Se2 as seen via scanning nanofocused x-ray diffraction. Phys. Rev. B 2011, 84, 060511. [Google Scholar] [CrossRef]
  82. Bianconi, A. Quantum materials: Shape resonances in superstripes. Nat. Phys. 2013, 9, 536–537. [Google Scholar] [CrossRef]
  83. Bendele, M.; Barinov, A.; Joseph, B.; Innocenti, D.; Iadecola, A.; Bianconi, A.; Takeya, H.; Mizuguchi, Y.; Takano, T.; Noji, T.; et al. Spectromicroscopy of electronic phase separation in KxFe2ySe2 superconductor. Sci. Rep. 2014, 4, 5592. [Google Scholar] [CrossRef]
  84. Ricci, A.; Poccia, N.; Joseph, B.; Innocenti, D.; Campi, G.; Zozulya, A.; Westermeier, A.; Schavkan, A.; Coneri, F.; Bianconi, A.; et al. Direct observation of nanoscale interface phase in the superconducting chalcogenide KxFe2−ySe2 with intrinsic phase separation. Phys. Rev. B 2015, 91, 020503. [Google Scholar] [CrossRef]
  85. Simonelli, L.; Mizokawa, T.; Sala, M.M.; Takeya, H.; Mizuguchi, Y.; Takano, Y.; Garbarino, G.; Monaco, G.; Saini, N.L. Temperature dependence of iron local magnetic moment in phase-separated superconducting chalcogenide. Phys. Rev. B 2014, 90, 214516. [Google Scholar] [CrossRef]
  86. Mangelis, P.; Lei, H.; McDonnell, M.; Feygenson, M.; Petrovic, C.; Bozin, E.; Lappas, A. On the Nanoscale Structure of KxFe2−yCh2 (Ch=S, Se): A Neutron Pair Distribution Function View. Condens. Matter 2018, 3, 20. [Google Scholar] [CrossRef]
  87. Bianconi, A. Multiplet splitting of final-state configurations in x-ray-absorption spectrum of metal VO2: Effect of core-hole-screening, electron correlation, and metal-insulator transition. Phys. Rev. B 1982, 26, 2741. [Google Scholar] [CrossRef]
  88. Marcelli, A.; Coreno, M.; Stredansky, M.; Xu, W.; Zou, C.; Fan, L.; Wangsheng, C.; Wei, S.; Cossaro, A.; Ricci, A.; et al. Nanoscale phase separation and lattice complexity in VO2: The metal-insulator transition investigated by XANES via Auger electron yield at the vanadium L23-edge and resonant photoemission. Condens. Matter 2017, 2, 38. [Google Scholar] [CrossRef]
  89. Gioacchino, D.; Marcelli, A.; Puri, A.; Zou, C.; Fan, L.; Zeitler, U.; Bianconi, A. Metastability phenomena in VO2 thin films. Condens. Matter 2017, 2, 10. [Google Scholar] [CrossRef]
  90. Corder, S.N.G.; Jiang, J.; Chen, X.; Kittiwatanakul, S.; Tung, I.C.; Zhu, Y.; Zhang, J.; Bechtel, H.; Martin, M.C.; Carr, L.G.; et al. Controlling phase separation in vanadium dioxide thin films via substrate engineering. Phys. Rev. B 2017, 96, 161110. [Google Scholar] [CrossRef] [Green Version]
  91. Vidas, L.; Günther, C.M.; Miller, T.A.; Pfau, B.; Perez-Salinas, D.; Martínez, E.; Schneider, M.; Guhrs, E.; Gargiani, P.; Valvidares, M.; et al. Imaging Nanometer Phase Coexistence at Defects During the Insulator-Metal Phase Transformation in VO2 Thin Films by Resonant Soft X-ray Holography. Nano Lett. 2018, 18, 3449–3453. [Google Scholar] [CrossRef] [PubMed]
  92. Grandi, F.; Amaricci, A.; Fabrizio, M. Unraveling the Mott-Peierls intrigue in Vanadium dioxide. arXiv 2019, arXiv:1906.10632. [Google Scholar]
  93. Thomsen, C. Light Scattering in Solids; Cardona, M., Guntherodt, G., Eds.; Springer-Verlag: Berlin, Germany, 1991; Volume VI, pp. 285–359. [Google Scholar]
  94. Iliev, M.N.; Thomsen, C.; Hadjiev, V.; Cardona, M. Resonant Raman scattering of oxygen-deficient YBa2Cu3O7−δ: Evidence for the coexistence of ortho-I, ortho-II, and tetragonal microstructures. Phys. Rev. B 1993, 47, 12341–12344. [Google Scholar] [CrossRef]
  95. Poulakis, N.; Palles, D.; Liarokapis, E.; Conder, K.; Kaldis, E.; Müller, K.A. Phase separation and softening of the O2,3 in-phase mode in the YBa2Cu3Ox superconductors. Phys. Rev. B 1996, 53, 219–220. [Google Scholar] [CrossRef]
  96. Mihailovic, D.; McCarty, K.F.; Ginley, D.S. Anharmonic properties and the two-particle continuum in the Raman spectra of YBa2Cu3O6.9, TlBa2CaCu2O7, and Tl2Ba2CaCu2O8. Phys. Rev. B 1993, 47, 8910–8916. [Google Scholar] [CrossRef]
  97. Ruani, G.; Taliani, C.; Muccini, M.; Conder, K.; Kaldis, M.; Keller, H.; Zech, D.; Muller, K.A. Apex anharmonicity observed by Raman scattering in 18O substituted YBa2Cu3O6+x. Physica C 1994, 226, 101–105. [Google Scholar] [CrossRef]
  98. Palles, D.; Poulakis, N.; Liarokapis, E.; Conder, K.; Kaldis, E.; Muller, K.A. Raman study of the oxygen anharmonicity in YBa2Cu3Ox (6.4 < x < 7.0) superconductors. Phys. Rev. B 1996, 54, 6721–6727. [Google Scholar]
  99. Cooper, S.L.; Slakey, F.; Klein, M.V.; Rice, J.P.; Bukowski, E.D.; Ginsberg, D.M. Gap anisotropy and phonon self-energy effects in single-crystal YBa2Cu3O7−δ. Phys. Rev. B 1988, 38, 11934–11937. [Google Scholar] [CrossRef]
  100. Deveraux, T.P.; Einzel, D.; Stadlober, B.; Hackl, R.; Leach, D.H.; Neumeier, J.J. Electronic Raman scattering in high-Tc superconductors: A probe of dx2−y2 pairing. Phys. Rev. Lett. 1994, 72, 396–399. [Google Scholar] [CrossRef] [PubMed]
  101. Heyen, E.T.; Wegerer, R.; Schonherr, E.; Cardona, M. Raman study of the coupling of crystal-field excitations to phonons in NdBa2Cu3O7−δ. Phys. Rev. B 1991, 44, 10195–10205. [Google Scholar] [CrossRef] [PubMed]
  102. Parisiades, P.; Lampakis, D.; Palles, D.; Liarokapis, E.; Zhigadlo, N.D.; Katrysh, S.; Karpinski, J. Two-mode behavior for the E2g broad band in Mg(B1−xCx)2. Phys. C 2008, 468, 1064–1069. [Google Scholar] [CrossRef]
  103. Calamiotou, M.; Margiolaki, I.; Gantis, A.; Siranidi, E.; Ren, Z.A.; Zhao, Z.X.; Liarokapis, E. Lattice anomalies in the FeAs4 tetrahedra of the NdFeAsO0.85 superconductor that disappear at Tc. Eur. Phys. Lett. 2010, 91, 57005. [Google Scholar] [CrossRef]
  104. Liarokapis, E.; Calamiotou, M.; Zhigadlo, N.D.; Katrysh, S.; Karpinski, J. Non-linera lattice response of Sm oxypnictides to hydrostatic pressure. J. Phys. Chem. Solid 2013, 74, 1465–1469. [Google Scholar] [CrossRef]
  105. Calamiotou, M.; Lampakis, D.; Zhigadlo, N.D.; Katrysh, S.; Karpinski, J.; Fitch, A.; Tsiaklaganos, P.; Liarokapis, E. Local lattice distortions vs. structural phase transition in NdFeAsO1−xFx. Phys. C 2016, 527, 55–62. [Google Scholar] [CrossRef]
  106. Thomsen, C.; Cardona, M. Physical Properties of High Temperature Superconductors; Ginsberg, D.M., Ed.; World Scientific: Singapore, 1994; p. 409. [Google Scholar]
  107. Plakida, N.M. High Temperature Superconductivity; Springer-Verlag: Berlin, Germany, 1995. [Google Scholar]
  108. FacFarlane, R.M.; Rosen, H.J.; Engler, E.M.; Jacowitz, R.D.; Lee, V.Y. Raman study of the effect of oxygen stoichiometry on the phonon spectrum of the high-Tc superconductor YBa2Cu3Ox. Phys. Rev. B 1988, 38, 284–289. [Google Scholar] [CrossRef]
  109. Wille, L.T.; Berera, A.; deFontaine, D. Thermodynamics of oxygen ordering in YBa2Cu3Oz. Phys. Rev. Lett. 1988, 60, 1065–1068. [Google Scholar] [CrossRef]
  110. Cava, R.J.; Hewat, A.W.; Hewat, E.A.; Batlogg, B.; Marezio, M.; Rabe, K.M.; Krajewski, J.J.; Peck, W.F.; Rupp, L.W., Jr. Structural anomalies, oxygen ordering and superconductivity in oxygen deficient Ba2YCu3Ox. Physica C 1990, 165, 419–433. [Google Scholar] [CrossRef]
  111. Schleger, P.; Casalta, H.; Hadfield, R.; Poulsen, H.F.; von Zimmermann, M.; Andersen, N.H.; Schneider, J.R.; Liang, R.; Dosanjh, P.; Hardy, W.N. Observation of ortho-III correlations by neutron and hard X-ray scattering in an untwined YBa2Cu3O6.77 single crystal. Physica C 1995, 241, 103–110. [Google Scholar] [CrossRef]
  112. Dagotto, E. Correlated electrons in high-temperature superconductors. Rev. Mod. Phys. 1992, 66, 763–841. [Google Scholar] [CrossRef]
  113. Heyen, E.T.; Kircher, J.; Cardona, M. Resonant Raman scattering in insulating YBa2Cu3O6 as a probe of its electronic structure. Phys. Rev. B 1992, 45, 3037–3047, and references therein. [Google Scholar] [CrossRef] [PubMed]
  114. Denisov, V.N.; Taliani, C.; Mal’shukov, A.G.; Burlakov, V.M.; Schonherr, E.; Ruani, G. Infrared excited Raman scattering and photoluminescence of deep intragap states in semiconducting YBa2Cu3O6+x. Phys. Rev. B 1993, 48, 16714–16721. [Google Scholar] [CrossRef] [PubMed]
  115. Mesot, J.; Allenspach, P.; Staub, U.; Furrer, A.; Mutka, H. Neutron spectroscopic evidence for cluster formation and percolative superconductivity in ErBa2Cu3Ox. Phys. Rev. Lett. 1993, 70, 865–868. [Google Scholar] [CrossRef]
  116. Liarokapis, E. Raman spectroscopy, structural modifications and phase transitions in the high-temperature superconductors. J. Supercond. 2000, 13, 889–893. [Google Scholar] [CrossRef]
  117. Pintschovius, L.; Reichardt, W. Inelastic neutron scattering studies of the lattice vibrations of high Tc compounds. In Physical Properties of High Temperature Superconductors; Ginsberg, D.M., Ed.; World Scientific: Singapore, 1994; pp. 295–374. [Google Scholar]
  118. Reyes-Gasga, J.; Krekels, T.; van Tendeloo, G.; van Landuyt, J.; Amelinckx, S.; Bruggink, W.H.M.; Verweij, H. 3-d vacancy ordered superstructures in “homogeneous” YBa2Cu3O7−δ. Physica C 1989, 159, 831–848. [Google Scholar] [CrossRef]
  119. Alario-Franco, M.A.; Chaillout, C.; Capponi, J.J.; Chenavas, J.; Marezio, M. A family of non-stoichiometric phases based on Ba2YCu3O7−δ (0 ≤ δ ≤ 1). Physica C 1988, 156, 455–460. [Google Scholar] [CrossRef]
  120. Sonntag, R.; Hohlwein, D.; Brückel, T.; Collin, G. First observation of superstructure reflections by neutron diffraction due to oxygen ordering in YBa2Cu3O6.35. Phys. Rev. Lett. 1991, 66, 1497–1500. [Google Scholar] [CrossRef]
  121. Zimmermann, V.M.; Schneider, J.R.; Frello, T.; Andersen, N.H.; Madsen, J.; Käll, M.; Poulsen, H.F.; Liang, R.; Dosanjh, P.; Hardy, W.N. Oxygen-ordering superstructures in underdoped YBa2Cu3O6+x studied by hard x-ray diffraction. Phys. Rev. B 2003, 68, 104515. [Google Scholar] [CrossRef]
  122. Ricci, A.; Poccia, N.; Campi, G.; Coneri, F.; Caporale, A.S.; Innocenti, D.; Burghammer, M.V.; Zimmermann, M.; Bianconi, A. Multiscale distribution of oxygen puddles in 1/8 doped YBa2Cu3O6.67. Sci. Rep. 2013, 3, 2383. [Google Scholar] [CrossRef]
  123. Ricci, A.; Poccia, N.; Campi, G.; Coneri, F.; Barba, L.; Arrighetti, G.; Polentarutti, M.; Burghammer, M.; Sprung, M.V.; Zimmermann, M.; et al. Networks of superconducting nano-puddles in 1/8 doped YBa2Cu3O6.5+y controlled by thermal manipulation. New J. Phys. 2014, 16, 053030. [Google Scholar] [CrossRef]
  124. Campi, G.; Ricci, A.; Poccia, N.; Bianconi, A. Imaging Spatial Ordering of the Oxygen Chains in YBa2Cu3O6+y at the Insulator-to-Metal Transition. J. Supercond. Nov. Mater. 2014, 27, 987–990. [Google Scholar] [CrossRef]
  125. Rohler, J.; Loeffen, P.W.; Mullender, S.; Conder, K.; Kaldis, E. Local structure studies of the underdoped-overdoped transition in YBa2Cu3Ox. In High-Tc Superconductivity 1996: Ten Years after the Discovery; Kaldis, E., Liarokapis, E., Muller, K.A., Eds.; NATO ASI Series; Kluwer Academic Publ.: Dordrecht, The Netherlands, 1997; pp. 469–502. [Google Scholar]
  126. Conder, K.; Zech, D.; Kruger, C.H.; Kaldis, E.; Keller, H.; Hewat, A.W.; Jilek, E. Indications for a phase separation in YBa2Cu3O7−x. In Phase Separation in Cuprate Superconductors; Sigmund, E., Muller, K.A., Eds.; Springer: Berlin/Herdelberg, Germany, 1994; pp. 210–224. [Google Scholar]
  127. Kaldis, E.; Röhler, J.; Liarokapis, E.; Poulakis, N.; Conder, K.; Loeffen, P.W. A displacive structural transformation in the CuO2 planes of YBa2Cu3Ox at the underdoped-overdoped phase separation line. Phys. Rev. Lett. 1997, 79, 4894–4897. [Google Scholar] [CrossRef]
  128. Calamiotou, M.; Gantis, A.; Palles, D.; Lampakis, D.; Liarokapis, E.; Koufoudakis, A. Phase separation and internal strains in the mixed La0.5R0.5Ba2Cu3Oy compounds (R = rare-earth element). Phys. Rev. B 1998, 58, 15238–15246. [Google Scholar] [CrossRef]
  129. Bogachev, G.; Abrashev, M.; Iliev, M.N.; Poulakis, N.; Liarokapis, E.; Mitros, C.; Koufoudakis, A.; Psyharis, V. Raman study of R0.5Pr0.5Ba2Cu3O7−δ. Phys. Rev. B 1994, 49, 12151–12158. [Google Scholar] [CrossRef] [PubMed]
  130. Palles, D.; Liarokapis, E.; Leventouri, T.H.; Chakoumakos, B.C. The effect of Ca substitution on the structure and the Raman active phonons in Y1−xCaxBa2Cu3O7−δ. J. Phys. Condens. Matter 1998, 10, 2515–2524. [Google Scholar] [CrossRef]
  131. Friedl, B.; Thomsen, C.; Cardona, M. Determination of the superconducting gap in RBa2Cu3O7−δ. Phys. Rev. Lett. 1990, 65, 915–918. [Google Scholar] [CrossRef]
  132. Gantis, A.; Calamiotou, M.; Palles, D.; Lampakis, D.; Liarokapis, E. Phase formation and lattice strain in superconducting compound Y1−xLaxBa2Cu3Oy (0 ≤ x ≤ 1). Phys. Rev. B 2003, 68, 064502. [Google Scholar] [CrossRef]
  133. Calamiotou, M.; Gantis, A.; Margiolaki, I.; Palles, D.; Siranidi, E.; Liarokapis, E. Phase separation, microstructure and superconductivity in the Y1−xPrxBa2Cu3Oy compounds. J. Condens. Matter 2008, 20, 395224. [Google Scholar] [CrossRef]
  134. Rosen, H.J.; Macfarlane, R.M.; Engler, E.M.; Lee, V.Y.; Jacowitz, R.D. Systematic Raman study of effects of rare-earth substitution on the lattice modes of high Tc superconductors. Phys. Rev. 1988, 38, 2460–2465. [Google Scholar] [CrossRef]
  135. Chang, I.F.; Mitra, S.S. Long wavelength optical phonons in mixed crystals. Adv. Phys. 1971, 20, 359–404. [Google Scholar] [CrossRef]
  136. Röhler, J.; Friedrich, C.; Granzow, T.; Kaldis, E.; Böttger, G. High Temperature Superconductivity; AIP Conf. Proc. No. 483; Barnes, S.E., Ed.; AIP: New York, NY, USA, 1999; pp. 320–323. [Google Scholar]
  137. Lampakis, D.; Liarokapis, E.; Karpinski, J.; Panagopoulos, C.; Nishizaki, T. Local lattice distortions and phase separation in cuprates. J. Supercond. 2004, 17, 121–125. [Google Scholar] [CrossRef]
  138. Liarokapis, E.; Lampakis, D.; Palles, D.; Karpinski, J.; Panagopoulos, C. A Raman view of lattice distortions an charge transfer in cuprates. J. Phys. Chem. Solid 2006, 67, 2065–2071. [Google Scholar] [CrossRef]
  139. Lampakis, D.; Palles, D.; Liarokapis, E.; Kazakov, S.M.; Karpinski, J. Hydrostatic pressure induced phase separation in the YBa2Cu4O8 superconductor. J. Phys. Rev. B 2005, 72, 014539. [Google Scholar] [CrossRef]
  140. Osada, M.; Kakihana, M.; Käll, M.; Börjesson, L. Pressure-induced effects in high-Tc superconductors: Raman scattering as a probe of charge-lattice dynamics under high pressure. Physica C 2001, 357–360, 142–145. [Google Scholar] [CrossRef]
  141. Liarokapis, E.; Lampakis, D.; Panagopoulos, C.; Nishizaki, T. High pressure study of Bi-2212. High Press. Res. 2003, 23, 111–115. [Google Scholar] [CrossRef]
  142. Orenstein, J.; Millis, A.J. Advances in the Physics of high-temperature superconductivity. Science 2000, 288, 468–474. [Google Scholar] [CrossRef]
  143. Lampakis, D.; Liarokapis, E.; Panagopoulos, C. Micro-Raman evidence for topological charge order across the superconducting dome of La2−xSrxCuO4. Phys. Rev. B 2006, 73, 174518. [Google Scholar] [CrossRef]
  144. Liarokapis, E.; Lampakis, D.; Nishizaki, T.; Panagopoulos, C. Raman studies of the high pressure effects in high Tc superconductors. High Press. Res. 2000, 18, 109–116. [Google Scholar] [CrossRef]
  145. Koch, U.; Lotter, N.; Wittig, J.; Assmus, W.; Gegenheimer, B.; Winzer, K. Pressure dependence of Tc for single crystal YBa2Cu3O7−x up to 10 GPa. Solid State Commum. 1988, 67, 959–963. [Google Scholar] [CrossRef]
  146. Scholtz, J.J.; van Eenige, E.N.; Wijngaarden, R.J.; Griessen, R. Pressure dependence of Tc and Hc2 of YBa2Cu3O8. Phys. Rev. B 1992, 45, 3077–3082. [Google Scholar] [CrossRef] [PubMed]
  147. Calamiotou, M.; Gantis, A.; Siranidi, E.; Lampakis, D.; Karpinski, J.; Liarokapis, E. Pressure-induced lattice instabilities and superconductivity in YBa2Cu4O8 and optimally doped YBa2Cu3O7−δ. Phys. Rev. B 2009, 80, 214517. [Google Scholar] [CrossRef]
  148. Karpinski, J.; Kazakov, S.M.; Angst, M.; Mironov, A.; Mali, M.; Roos, J. Influence of Sr substitution on the structure, charge distribution, and critical temperature of Y(Ba1−xSrx)2Cu4O8 single crystals. Phys. Rev. B 2001, 64, 094518. [Google Scholar] [CrossRef]
  149. Calamiotou, M.; Gantis, A.; Lampakis, D.; Siranidi, E.; Liarokapis, E.; Margiolaki, I.; Conder, K. Pressure-induced phase separation in the Y123 superconductor. Eur. Phys. Lett. 2009, 85, 26004. [Google Scholar] [CrossRef]
  150. Kenichi, T. Evaluation of the hydrostaticity of a helium-pressure medium with powder x-ray diffraction techniques. J. Appl. Phys. 2001, 89, 662–668. [Google Scholar]
  151. Calamiotou, M.; Parisiades, P.; Siranidi, E.; Lampakis, D.; Conder, K.; Liarokapis, E. Pressure induced lattice effects in pure and near optimally doped La2−xSrxCuO4. Physica C 2019, 565, 1353516. [Google Scholar] [CrossRef]
  152. Ohtani, T.; Himeda, Y.; Norimatsu, Y.; Akiyama, N. Tc increase of Bi2212 induced by exposure to organic liquids or by photocatalytic effect using TiO2. Bull. Okayama Univ. Sci. 2016, 52, 35–44. [Google Scholar]
  153. Raghuveer, V.; Thampi, R.; Xanthopoulos, N.; Mathieu, H.J.; Viswanathan, B. Rare earth cuprates as electrocatalysts for ethanol oxidation. Solid State Ion. 2001, 140, 263–274. [Google Scholar] [CrossRef]
  154. Gupta, M.; Gupta, R. Pressure dependence of the hole concentration in superconducting La1.85Sr0.15CuO4. Physica C 1991, 173, 381–386. [Google Scholar] [CrossRef]
  155. Ganguly, P.; Shah, N.; Phadke, M.; Ramaswamy, V.; Mulla, I.S. Deviations from Vegard’s law: Charges in the c-axis parameter in La2−xSrxCuO4−d in relation to the insulator-superconductor-metal transition. Phys. Rev. B 1993, 47, 991–995. [Google Scholar] [CrossRef]
Figure 1. Typical spectra of YBa2Cu3Ox at room temperature using the 514.5 nm laser line in the y(cc)y (a) and y(xx)y (b) polarization.
Figure 1. Typical spectra of YBa2Cu3Ox at room temperature using the 514.5 nm laser line in the y(cc)y (a) and y(xx)y (b) polarization.
Condensedmatter 04 00087 g001
Figure 2. Apex (a) and in-phase (b) phonon energy and width dependence on amount of oxygen content for YBa2Cu3Ox. Corresponding data for Y1−yCayBa2Cu3O7−δ are presented with % amount of Ca. Error bars indicate statistical distribution of data from various microcrystallites.
Figure 2. Apex (a) and in-phase (b) phonon energy and width dependence on amount of oxygen content for YBa2Cu3Ox. Corresponding data for Y1−yCayBa2Cu3O7−δ are presented with % amount of Ca. Error bars indicate statistical distribution of data from various microcrystallites.
Condensedmatter 04 00087 g002
Figure 3. Apex and in-phase oxygen modes excited with various laser lines for optimal doped (a) and underdoped (b) compounds.
Figure 3. Apex and in-phase oxygen modes excited with various laser lines for optimal doped (a) and underdoped (b) compounds.
Condensedmatter 04 00087 g003
Figure 4. Relative amounts of coexisting phases in YBa2Cu3Ox compound at different doping levels.
Figure 4. Relative amounts of coexisting phases in YBa2Cu3Ox compound at different doping levels.
Condensedmatter 04 00087 g004
Figure 5. Histograms of apex and in-phase peak positions of YBa2Cu3Ox showing various phases.
Figure 5. Histograms of apex and in-phase peak positions of YBa2Cu3Ox showing various phases.
Condensedmatter 04 00087 g005
Figure 6. Raman spectra from selected La concentrations for polarizations along the c-axis (a) or the ab-plane (b).
Figure 6. Raman spectra from selected La concentrations for polarizations along the c-axis (a) or the ab-plane (b).
Condensedmatter 04 00087 g006
Figure 7. Relative intensity of the three phases as obtained from the B1g-symmetry phonon.
Figure 7. Relative intensity of the three phases as obtained from the B1g-symmetry phonon.
Condensedmatter 04 00087 g007
Figure 8. Raman spectra of Y1−yCayBa2Cu3Ox for the y(xx)y scattering polarization.
Figure 8. Raman spectra of Y1−yCayBa2Cu3Ox for the y(xx)y scattering polarization.
Condensedmatter 04 00087 g008
Figure 9. Modification of the Cupl and Oap distance in Å for the Y1−yCayBa2Cu3Ox compound.
Figure 9. Modification of the Cupl and Oap distance in Å for the Y1−yCayBa2Cu3Ox compound.
Condensedmatter 04 00087 g009
Figure 10. Relative intensity of 428 cm−1 mode and its correlation with amount of oxygen.
Figure 10. Relative intensity of 428 cm−1 mode and its correlation with amount of oxygen.
Condensedmatter 04 00087 g010
Figure 11. Pressure dependence of the lattice constants (a-, b-, and c-axis) and lattice strain (bottom graphs) for YBa2Cu3O7−δ (Y1237), PrBa2Cu3O7−δ (Pr1237), and YBa2Cu4O8 (Y1248) compounds. Full (empty) symbols for increasing (decreasing) pressure.
Figure 11. Pressure dependence of the lattice constants (a-, b-, and c-axis) and lattice strain (bottom graphs) for YBa2Cu3O7−δ (Y1237), PrBa2Cu3O7−δ (Pr1237), and YBa2Cu4O8 (Y1248) compounds. Full (empty) symbols for increasing (decreasing) pressure.
Condensedmatter 04 00087 g011
Figure 12. Room temperature pressure Raman data for the Ag-symmetry phonons of the YBa2Cu3O7−δ (Y1237, triangles) and PrBa2Cu3O7−δ (Pr1237, squares) systems.
Figure 12. Room temperature pressure Raman data for the Ag-symmetry phonons of the YBa2Cu3O7−δ (Y1237, triangles) and PrBa2Cu3O7−δ (Pr1237, squares) systems.
Condensedmatter 04 00087 g012
Figure 13. Pressure dependence for Y1237 of Ba atom z-relative position (as compared with the c-axis) and the Cu2-Cu1 distance in Å (open triangles, for decompression).
Figure 13. Pressure dependence for Y1237 of Ba atom z-relative position (as compared with the c-axis) and the Cu2-Cu1 distance in Å (open triangles, for decompression).
Condensedmatter 04 00087 g013
Figure 14. Top: dependence of the c-axis for La1.85Sr0.15CuO4 at increasing (triangles), decreasing (circles) pressures, and for increasing pressure of La2CuO4 (squares). Bottom: pressure dependence of the La/Sr phonon energy (triangles) and width (circles) of La1.85Sr0.15CuO4.
Figure 14. Top: dependence of the c-axis for La1.85Sr0.15CuO4 at increasing (triangles), decreasing (circles) pressures, and for increasing pressure of La2CuO4 (squares). Bottom: pressure dependence of the La/Sr phonon energy (triangles) and width (circles) of La1.85Sr0.15CuO4.
Condensedmatter 04 00087 g014

Share and Cite

MDPI and ACS Style

Liarokapis, E. Probing Phase Separation and Local Lattice Distortions in Cuprates by Raman Spectroscopy. Condens. Matter 2019, 4, 87. https://doi.org/10.3390/condmat4040087

AMA Style

Liarokapis E. Probing Phase Separation and Local Lattice Distortions in Cuprates by Raman Spectroscopy. Condensed Matter. 2019; 4(4):87. https://doi.org/10.3390/condmat4040087

Chicago/Turabian Style

Liarokapis, Efthymios. 2019. "Probing Phase Separation and Local Lattice Distortions in Cuprates by Raman Spectroscopy" Condensed Matter 4, no. 4: 87. https://doi.org/10.3390/condmat4040087

Article Metrics

Back to TopTop