Evidence of Lattice Strain as a Precursor to Superconductivity in BaPb0.75Bi0.25O3
Abstract
In this work, we have investigated the precursor effects to superconductivity in BaPb0.75Bi0.25O3 using temperature dependent resistivity, x-ray diffraction technique and photoemission spectroscopy. The present compound exhibits superconductivity around 11 K (). The synthesis procedure adopted is much simpler as compared to the procedure available in the literature. In the temperature range (10 K–25 K) i.e. above , our results show an increase in both the orthorhombic and tetragonal strain. The well screened features observed in Bi and Pb 4 core levels are indicative of the metallic nature of the sample. The compound exhibits finite intensity at the Fermi level at 300 K and this intensity decreases with decrease in temperature and develops into a pseudogap; the energy dependence of the spectral density of states suggests disordered metallic state. Furthermore, our band structure calculations reveal that the structural transition upon Pb doping results in the closing of the band gap at the Fermi level.
pacs:
61.05.C-,74.25.JbI Introduction
The lead doped BaBiO3, BaPb1-xBixO3 (BPBO), is a well-known 3-dimensional perovskite, which exhibits superconductivity for bismuth compositions in the range Climent-Pascual et al. (2011). This family of compounds has attracted much attention because of its relatively high transition temperature (13 K, at Sleight et al. (1993)), despite the low carrier concentrationThanh et al. (1980). Unlike the cuprates or lanthanides, BPBO family of superconductors are (i) three-dimensional, (ii) without transition metal ions like Cu, La, etc, and (iii) non-magnetic. In the superconducting compositions, the compound is diphasic, with the coexistence of the tetragonal and orthorhombic phasesMarx et al. (1992); Climent-Pascual et al. (2011); the volume fraction of tetragonal phase dictates the , thus being the superconducting phase. Recently, Nicoletti et al.Nicoletti et al. (2017) have shown that a short range CDW order exists adjacent to the superconducting phase (i.e., above 7 K) in superconducting BaPb0.78Bi0.22O3. Further, recent study by Parra et al.Parra et al. (2021), shows that an electronic reorganization into a 2D granular superconductor takes place when BPBO approaches the superconductor-insulator transition before ultimately transforming to an insulator. Superconductivity is generally seen adjacent to exotic symmetry-breaking ground states, like AFM and CDW ordersNicoletti et al. (2017). Usually, a complex but subtle interplay between spin, charge, crystal and electronic structures shape the resulting electronic properties. BPBO, which manifests such emergent electronic properties, is an ideal candidate to engineer novel superconducting devices such as quantum computing circuitsKim et al. (2019).
The crystal structure of superconducting compositions have been a major point of confusion – Cox and SleightCox and Sleight (1976) have reported that the compound is tetragonal in the range ; Khan et al.Khan et al. (1977) observe that the compound has a distorted orthorhombic crystal structure belonging to the phase for all compositions, whereas, Oda et al.Oda et al. (1985) concluded that the compound is orthorhombic for all compositions with along with a structural transition from an orthorhombic to a monoclinic phase at about 160 K in case of BaPb0.75Bi0.25O3; in a later study, Asano et al.Asano et al. (1988) and Oda et al.Oda et al. (1986) proposed that the differences in the sample preparation techniques could result in the compounds stabilising in either orthorhombic or tetragonal phase, each having a distinct .
Investigations of the phase separation of BaPb1-xBixO3 by Giraldo-Gallo et al. using high quality single crystalsGiraldo-Gallo et al. (2015) have lead to the observation that the structural dimorphism takes the form of partially disordered nanoscale stripes for optimal composition of . As the bismuth composition deviates from its optimal value, the volume of tetragonal phase reduces and the tetragonal stripes reduce to islands embedded in a matrix of orthorhombic BaPb1-xBixO3. Point-contact spectroscopic measurements performed on BaPb1-xBixO3 () indicate that the metal-insulator transition (MIT) is driven by disorderLuna et al. (2014). Magnetoresistance studies performed on BaPb0.75Bi0.25O3 epitaxial thin films grown on LaLuO3 by Harris et al. Harris et al. (2018), show superconducting fluctuations in epitaxial thin films well above ; and in case of the thickest films with thickness above 100 nm these fluctuations were observed till 27 K.
Several photoemission spectroscopy measurement studies on BaPb1-xBixO3 have been reported, some performed on single crystals Wertheim et al. (1982); Matsuyama et al. (1989); Winiarski et al. (1991); Namatame et al. (1993), and few more of polycrystalline samplesSakamoto et al. (1987); Kostikova et al. (2001); Korolkov et al. (2002). Most studies do not reveal any charge disproportionation in the compoundWertheim et al. (1982); Sakamoto et al. (1987); Matsuyama et al. (1989); Winiarski et al. (1991); Namatame et al. (1993), notable exception being Kostikova et alKostikova et al. (2001). and Korolkov et al.Korolkov et al. (2002), who noticed the presence of both PbII and PbIV, along with BiIII and BiV in their compound under study – BaPb0.8Bi0.2O3. Room temperature (RT) x-ray photoelectron spectroscopic (XPS) studies by Wertheim et al.Wertheim et al. (1982) and Sakamoto et al.Sakamoto et al. (1987) show no indication of any density of states (DOS) at Fermi level (). However, later ultraviolet photoemission studies (UPS) by Matsuyama et al.Matsuyama et al. (1989) (at RT and 180 K), and Namatame et al.Namatame et al. (1993) (at liquid-nitrogen temperature (LNT)) indicate a finite DOS at . The UPS studies by Matsuyama et al.Matsuyama et al. (1989) show that BaPb0.85Bi0.15O3 has a clear Fermi-edge structure characteristic of normal metal; suggesting that the superconductivity would be driven by Cooper pairing of electrons in the Fermi-liquid states. According to the BCS theory, good metals do not become superconductors because of weak-electron phonon coupling, which makes their claim a very interesting one. It is also important to note that their measurements were done at 180 K, a temperature significantly higher than the of the compound. The electronic structure studies by Namatame et al.Namatame et al. (1993) was performed at LNT in an attempt to obtain a picture of the electronic structure of BaPb1-xBixO3 for a wide range of compositions. Further, studies exploring effects such as superconducting fluctuations discovered by Harris et al.Harris et al. (2018) indicate a presence of precursor effects before the onset of bulk superconductivity. To understand the manifestation of co-existence of two structural phases in the electronic structure or precursor effects to superconductivity in BaPb0.75Bi0.25O3 we have carried out temperature dependent crystal structure and electronic structure studies on the compounds above .
Another interesting aspect of the BaPb1-xBixO3-δ series is the strong dependence of the superconductivity on the oxygen content in the sampleHashimoto and Kawazoe (1993). The authors found that an oxygen deficiency of greater than destroyed the superconductivity. Hence, an elaborate preparation procedure is generally used to ensure that sample is well-oxygenated. The typical preparation route is via solid state reactions of powdered raw materials – BaCO3, Pb3O4 and Bi2O3 – taken in stoichiometric proportions. These raw materials are mixed in ethanol and ground well using a ball-milling machine. The mixture is then heated at 720 for 12 h under flowing oxygen. Further, the calcined sample is ground again, pressed into a pellet and then sintered at 800 for 12 h, in oxygen atmosphere.
In this paper, we investigate the possible precursor effects leading to the superconductivity. We present a detailed investigation of temperature dependent resistivity, crystal structure and electronic structure studies on the compound under study. The electrical transport and magnetic susceptibility measurements indicate a superconducting transition around 11 K. Further, these studies reveal that the disorder plays a very important role in the electrical resistivity and the electronic structure of the compound. The spectral DOS near the Fermi level displays a square-root of energy dependence indicating the compound to be a disordered metal above . The temperature evolution of the symmetrised spectral density of states indicates the development of a disorder-induced pseudogap at the Fermi level. Band structure calculations performed using the TB-mBJ correlation functional indicates that the monoclinic to tetragonal structural phase transition upon Pb doping of the parent compound plays a major role in the closing of the band gap observed in the compound under study. Additionally, we have adopted a much simpler synthesis procedure as compared to the ones available in the literature.
II Experimental
The polycrystalline samples of BaPb0.75Bi0.25O3 were prepared via solid state route from BaCO3, Bi2O3 and Pb3O4. The raw materials were first preheated at 450 for 2 hours to remove any traces of moisture absorbed by them. The proper stoichiometric ratios of the raw materials were then ground thoroughly in a agate mortar for 24-48 hours. This finely ground powder was then calcined in a box furnace at 900 for 48 hours. The heating rate was maintained at 5 /min and the samples were allowed to cool down to room temperature naturally inside the box furnace. The calcined powders were then sintered in the box furnace at 900 for around 150 hours, with intermediate grindings every 24 hours. The final sintering was performed at 950 . After each grinding, the powders were then pressed into pellets of 10 mm diameter under a hydrostatic pressure of 5 tons. It was found that covering the pellets with a thin layer of powder of the sample ensured a good stoichiometric compound, and prevented the escape of Bi or Pb from the surface of the pellet. It is to be noted that despite not sintering the sample in oxygen atmosphere, we were able to obtain a superconducting transition.
The temperature dependent xrd (T-XRD) measurements were performed using Rigaku’s Smart Lab x-ray diffractometer powered by a 9kW rotating anode x-ray generator. The T-XRD patterns were collected using Cu Kα radiations and were collected in the 2 range of 19∘ to 87∘ at a scanning speed of 1∘ per minute with a step size of 0.02∘. The Magnetic Properties Measurement System (MPMS) from Quantum Design, Inc. was used in the measurement of temperature dependent dc magnetization measurements. The measurement was performed at an applied magnetic field of 0.5 T, in the temperature range of 300K to 2K. Electrical resistivity measurements were performed on the sample using Physical Properties Measurement System (PPMS) setup from Quantum Design, Inc, in the same temperature range, using a standard 4-probe DC resistivity method. The probes were attached to the sample using high quality silver paste, and the data collection was done during the cooling cycle. The field emission scanning electron microscopy (FE-SEM) micrographs were procured using Nova Nano SEM-450 at room temperature at a chamber pressure of mbar, with a scanning voltage of 10 kV.
The temperature dependent photoemission spectra were collected on Scienta R4000 hemispherical analyser using a monochromatic Al Kα (1486.6 eV) x-ray source, He I (21.2 eV), and He II (40.8 eV) ultraviolet radiations at various temperatures between 300 K to 30 K. The binding energy scale was calibrated by measuring the Fermi level of Ag pellet, cleaned in situ by argon ion sputtering, using monochromatic Al Kα and HeI. A clean sample surface was obtained by fracturing the mounted samples in a chamber having vacuum better than mbar. The base pressure during the measurement was mbar.
III Computational
Band structure calculation using self-consistent full potential linear augmented-plane-wave (LAPW) were performed for BaPb0.75Bi0.25O3 using the code implemented in Elkelk . We have used local density approximation (LDA)Perdew and Wang (1992) for the exchange potential and Tran-Blaha modified Becke-Johnson (TB-mBJ) potentialTran and Blaha (2009) for the correlation part. The calculations were performed using the structural parameters obtained from the refinement of the xrd patterns of phase of BaPb0.75Bi0.25O3 taken at room temperature. A supercell was constructed and one atom of Pb was replaced by Bi, to account for the 75% doping of Pb at the Bi-site. The muffin-tin shape approximation for the potential well in the crystal lattice employed in the calculations used radii of 2.8, 2.56, 2.43 and 1.43 bohr for Ba, Bi, Pb and O, respectively. The difference in total energy required for the termination of self-consistent cycles was set to be less than Hartree/cell. In order to understand the evolution of the states at , we have also performed TB-mBJ calculations on BaBiO3 using the structural parameters of BaPb0.75Bi0.25O3, using a muffin-tin radii of 2.8, 2.29 and 1.72 bohr for Ba, Bi and O, respectively.
IV Results and Discussions
IV.1 General Characterization

Figure 1(a) shows the crystal structure of the compound under study; also shown are the Bi(Pb)-O-Bi(Pb) angles of the apical and basal planes of the two phases. Panels (b) and (c) for the same figure (1) show FE-SEM micrographs of the BaPb0.75Bi0.25O3 pellet. The average grain size was estimated to be 8 m.

In figure 2, we show the typical Rietveld fitting of the xrd patterns of BaPb0.75Bi0.25O3 collected at 10 K and 300 K. A two-phase Rietveld refinement, consisting of tetragonal I4/mcm phase and orthorhombic Ibmm phase, has been performed to fit the experimental data; the crystal structure and the room temperature lattice parameters are in line with the earlier worksMarx et al. (1992); Climent-Pascual et al. (2011); Giraldo-Gallo et al. (2015); Nicoletti et al. (2017). The absence of unindexed peaks points to the purity of the sample. The lattice parameters and atomic positions obtained from the refinement of the patterns collected at 300 K and 10 K are shown in Table 1.
Tetragonal Phase | Orthorhombic phase | |||||
10 K | 300 K | 10 K | 300 K | |||
Lattice parameters | ||||||
a (Å) | 6.02815 | 6.05055 | 6.07851 | 6.08959 | ||
b (Å) | - | - | 6.04651 | 6.06051 | ||
c (Å) | 8.62396 | 8.60776 | 8.51402 | 8.55560 | ||
Atomic Positions | ||||||
Ba | x | 0.48983 | 0.50653 | |||
O1 | x | 0.0034 | 0.0472 | |||
O2 | x | 0.26216 | 0.22212 | |||
y | 0.76216 | 0.72212 | ||||
z | 0 | 0 | 0.9892 | 0.6102 |
The quality of fit, when modelled using a single phase – either tetragonal I4/mcm or orthorhombic Ibmm – is significantly lower as compared to the two-phase fit. The inset (b1) of figure 2(b) shows the fit of the peaks in the range when modelled using a single tetragonal phase. A similar fit is observed when the xrd patterns are refined using a single orthorhombic phase (not shown). Inset (b2) of figure 2(b) shows the peaks in the range when fit using a diphasic model. A sample with tetragonal symmetry shows a single pair of (404) peak (corresponding to Cu Kα1 and Cu Kα2)Hashimoto and Kawazoe (1993). On the other hand, in case of orthorhombic symmetry, this peak splits into (404) and (044). To ensure a proper fit, it is necessary to model data using two phases, which result in the decomposition of twin peaks into its 6 constituent peaks – one pair corresponding to (404) reflection from tetragonal phase, one pair each corresponding to (404) and (044) reflections from the orthorhombic phase – thus indicating the coexistence of I4/mcm and Ibmm phases.

In figure 3(a), we show the resistivity versus temperature data for the compound under study. As the temperature drops, the resistivity gradually increases till 250 K, below which a change of slope is observed. The resistivity increases rapidly until 10.5 K (), below which it drops sharply indicating the onset of superconductivity, as seen in the inset of figure 3(a). The figure 3(b) shows the field-cooled (FC) magnetic moment of the sample as function of temperature. As reported in the literatureUchida et al. (1988); Hashimoto et al. (1994), the sample is diamagnetic in the entire temperature range. The magnetic moment shows sharp drop at around 11 K indicating the onset of the superconducting transition. The combined resistivity and dc magnetic moment measurements suggest the onset of superconductivity at K.

Figures 4(a)-(b) illustrate the fit of the resistivity data in two temperature ranges, using different models. The resistivity data in figure 3(a) is roughly divided into 3 regions based on this. In region 1, ( K - 300 K), the electrical transport is dominated by activated behaviour, which is appreciable when sufficient thermal energy is available for the electron to be excited across the band gapSchnakenberg (1968); Greaves (1973); Kumar et al. (2007). It is important to note that the studies by Uchida et al.Uchida et al. (1987), indicate that temperature dependence of normal-state resistivity of BaPb1-xBixO3 is like that of semiconductors for . Figure 4(a) shows the variation of as a function of , modelled by equation 1.
(1) |
where, is a pre-exponential factor, is the activation energy, and JK-1 is the Boltzmann’s constant. From the straight line fit of the data, we obtain an activation energy of meV.
As the temperature drops below 250 K, it’s not possible to model the data using Arrhenius-type behaviour. We observe a transition of the resistivity behaviour from Arrhenius to variable range hopping (VRH). Mott variable-range hopping model has a characteristic temperature dependence of resistivityMott (1972); Mott and Davis (2012) given by
(2) |
where is a parameter that describes the dimensionality of the conduction in the system, and is a pre-exponential constant. is the characteristic Mott temperature, which describes the energy barrier of an electron hopping from one localized state to another, given by the equation 3.
(3) |
Here, it is important to note that it was not possible to fit the entire temperature range below 250 K, till the , with a single curve. It was necessary to split them into separate regions with distinct characteristic temperatures and localization lengths. However, while we were able to fit the data using equation 2 mathematically, we found that the ratio of mean hopping distance to electron localization length, given by equation 4, is much less than 1, which is an indication that equation 2 is not applicable in our caseRosenbaum (1991).
(4) |
In the intermediate temperature range, the resistivity data can be fit using Greaves VRHGreaves (1973). The acoustic phonon contribution to the resistivitySchnakenberg (1968); Greaves (1973), takes the form as given in equation 5,
(5) |
where, is the pre-exponential constant and is the characteristic temperature of Greaves VRH.
For Greaves VRH to be valid, the plot of versus should be linear. In this intermediate temperature range () where the deviation from the fit is observed has been taken to be in the higher temperature limit, and in the lower temperature limitSakata et al. (1999); Kumar et al. (2007). Below this temperature, we have been unable to model the resistivity using this equation. Thus, the best fit was observed in the temperature range 76 K to 166 K, and we obtained the characteristic temperature of K. Using the value of states/cm3, as reported by Kitazawa, et al.Kitazawa et al. (1985), we obtain the electron localization length of Å. The deviation of the resistivity from the fit begins at 76 K at the lower end, and 166 K at the higher end. Thus, we may estimate that the Debye temperature of the compound lies between 304 K to 332 K.
IV.2 Temperature dependent xrd studies
To understand the structural link with the transport properties and understand the behaviour of phase separation, we have carried out temperature dependent xrd on the compound. Figures 5-7 show the variation of the lattice parameters, bond lengths, bond angles, lattice strain, and unit cell volume with temperature. The xrd patterns were analysed using Rietveld profile refinement and the structural parameters were obtained as a function of temperature. The goodness of fit obtained is in the range 1.45 – 1.63 for all temperatures. As the temperature drops from 300 K to 10 K, we observe that the number of peaks remain the same. This suggests the absence of any structural transitions at lower temperatures, in contrast to some of the earlier studiesOda et al. (1985); Asano et al. (1988). Moreover, we observe that the fraction of the tetragonal phase is 70%, and is largely temperature independent (see figure 7(a)).
Keeping these results in mind, revisiting the resistivity studies of the previous section gives an opportunity to understand the high transition temperature of VRH from Arrhenius type behaviour. Firstly, at low temperature, the thermal energy available for excitation of electron is insufficient, thus leading the electrical transport behaviour away from activated behaviour. In such a scenario, it is favorable for the electron to hop to a site farther than the nearest neighbour with a lower potential, giving rise to observed VRH behaviourKumar et al. (2007). Furthermore, we hypothesize that this high transition temperature is due to the role of disorder in the system. Two leading contributions to disorder are (a) the structural phase coexistence of orthorhombic and tetragonal phases at all temperatures, with fraction of the tetragonal phase constant at 70%, and largely temperature independent, and (b) the composition lying within the transition region between metal and semiconducting regions. The work on BaPb1-xBixO3 by Uchida et al.Uchida et al. (1987) suggests that the compound under study lies in the region where both metal and semiconducting regions co-exist.
Figures 5(a) and 5(b) illustrate the variation of the lattice parameters of I4/mcm and Ibmm phases with temperature. In order to facilitate the comparison of the structural data with resistivity measurements, the entire temperature range have been divided into three regions, identical to that in figure 3(a).

In the activated region, the parameter of the tetragonal phase, decreases by 0.0014%, whereas increases by 0.001%. As the temperature drops to region II, the decrement in is 0.0022%, and the increment in is 0.001%. In region III, the lattice parameters are nearly constant within the range of experimental error.
The antiphase Bi(Pb)-O6 octahedral rotations about the -axis increase with decreasing temperature. The decreasing Bi(Pb)-O2-Bi(Pb) bond angle results in the contraction of the parameter. This reduction in the bond angle effectively contracts the interstitial volume around the Barium atom, thus increasing the O1-Ba-O2 angle, which drives the increase in the parameter with decreasing temperature.

The lattice parameters of the Ibmm phase, however, paint a more complex picture. The figure 5(a) and (b) respectively show the variation of , , and parameters of the Ibmm phase. The lattice parameters and reduce monotonically until 75 K, dictated by the reducing bond lengths and bond angles; further, decreases monotonically until the beginning of region II, and remains constant in region II. Below 75 K, increases, while decreases; remains nearly constant. The interplay between the bond lengths and bond angles leading to the observed behaviour of the lattice parameters are easily understood from graphs in panels (e) and (f) of figure 5.
From figures 6(a) and (b), we clearly observe that the nature of Bi(Pb)-O6 octahedra is different in the two phases. In the case of tetragonal phase, there are two long apical Bi-O bonds and four short basal Bi-O bonds, while in the case of orthorhombic phase, the trend is reversed. It is interesting to note that, Gallo et al.Giraldo-Gallo et al. (2015), have observed that the phase separation takes the form of partially disordered nanoscale stripes. It is possible that the difference in the nature of distortion could be one of the reasons for such stripes. These experiments were carried out by the authors on single crystalline BaPb0.75Bi0.25O3 and have observed nanoscale structural phase separation.

To understand the nature of phase separation in the present work, one can propose two scenariosChmaissem et al. (2003); Sharma et al. (2022) (a) two phases coexisting in the same grain; (b) two phases lying in two different grains. In the first scenario, it is expected that one phase grows at the cost of the other while in the second case the phase fraction remains the same. Towards this, we have calculated the average size of the structural phases as a function of temperature using Scherrer formulaPatterson (1939) (equation 6).
(6) |
where, is a constant, Å is the x-ray wavelength, is the FWHM of the peak at angle . Our results show that the size of the structural phases remains the same through temperature range of study. This behaviour suggests the second scenario proposed above. However, temperature dependent transmission electron microscopy (TEM) studies will be helpful to unravel the nature of phase separation in this polycrystalline compound.
Graphs in figure 6(c) and (d) show the temperature variation of tetragonal and orthorhombic strain. In the forth coming section, we will be using the term lattice strain synonymous to tetragonal/orthorhombic strain. In the tetragonal phase (figure 6(c)), we observe the lattice strain increases with decrease in temperature and in the orthorhombic phase (figure 6(d)), the lattice strain decreases until around 125 K and below this temperature it is found to increase. It is interesting to note that in the temperature range close to , the lattice strain is maximum possibly playing a role as precursor for superconductivity. Such behaviour has been observed in the case of YBa2Cu3O7 compoundHorn et al. (1987), where the orthorhombic strain is found to be maximum around with little to no anomaly in the unit cell volume. However, in the present study we observe an anomaly in the unit cell volume around 35 K and 65 K, in the tetragonal and orthorhombic phases, respectively, (see figure 7(c)). In this figure, the anharmonic part of the lattice contribution to the unit cell volumes was obtained by fitting with Debye model, as given by equation 7.
(7) |
where, is the volume at absolute zero, is the number of atoms per unit cell, is the Grünesian parameter and is the Debye temperature. The three fitting parameters – , and – were determined using the least square fitting, are shown in table 2. We readily see that the K, the Debye temperature of the majority phase fraction of the compound, obtained from the fit matches well with the estimate obtained from the resistivity data, as previously discussed.
Phase | |||||
---|---|---|---|---|---|
Tetragonal | 313.38 | 0.026 | 330.4 | ||
Orthorhombic | 312.92 | 0.098 | 916.8 |
V Electronic structure studies
The results discussed thus far indicate that the disorder plays a very important role in the sample, especially in the region just above . Signature of disorder is visible not only in transport measurements, but also is expected to manifest in electronic structure. According to BCS theory, good conductors do not show superconductivity because of weak electron phonon coupling. In the present compound, the observed resistivity values lie in the range of metals. It will be interesting to study the temperature dependent behaviour of the electronic states close to the Fermi level and also behaviour of core levels prior to . To visualize and understand the electronic structure of BaPb1-xBixO3, we have performed photoemission spectroscopy measurements on our sample at various temperatures between 300 K and 30 K, using Al Kα x-ray radiation, HeI and HeII ultraviolet radiations. Figure 8 shows the core levels of the compound collected at RT and 30 K.

V.1 Core level studies
Panels (a)–(e) of figure 8 show the core level spectra of Ba 3, Ba 4, O 1, Bi 4, and Pb 4, respectively, at 300 K and 30 K. Our results show that with decrease in temperature there is a decrement in the width of the core level peaks. This is very clearly observed in the case of Ba and O 1s core levels. The Bi and Pb 4 core levels exhibits two features labelled as A and A*. Such features have been observed in Bi core level spectra by Prachi et al.Telang et al. (2022) in the case of Bi based pyrochlore iridate where the sample is metallic and have attributed the features A and A* to be of poorly and well screened features. Such two features have also been observed in the case of Pb 4 by Payne et al.Payne et al. (2007), in the case of PbO2 sample that is metallic. It is interesting to note that in the case of BaBiO3, that is semi conducting, only poorly screened feature is observedBharath et al. (2019); Plumb et al. (2016). The well screened feature arises when the Bi/Pb 4 core hole is screened by the transfer of electrons from the ligand and the poorly screened feature arises when no such transfer occurs.
V.2 xps Valence Band and Band Strucutre Studies

The comparison of xps and HeII valence band spectra collected at 300 K is shown in figure 9. Finite DOS at is clearly visible in both xps and HeII spectra. We observe six features in the spectra labelled as A, B1, B2, B3, C and D. The xps valence band represents the bulk features of the sample. To identify the features, band structure calculations were carried out on BaPb0.75Bi0.25O3, in the tetragonal phase using DFT method, as shown in figure 10. Our results show that the total density of states (TDOS) of TB-mBJ is shifted towards lower energy as compared to the LDA calculations, as seen in panel (a) of figure 10. The partial density of states (PDOS) are plotted in panels (b)–(f) of figure 10. While LDA calculations generally sufficient in case of metallic compounds, it has been well documented that LDA underestimates the band gap at the fermi levelPerdew and Levy (1983); Sham and Schlüter (1983) of semiconducting or insulating compounds. This was the case of the parent compound, BaBiO3, where the gap obtained in the case of LDA calculations was underestimated as compared to the TB-mBJ calculationsBharath et al. (2019). However, the compound under study, BaPb0.75Bi0.25O3, exhibits both metallic and semiconducting behaviours, as discussed previously. Further, we’re also probing the effects of Pb doping in BaBiO3 to understand how the band gap observed in BaBiO3 closes upon Pb doing. Thus, TB-mBJ calculations have been employed to understand the electronic structure of BaPb0.75Bi0.25O3. The experimental spectra were matched with the TB-mBJ calculations with a finite energy shift. Towards this, in the present work, a rigid shift of -0.8 eV was given to the DOS obtained from the TB-mBJ calculation to match with the experimental spectrum.

Our results show that there is finite DOS at the Fermi level that is in line with the experimental spectra. The calculated DOS can be divided into five regions. In region I (-14 to -12 eV), there is significant contribution from Ba states and weak contributions from Bi/Pb and Pb states. Region II (-12 to -7 eV) covers dominant contributions from Bi/Pb and weak contributions from rest of the DOS. Region III (-7eV to -1.5 eV) has contributions from Bi/Pb , , , and O states. In region IV (-1.5 to ) there is dominant contribution from Bi and weak contributions from other states. The region V (above to 2 eV) has dominant contribution from Pb and weak contributions from rest of the states. We now identify the different features based on the above results. The feature represents surface cleanliness. In the case of xps, the intensity of this feature is low as compared to the HeII spectra due to higher sensitivity to surface oxygen. The features B1, B2 and B3 covers region III and the feature encompasses region II. The photoionisation cross section for Al Kα source is more for Bi/Pb as compared to the photoionisation cross section of O , while in the case of HeII spectra, the reverse is true. Hence, the higher intensity of feature observed in xps as compared to HeII spectra suggests the contributions from Bi/Pb states.
Pb doping introduces a significant change in the electronic structure near as compared to BaBiO3. An insulating gap at observed in BaBiO3Bharath et al. (2019) closes on 75% Pb doping. The region between -3 to -1.4 eV which was identified as feature in BaBiO3Bharath et al. (2019) splits into 3 features upon Pb doping, namely, B1, B2 and B3. All the three features have dominant contributions from O states. In figure 10 (g), we have presented the TB-mBJ calculations of monoclinic BaBiO3 (labelled as BBO (Mono)), BaBiO3 using the structural parameters of BaPb0.75Bi0.25O3 (labelled BBO (Tetra)) and BaPb0.75Bi0.25O3. By performing TB-mBJ calculations on BaBiO3 using the structural parameters of BaPb0.75Bi0.25O3, we intend to understand the origin of states at . Our results show that there is finite DOS at in this case. This suggests that the finite DOS appearing at is mainly driven by the crystal structure. Further, the DOS at Fermi level was observed to decrease upon Pb doping, as seen clearly in the inset of figure 10 (g). To ascertain that the position of bismuth does not play an important role in the electronic structure of BaPb0.75Bi0.25O3, we repeated the TB-mBJ calculations by substituting the bismuth atom in each of the different positions of lead. The total DOS was observed to be identical in all the cases.
V.3 UPS Valence Band Studies near


Having identified the states that contribute to the Fermi level (), we now look into the evolution of the states close to . In figure 11(a), we show the temperature dependent valence band collected using HeI spectra. All the spectra are normalised at 100 meV. Our results show that all the spectra show a finite intensity at and this intensity is found to decrease on lowering the temperature. To observe more clearly the temperature behaviour of the states close to , the spectral density of states was obtained by dividing the experimental spectra by the Fermi Dirac distribution and symmetrising it, figure 11(b). The figure shows significant reduction in the intensity as a dip in the vicinity of with decrease in temperature. This dip increases as the temperature drops and a systematic increase in the spectral DOS above 100 meV is observed with decrease in temperature suggesting the stabilization of pseudogap.
It is important to note that Matsuyama et al.Matsuyama et al. (1989) have collected He I and II spectra on BaPb0.85Bi0.15O3 at around 300 K and 180 K. Based on the Fermi edge observed in this compound, they proposed that the superconductivity would be driven by cooper pairing of electrons in Fermi liquid states. To check for this possibility, the SDOS was plotted as a function of . Our results show that the value of thus obtained was 0.5 suggesting the role of disorder existing in the compound, as seen in figure 12(a). Earlier studiesSarma et al. (1998); Kobayashi et al. (2007) have shown that such decrement in the intensity could arise due to the localization of the states at induced by disorder. Under this situation, it is expected that the DOS at follows , where is the temperature; such a behaviour can be clearly observed in panel (b) of figure 12, thus suggesting that the compound is a disordered metal for . However, to observe whether there is any characteristic peak observed close to representing the superconducting state, experiments below are required.
VI Summary
In conclusion, we have studied the structural, transport and electronic properties of the polycrystalline superconducting BaPb0.75Bi0.25O3 sample. The analysis of the temperature dependent resistivity data reveals the compound undergoes a superconducting transition around 11 K. The VRH behaviour of the resistivity at low temperatures in an indication that disorder plays a very important role in the electronic properties of the compound. We believe that the structural dimorphism present in the sample at all temperatures, along with the coexistence of metallic and semiconducting regions in the sample are the likely causes for disorder in the compound under study. To explore the effect of phase separation, structural studies were conducted using xrd which reveal that the compound is dimorphic at all temperatures ranging from 300 K down to 10 K. Further, the evolution of the tetragonal and orthorhombic strain with temperature indicate that the strain is maximum as we approach the . The observation of well screened features in the Bi and Pb spin orbit split 4f core level suggests the metallic nature of the sample. Additionally, we observed a finite density of states at Fermi level. Our band structure studies suggest that the closing of the gap upon Pb doping is due to structural transition from the monoclinic I2/m phase to the tetragonal I4/mcm phase. The temperature-dependent behaviour of the electronic states close to the fermi level was studied using UPS measurements. The signature of disorder was observed in the form of dependence of SDOS near the Fermi level suggests the compound under study for is a disordered metal. Furthermore, the DOS at Fermi level decreased progressively with temperature – an evidence of opening of disorder-induced pseudogap. Our combined crystal structure, transport and electronic structure studies suggest that lattice strain and disorder act as precursor to superconductivity in BaPb0.75Bi0.25O3 compound.
VII Acknowledgements
The authors, Bharath M and R Bindu thank Science and Engineering Research Board (SERB), Department of Science and Technology, Government of India for funding this work. This work is funded under the SERB project sanction order No. EMR-2016-001144.
References
- Climent-Pascual et al. (2011) E. Climent-Pascual, N. Ni, S. Jia, Q. Huang, and R. Cava, Physical Review B 83, 174512 (2011).
- Sleight et al. (1993) A. W. Sleight, J. Gillson, and P. Bierstedt, Solid State Communications 88, 841 (1993).
- Thanh et al. (1980) T. D. Thanh, A. Koma, and S. Tanaka, Applied Physics 22, 205 (1980).
- Marx et al. (1992) D. Marx, P. Radaelli, J. Jorgensen, R. Hitterman, D. Hinks, S. Pei, and B. Dabrowski, Physical Review B 46, 1144 (1992).
- Nicoletti et al. (2017) D. Nicoletti, E. Casandruc, D. Fu, P. Giraldo-Gallo, I. Fisher, and A. Cavalleri, Proceedings of the National Academy of Sciences 114, 9020 (2017).
- Parra et al. (2021) C. Parra, F. C. Niestemski, A. W. Contryman, P. Giraldo-Gallo, T. H. Geballe, I. R. Fisher, and H. C. Manoharan, Proceedings of the National Academy of Sciences 118, e2017810118 (2021).
- Kim et al. (2019) J. Kim, J. Mun, B. Kim, H. G. Lee, D. Lee, T. H. Kim, S. Lee, M. Kim, S. H. Chang, and T. W. Noh, Physical Review Materials 3 (2019), URL https://doi.org/10.1103%2Fphysrevmaterials.3.113606.
- Cox and Sleight (1976) D. Cox and A. Sleight, Solid State Communications 19, 969 (1976).
- Khan et al. (1977) Y. d. Khan, K. Nahm, M. Rosenberg, and H. Willner, physica status solidi (a) 39, 79 (1977).
- Oda et al. (1985) M. Oda, Y. Hidaka, A. Katsui, and T. Murakami, Solid state communications 55, 423 (1985).
- Asano et al. (1988) H. Asano, M. Oda, Y. Endoh, Y. Hidaka, F. Izumi, T. Ishigaki, K. Karahashi, T. Murakami, and N. Watanabe, Japanese journal of applied physics 27, 1638 (1988).
- Oda et al. (1986) M. Oda, Y. Hidaka, A. Katsui, and T. Murakami, Solid State Communications 60, 897 (1986), ISSN 0038-1098, URL https://www.sciencedirect.com/science/article/pii/0038109886903819.
- Giraldo-Gallo et al. (2015) P. Giraldo-Gallo, Y. Zhang, C. Parra, H. Manoharan, M. Beasley, T. Geballe, M. Kramer, and I. Fisher, Nature communications 6, 1 (2015).
- Luna et al. (2014) K. Luna, P. Giraldo-Gallo, T. Geballe, I. Fisher, and M. Beasley, Physical Review Letters 113, 177004 (2014).
- Harris et al. (2018) D. T. Harris, N. Campbell, R. Uecker, M. Brützam, D. G. Schlom, A. Levchenko, M. S. Rzchowski, and C.-B. Eom, Physical Review Materials 2, 041801 (2018).
- Wertheim et al. (1982) G. Wertheim, J. Remeika, and D. Buchanan, Physical Review B 26, 2120 (1982).
- Matsuyama et al. (1989) H. Matsuyama, T. Takahashi, H. Katayama-Yoshida, Y. Okabe, H. Takagi, and S. Uchida, Physical Review B 40, 2658 (1989).
- Winiarski et al. (1991) A. Winiarski, G. Wübbeler, C. Scharfschwerdt, E. Clausing, and M. Neumann, Fresenius’ journal of analytical chemistry 341, 296 (1991).
- Namatame et al. (1993) H. Namatame, A. Fujimori, H. Takagi, S. Uchida, F. De Groot, and J. Fuggle, Physical Review B 48, 16917 (1993).
- Sakamoto et al. (1987) H. Sakamoto, H. Namatame, T. Mori, K. Kitazawa, S. Tanaka, and S. Suga, Journal of the Physical Society of Japan 56, 365 (1987).
- Kostikova et al. (2001) G. Kostikova, D. Korol’Kov, and Y. P. Kostikov, Russian journal of general chemistry 71, 1010 (2001).
- Korolkov et al. (2002) D. V. Korolkov, G. P. Kostikova, and Y. P. Kostikov, Physica C: Superconductivity 383, 117 (2002).
- Hashimoto and Kawazoe (1993) T. Hashimoto and H. Kawazoe, Solid state communications 87, 251 (1993).
- (24) Elk software, URL https://elk.sourceforge.net/.
- Perdew and Wang (1992) J. P. Perdew and Y. Wang, Physical Review B 45, 13244 (1992).
- Tran and Blaha (2009) F. Tran and P. Blaha, Physical Review Letters 102, 226401 (2009).
- Uchida et al. (1988) S. Uchida, H. Hasegawa, K. Kitazawa, and S. Tanaka, Physica C: Superconductivity 156, 157 (1988).
- Hashimoto et al. (1994) T. Hashimoto, H. Kawazoe, and H. Shimamura, Physica C: Superconductivity 223, 131 (1994).
- Schnakenberg (1968) J. Schnakenberg, Physica Status Solidi (b) 28, 623 (1968), URL https://doi.org/10.1002%2Fpssb.19680280220.
- Greaves (1973) G. Greaves, Journal of Non-Crystalline Solids 11, 427 (1973).
- Kumar et al. (2007) R. Kumar, R. Choudhary, M. Ikram, D. Shukla, S. Mollah, P. Thakur, K. Chae, B. Angadi, and W. Choi, JOURNAL OF APPLIED PHYSICS 102, 073707 (2007).
- Uchida et al. (1987) S. Uchida, K. Kitazawa, and S. Tanaka, Phase transitions 8, 95 (1987).
- Mott (1972) N. Mott, Philosophical Magazine 26, 1015 (1972).
- Mott and Davis (2012) N. F. Mott and E. A. Davis, Electronic processes in non-crystalline materials (Oxford university press, 2012).
- Rosenbaum (1991) R. Rosenbaum, Physical Review B 44, 3599 (1991).
- Sakata et al. (1999) H. Sakata, K. Sega, and B. K. Chaudhuri, Physical Review B 60, 3230 (1999), URL https://doi.org/10.1103%2Fphysrevb.60.3230.
- Kitazawa et al. (1985) K. Kitazawa, S. Uchida, and S. Tanaka, Physica B+ C 135, 505 (1985).
- Chmaissem et al. (2003) O. Chmaissem, B. Dabrowski, S. Kolesnik, J. Mais, J. Jorgensen, and S. Short, Physical Review B 67, 094431 (2003).
- Sharma et al. (2022) P. Sharma, S. Pathak, H. Pant, and R. Bindu, Applied Physics A 128, 271 (2022).
- Patterson (1939) A. Patterson, Physical review 56, 978 (1939).
- Horn et al. (1987) P. Horn, D. Keane, G. Held, J. Jordan-Sweet, D. Kaiser, F. Holtzberg, and T. Rice, Physical review letters 59, 2772 (1987).
- Telang et al. (2022) P. Telang, A. Bandyopadhyay, K. Mishra, D. Rout, R. Bag, A. Gloskovskii, Y. Matveyev, and S. Singh, Journal of Physics: Condensed Matter 34, 395601 (2022).
- Payne et al. (2007) D. J. Payne, R. G. Egdell, D. S. Law, P.-A. Glans, T. Learmonth, K. E. Smith, J. Guo, A. Walsh, and G. W. Watson, Journal of Materials Chemistry 17, 267 (2007).
- Bharath et al. (2019) M. Bharath, P. Sharma, J. Brar, R. Maurya, and R. Bindu, Journal of Physics: Condensed Matter 32, 055504 (2019).
- Plumb et al. (2016) N. Plumb, D. Gawryluk, Y. Wang, Z. Ristić, J. Park, B. Lv, Z. Wang, C. Matt, N. Xu, T. Shang, et al., Physical Review Letters 117, 037002 (2016).
- Perdew and Levy (1983) J. P. Perdew and M. Levy, Physical Review Letters 51, 1884 (1983).
- Sham and Schlüter (1983) L. J. Sham and M. Schlüter, Physical review letters 51, 1888 (1983).
- Sarma et al. (1998) D. Sarma, A. Chainani, S. Krishnakumar, E. Vescovo, C. Carbone, W. Eberhardt, O. Rader, C. Jung, C. Hellwig, W. Gudat, et al., Physical review letters 80, 4004 (1998).
- Kobayashi et al. (2007) M. Kobayashi, K. Tanaka, A. Fujimori, S. Ray, and D. Sarma, Physical review letters 98, 246401 (2007).