The recent discovery of superconductivity in hole-doped infinite-layer nickelates has triggered a great interest in the synthesis of novel nickelate phases, which have primarily been examined in thin film samples. Here, we report the high-pressure optical floating zone growth of various perovskite and perovskite-derived rare-earth nickelate single-crystals and investigate the effects of hole-, electron-, and self-doping. For hole-doping with Ca and Sr, we observe phase separations during the growth process when a substitution level of 8% is exceeded. A similar trend emerges for electron-doping with Ce and Zr. Employing lower doping levels allows us to grow sizable crystals in the perovskite phase, which exhibit significantly different electronic and magnetic properties than the undoped parent compounds, such as decreased resistivity and a suppressed magnetic response. Our insights into the doping-dependent phase formation and the resulting properties of the synthesized crystals reveal limitations and opportunities for the exploration and manipulation of electronic states in rare-earth nickelates.

The Ruddlesden–Popper (RP) nickelates Rn+1NinO3n+1 (R = rare-earth ion) with the perovskite (n = ∞) or perovskite-derived crystal structures (n ≠ ∞) are prototypical quantum materials,1 showing a plethora of electronic, orbital, and magnetic phases already without charge carrier doping.2–6 The perovskite-derived phases with n = 1–3 have been actively studied since the 1980s.7 To synthesize the perovskite phase RNiO3, high pressure is required to stabilize the Ni3+ oxidation state and the distorted crystal structure, where the octahedral tilt angles are determined by the radius of the R ion. The least distorted perovskite nickelate LaNiO3 was first synthesized in 1957,8 whereas nickelates with R other than La were realized in 1991.9 The RNiO3 compounds have been of long-standing interest due to their sharp metal-to-insulator transition10–12 and magnetic ground state with an unconventional spin spiral.13,14 Additionally, the metal-to-insulator transition is accompanied by an orthorhombic to monoclinic structural phase transition.12 The transition temperatures of these phases decrease as the size of the R ion increases. One exception is the compound with the largest anion, LaNiO3, which exhibits a rhombohedral structure and remains metallic and paramagnetic down to the lowest temperatures, provided that the oxygen content is stoichiometric.15,16

Upon hole- or electron-doping of the Ruddlesden–Popper nickelate, new electronic phases can emerge,17–22 along with potential functional properties.23,24 In RNiO3 powders, hole- and electron doping up to 10% has been achieved through the substitution of the R ion by divalent Sr/Ca and tetravalent Th/Ce ions, respectively.25–28 Typically, the powder synthesis is carried out under high external oxygen gas pressures.9 Recent studies have reported substitutions in powders as high as 40%.29 However, further investigations are needed to determine if such high substitution concentrations are homogeneously incorporated into the microstructure of the RNiO3 powder grains.

Since 2019, the field has been reinvigorated30 by the discovery of superconducting behavior in hole-doped rare-earth nickelates with an infinite-layer crystal structure.31–36 This structure can be achieved through the topotactic oxygen deintercalation of the perovskite phase.37–39 These infinite-layer nickelates with Ni1+ are nominally isoelectronic and isostructural to cuprate high-temperature superconductors with Cu2+ ions.40 Additionally, superconductivity has been observed in the undoped nickelate Nd6Ni5O12,41 which, however, can be considered a self-doped cuprate analog.42 Yet, superconductivity in the material family of nickelates has only been observed in thin film samples,29,43 even though the latest studies have indicated that it could be an intrinsic property of the bulk phase of doped infinite-layer nickelates.44 Furthermore, theoretical studies propose that superconductivity may also arise in electron-doped nickelates, such as La2.4Zr0.6Ni2O6, where La3+ is partially substituted by tetravalent Zr4+ ions.20 

Currently, a direct synthesis of the infinite-layer phase of nickelates is unfeasible due to the highly metastable Ni1+ state in square-planar NiO2 units. Similarly, the direct synthesis of other Ruddlesden–Popper derived nickelates with NiO2 planes, such as Nd6Ni5O12, has yet to be achieved. However, well-established routes exist for the topochemical removal of the apical oxygen from the parent nickelate phases, such as using H2/Ar gas or CaH2 powder as a reducing agent.37,41,45 In particular, the CaH2-assisted reduction has been successfully demonstrated not only on perovskite thin films31–36 and polycrystalline powders43,46,47 but also on single-crystals48,49 with volumes up to 1 mm3.50 

A technical breakthrough enabling the growth of large Rn+1NinO3n+1 single-crystals was the advent of the high-oxygen pressure high-temperature optical floating zone (OFZ) technique, yielding bulk single-crystals of LaNiO3,15,16,50–53 PrNiO3,54 La3Ni2O7,55 and (La,Pr)4Ni3O10.56,57 However, a problem with oxygen transport toward the center of the boule was observed in this synthesis method.16 As a result, oxygen deficient phases may form, but they can be alleviated by post-annealing under high gas pressure in autoclaves.

Overall, the technical advances in perovskite single-crystal growth, as well as the demonstrated topotactic reduction of LaNiO3 single-crystals to infinite-layer LaNiO2,50 call for the previously unexplored synthesis of doped perovskite single-crystals that provide the starting material for the reduction to the infinite-layer phase. In this work, we present the optical floating zone growth of single-crystals of doped perovskite and Ruddlesden–Popper phase nickelates under 300 bars of oxygen partial pressure and discuss the opportunities and limitations of this method. We begin with an overview of the synthesis of the undoped compounds LaNiO3 and PrNiO3, as well as an attempt at the growth of the n = 5 Ruddlesden–Popper nickelate Pr6Ni5O16. As a key result of the study, we report the phase formations in Ce- and Zr-doped (La,Pr)NiO3, which are nominally electron-doped. Furthermore, we discuss the first OFZ growth of Sr-doped LaNiO3, which is nominally hole-doped.

Precursor powders were prepared by mixing the corresponding stoichiometries of La2O3 (99.99% Alfa Aesar), Pr6O11 (Alfa Aesar, 99.99%), and NiO (99.998% Alfa Aesar), as well as CeO2 (99.99% Alfa Aesar), CaCO3 (99.999% Alfa Aesar), SrCO3 (99.998% Roth), ZrO2 (99.978% Alfa Aesar), and Eu2O3 (99.995% Roth). Subsequently, the powders were ball-milled for 20 min, and the mixtures were transferred in alumina crucibles to box furnaces, followed by heating to 1100 °C. Cylindrically shaped feed and seed rods were prepared by ball-milling the sintered materials, which were filled into rubber forms with a 6 mm diameter. The rubber was evacuated and pressed in a stainless steel form filled with water using a Riken type S1-120 70 kN press. All rods were heat treated at 900 °C to avoid cracking or breaking due to the oxidization process when the density of the rod becomes too high.

The single-crystal growth was carried out in a high pressure, high-temperature optical floating zone furnace (model HKZ, SciDre GmbH, Dresden, Germany) that allows for gas pressures in the growth chamber up to 300 bars. The growth chamber has a length of 72 mm and a 20 mm wall thickness. A xenon arc lamp operating at 5 kW was used as a heating source with the rare vertical mirror alignment of the HKZ. The rods were then aligned in the HKZ on steel holders, followed by the installation of the high pressure chamber. Subsequently, the chamber was filled up to 14/30/85/200/300 bars of oxygen pressure and held at a flow rate of 0.1 l/min. The crucial part in the nickelate growth is the initial melt connection, which is complicated as we typically start with a mixture of the Ruddlesden–Popper phase (La,Pr)2NiO4 and NiO, which both have lower melting points than the desired perovskite phase. However, preannealing at high pressures leads to cracking and even breaking of the rods. Therefore, we either increased the power until a homogeneous zone melt was achieved or premelted the rods at the given pressure to perform the final phase. In the former case, achieving a homogeneous zone melt is necessary before connecting the two rods, but the starting process of the growth was often very challenging. Both procedures have been employed for the materials synthesized in this work and have proven to be equally successful. Notably, a premelting of the rods does not increase the oxygen content in the obtained boule, as the given oxygen pressure and oxygen transport in the melt stabilize this, and the subsequent growth is easier.

All growths were conducted under an oxidizing atmosphere, with the partial oxygen pressure tailored to yield the best outcome for each composition (see Tables I and II). Common to all compounds investigated here is the increase in melting temperature when the phase containing Ni3+ is formed. This effect is amplified by the endothermal oxidation process, which we counteract with a continuous increase in lamp power, particularly during the early stage of the growth. This means that newly introduced material will be largely overheated and become very fluid. Therefore, without premelting, the diameter of the grown boule is dictated by the flow of the low-viscosity melt, which requires constant feeding. For all growths, we found that a growth rate of 2 mm/h with an additional feed rate of 2 mm/h results in stable conditions.

TABLE I.

Summary of the materials without doping grown with the HKZ. The columns indicate the desired nominal composition of the material; n labels the formed Ruddlesden–Popper (RP) phase; wt. % is the weight percentage of the RP phase; and a, b, and c are the lattice constants determined from Rietveld refinements of PXRD data. Furthermore, the space group (structure) and the oxygen partial pressure of the growth are indicated. If a wt. % below 100 is given, the remaining phase is NiO.

Nominal comp.n in RPwt. %a (Å)b (Å)c (Å)Structurep(O2) (bar)
LaNiO3 ∞ 97.9(7) 5.454 08(7) ⋯ 13.1255(2) R3̄c 85 
PrNiO3 ∞ 97.3(6) 5.409 5(1) 5.3703(1) 7.6147(2) Pbnm 300 
Pr6Ni5O16 72.5(6) 5.371 7(2) 5.4565(1) 27.534(1) P21/a 300 
Pr6Ni5O16 27.5(3) 5.440 6(3) 5.3763(3) 7.6166(4) Pbnm 300 
La3Eu3Ni5O16 74.9(7) 5.365 7(2) 5.4599(2) 27.4703(9) P21/a 300 
La3Eu3Ni5O16 11.1(3) 3.808(2) ⋯ 12.515(1) I4/mmm 300 
Nominal comp.n in RPwt. %a (Å)b (Å)c (Å)Structurep(O2) (bar)
LaNiO3 ∞ 97.9(7) 5.454 08(7) ⋯ 13.1255(2) R3̄c 85 
PrNiO3 ∞ 97.3(6) 5.409 5(1) 5.3703(1) 7.6147(2) Pbnm 300 
Pr6Ni5O16 72.5(6) 5.371 7(2) 5.4565(1) 27.534(1) P21/a 300 
Pr6Ni5O16 27.5(3) 5.440 6(3) 5.3763(3) 7.6166(4) Pbnm 300 
La3Eu3Ni5O16 74.9(7) 5.365 7(2) 5.4599(2) 27.4703(9) P21/a 300 
La3Eu3Ni5O16 11.1(3) 3.808(2) ⋯ 12.515(1) I4/mmm 300 
TABLE II.

Summary of the materials with doping grown with the HKZ. The columns indicate the desired nominal composition of the material: n labels the formed Ruddlesden–Popper (RP) phase (where DP stands for double perovskite), wt. % is the weight percentage of the RP phase, x is the substitution content determined by EDS, and a, b, and c are the lattice constants determined from Rietveld refinements of PXRD data. Furthermore, the space group (structure) and the oxygen partial pressure of the growth are indicated. If a wt. % below 100 is given, the remaining phase is NiO. The labels (in), (mid), and (out) indicate whether the powder for the PXRD analysis was extracted from the inner, middle, or outer part of the boule, as depicted in Fig. 7.

Nominal comp.n in RPwt. %x from EDSa (Å)b (Å)c (Å)Structurep(O2) (bar)
Pr0.95Ce0.05NiO3 ∞ 41.3(3) 0.02(1) 5.409 7(1) 5.3748(2) 7.620 1(3) Pbnm 300 
La0.95Ce0.05NiO3 ∞ 77.5(4) 0.046(3) 5.453 50(8) ⋯ 13.123 5(2) R3̄c 85 
La0.91Ce0.09NiO3 ∞ 85.81(5) 0.057(7) 5.453 0(1) ⋯ 13.120 1(3) R3̄c 85 
La0.8Ce0.2NiO3 ∞ 69.2(4) 0.07(1) 5.453 2(1) ⋯ 13.119 3(3) R3̄c 200 
La2.4Zr0.6Ni2O7 ∞ 46.14(4) 0.12(1) 5.501 9(2)  13.253 1(5) R3̄c 14 
La2.4Zr0.6Ni2O7 DP 52.6(5) ⋯ 5.623 1(2) 5.7053(3) 7.989 1(4) P21/n 14 
La2.4Zr0.6Ni2O7 DP 60.5(7) ⋯ 5.640 4(2) 5.7442(3) 8.020 9(4) P21/n 4.5 
La2.4Zr0.6Ni2O7 19.3(2) ⋯ 3.863 7(1) ⋯ 12.685 7(7) I4/mmm 4.5 
La0.98Zr0.02NiO3 ∞ 99.5(2) 0.026(6) 5.481 44(2) ⋯ 13.189 48(7) R3̄c 85 
La0.95Zr0.05NiO3 ∞ 99.5(2) 0.071(4) 5.481 44(2) ⋯ 13.189 48(7) R3̄c 85 
La0.8Zr0.2NiO3 ∞ 57.8(5) 0.12(1) 5.502 7(1) ⋯ 13.259 6(4) R3̄c 30 
La0.8Zr0.2NiO3 DP 39.39(4) ⋯ 5.622 3(2) 5.6987(3) 7.978 1(4) P21/n 30 
La0.92Sr0.08NiO3 ∞ 65.2(6) 0.12(5) 5.447 1(2) ⋯ 13.169 1(1) R3̄c 200 
La0.84Sr0.16NiO3 (in) ∞ 27.5(2) 0.13(1) 5.445 51(7) ⋯ 13.139 5(3) R3̄c 300 
La0.84Sr0.16NiO3 (in) 67.1(4) ⋯ 3.825 78(4) ⋯ 12.700 2(1) I4/mmm 300 
La0.84Sr0.16NiO3 (mid) ∞ 54.9(3) ⋯ 5.443 72(6) ⋯ 13.134 1(2) R3̄c 300 
La0.84Sr0.16NiO3 (mid) 31.7(2) ⋯ 3.810 85(6) ⋯ 12.714 1(9) I4/mmm 300 
La0.84Sr0.16NiO3 (out) ∞ 72.1(4) ⋯ 5.445 39(4) ⋯ 13.138 2(1) R3̄c 300 
La0.84Sr0.16NiO3 (out) 29.4(2) ⋯ 3.818 60(5) ⋯ 12.717 2(3) I4/mmm 300 
Nominal comp.n in RPwt. %x from EDSa (Å)b (Å)c (Å)Structurep(O2) (bar)
Pr0.95Ce0.05NiO3 ∞ 41.3(3) 0.02(1) 5.409 7(1) 5.3748(2) 7.620 1(3) Pbnm 300 
La0.95Ce0.05NiO3 ∞ 77.5(4) 0.046(3) 5.453 50(8) ⋯ 13.123 5(2) R3̄c 85 
La0.91Ce0.09NiO3 ∞ 85.81(5) 0.057(7) 5.453 0(1) ⋯ 13.120 1(3) R3̄c 85 
La0.8Ce0.2NiO3 ∞ 69.2(4) 0.07(1) 5.453 2(1) ⋯ 13.119 3(3) R3̄c 200 
La2.4Zr0.6Ni2O7 ∞ 46.14(4) 0.12(1) 5.501 9(2)  13.253 1(5) R3̄c 14 
La2.4Zr0.6Ni2O7 DP 52.6(5) ⋯ 5.623 1(2) 5.7053(3) 7.989 1(4) P21/n 14 
La2.4Zr0.6Ni2O7 DP 60.5(7) ⋯ 5.640 4(2) 5.7442(3) 8.020 9(4) P21/n 4.5 
La2.4Zr0.6Ni2O7 19.3(2) ⋯ 3.863 7(1) ⋯ 12.685 7(7) I4/mmm 4.5 
La0.98Zr0.02NiO3 ∞ 99.5(2) 0.026(6) 5.481 44(2) ⋯ 13.189 48(7) R3̄c 85 
La0.95Zr0.05NiO3 ∞ 99.5(2) 0.071(4) 5.481 44(2) ⋯ 13.189 48(7) R3̄c 85 
La0.8Zr0.2NiO3 ∞ 57.8(5) 0.12(1) 5.502 7(1) ⋯ 13.259 6(4) R3̄c 30 
La0.8Zr0.2NiO3 DP 39.39(4) ⋯ 5.622 3(2) 5.6987(3) 7.978 1(4) P21/n 30 
La0.92Sr0.08NiO3 ∞ 65.2(6) 0.12(5) 5.447 1(2) ⋯ 13.169 1(1) R3̄c 200 
La0.84Sr0.16NiO3 (in) ∞ 27.5(2) 0.13(1) 5.445 51(7) ⋯ 13.139 5(3) R3̄c 300 
La0.84Sr0.16NiO3 (in) 67.1(4) ⋯ 3.825 78(4) ⋯ 12.700 2(1) I4/mmm 300 
La0.84Sr0.16NiO3 (mid) ∞ 54.9(3) ⋯ 5.443 72(6) ⋯ 13.134 1(2) R3̄c 300 
La0.84Sr0.16NiO3 (mid) 31.7(2) ⋯ 3.810 85(6) ⋯ 12.714 1(9) I4/mmm 300 
La0.84Sr0.16NiO3 (out) ∞ 72.1(4) ⋯ 5.445 39(4) ⋯ 13.138 2(1) R3̄c 300 
La0.84Sr0.16NiO3 (out) 29.4(2) ⋯ 3.818 60(5) ⋯ 12.717 2(3) I4/mmm 300 

Powder x-ray diffraction (PXRD) was performed at room temperature using a Rigaku MiniFlex diffractometer with Bragg Brentano geometry, Cu Kα radiation, and a Ni filter. Rietveld refinements were conducted with the FullProf software suite.58 The x-ray Laue diffraction images were collected with a Photonic Science CCD detector using a standard W broad x-ray source operated at 35 kV and 40 mA. For indexing the Laue patterns, the software OrientExpress was used.

Electron microscopy images with both secondary electrons (SE) and backscattering electrons (BSE) were taken with a Zeiss Merlin electron microscope operated at 12 kV, 600 mA, at a sample distance of 5 mm. Energy-dispersive x-ray spectra (EDS) were recorded with a NORAN System 7 (NSS212E) detector in a Tescan Vega (TS-5130MM) SEM.

Magnetic susceptibility measurements were performed using a vibrating sample magnetometer [MPMS VSM superconducting quantum interference device (SQUID), Quantum Design] and electrical transport measurements with a Physical Property Measurement System (PPMS, Quantum Design).

X-ray photoelectron spectroscopy (XPS) data were collected using a commercial Kratos AXIS Ultra spectrometer and a monochromatized Al Kα source (photon energy, 1486.6 eV). The base pressure during XPS was in the low 10−10 mbar range. The spectra were collected using an analyzer with a pass energy of 20 eV. XPS spectra were analyzed using the CASAXPS software.59 All samples were cleaved in a glove box, mounted on carbon tape or In foil, and transported under an inert atmosphere to the XPS chamber.

Figure 1(a) shows the as-grown LaNiO3 boule together with the characterization by PXRD and a Laue diffraction image [Fig. 1(e)]. The OFZ synthesis with an HKZ-type furnace of LaNiO3 as well as PrNiO3, La3Ni2O7, and (La,Pr)4Ni3O10 has been reported previously,15,16,50–57 although only a few details were given about the growth, phase diagram, and formation of secondary phases. As can be seen in Fig. 1(a), the PXRD reveals the presence of a small amount of NiO in addition to the primary LaNiO3 phase in a pulverized piece broken off from the boule. The emergence of secondary phases such as NiO will be discussed in detail for the other nickelate compounds in this study. In accordance with previous studies on LaNiO3, we refine the PXRD data in the rhombohedral space group R3̄c, which also describes the structure of the doped La-based perovskites in our study (see Tables I and II). As apparent from Laue diffraction [Fig. 1(e)], the growth direction of LaNiO3 corresponds to the rhombohedral (110) direction. Under the application of a relatively strong force, the LaNiO3 boule can be cleaved. The exposed surfaces are rough with small facets [Figs. 1(c) and 1(d)] that are mostly aligned in parallel to the (110) plane. The boule can also be cleaved along the orthogonal direction, which is displayed in Fig. 1(d), where the (11̄0) direction points out of the picture while the c-axis lies in the horizontal plane [see also the corresponding sketch of the crystal structure in Fig. 1(b)].

FIG. 1.

(a) Image of the grown LaNiO3 boule and the PXRD pattern. The solid black line corresponds to the calculated intensity from the Rietveld refinement. The calculated Bragg peak positions of the LaNiO3 majority [97.9(7) wt. %] and NiO minority phase [2.1(2) wt. %] are indicated as green and red vertical bars, respectively. The blue line corresponds to the difference between the experimental and calculated intensity. (b) Schematic of the rhombohedral unit cell (hexagonal axes) of perovskite LaNiO3, viewed along the same direction as the cleave shown in panel (d). (c) Picture of the cross section of the boule obtained after cleaving a single crystal. (d) Picture of a cleave along the growth direction of a single crystalline piece with indexed orientations determined from the Laue diffraction pattern in panel (e).

FIG. 1.

(a) Image of the grown LaNiO3 boule and the PXRD pattern. The solid black line corresponds to the calculated intensity from the Rietveld refinement. The calculated Bragg peak positions of the LaNiO3 majority [97.9(7) wt. %] and NiO minority phase [2.1(2) wt. %] are indicated as green and red vertical bars, respectively. The blue line corresponds to the difference between the experimental and calculated intensity. (b) Schematic of the rhombohedral unit cell (hexagonal axes) of perovskite LaNiO3, viewed along the same direction as the cleave shown in panel (d). (c) Picture of the cross section of the boule obtained after cleaving a single crystal. (d) Picture of a cleave along the growth direction of a single crystalline piece with indexed orientations determined from the Laue diffraction pattern in panel (e).

Close modal

Figures 2(a) and 2(b) depict the boules and the PXRD patterns of our PrNiO3 growth and the attempted growth of Pr6Ni5O16. To achieve a boule of PrNiO3 with a length of about 8 cm [see Fig. 2(a)], a continuous pressure of 300 bars was held for several days. The results of our PXRD analysis of PrNiO3 and our other growths are summarized in Table I, which displays the lattice constants and phase compositions in wt. %. In agreement with previous studies, we refine the structure of PrNiO3 in the orthorhombic space group Pbnm. In Fig. 2(c), we present the first phase diagram of NiO–PrO2, which at the given pressure of 300 bars of oxygen only contains four phases: NiO, PrNiO3, Pr4Ni3O10, and PrO2.

FIG. 2.

(a) Image of the grown PrNiO3 boule and the PXRD pattern. The calculated Bragg peak positions of the PrNiO3 majority [97.3(7) wt. %] and NiO minority phase [2.7(2) wt. %] are indicated as green and purple vertical bars, respectively. (b) Image of the boule of the attempted Pr6Ni5O16 growth and the PXRD pattern, which indicates only the Pr4Ni3O10 [72.5(6) wt. %] and the PrNiO3 phase [27.5(3) wt. %]. (c) Sketch of the growth phase diagram for the phase formations of the OFZ melt under 300 bars of oxygen partial pressure for NiO and PrO2. (d) and (e) Simultaneously acquired SE (left) and BSE (right) images of broken surfaces from the boules grown from a precursor corresponding to nominally stoichiometric PrNiO3 (d) and a Pr-excess precursor nominally corresponding to Pr1.05NiO3 (e). The inclusions with a dark contrast in (d) correspond to NiO, while regions with a light contrast in (e) correspond to intergrown Pr4Ni3O10.

FIG. 2.

(a) Image of the grown PrNiO3 boule and the PXRD pattern. The calculated Bragg peak positions of the PrNiO3 majority [97.3(7) wt. %] and NiO minority phase [2.7(2) wt. %] are indicated as green and purple vertical bars, respectively. (b) Image of the boule of the attempted Pr6Ni5O16 growth and the PXRD pattern, which indicates only the Pr4Ni3O10 [72.5(6) wt. %] and the PrNiO3 phase [27.5(3) wt. %]. (c) Sketch of the growth phase diagram for the phase formations of the OFZ melt under 300 bars of oxygen partial pressure for NiO and PrO2. (d) and (e) Simultaneously acquired SE (left) and BSE (right) images of broken surfaces from the boules grown from a precursor corresponding to nominally stoichiometric PrNiO3 (d) and a Pr-excess precursor nominally corresponding to Pr1.05NiO3 (e). The inclusions with a dark contrast in (d) correspond to NiO, while regions with a light contrast in (e) correspond to intergrown Pr4Ni3O10.

Close modal

Similarly to LaNiO3 [Fig. 1(b)], we detect a minor amount of NiO in the stoichiometrically grown PrNiO3 [Fig. 2(b)]. Note that an admixture of NiO to the perovskite phase can become relevant after a topotactic reduction where ferromagnetic Ni forms, which can dominate the signal in magnetic susceptibility measurements.47 In cases of very small NiO admixtures, their presence can be below the detection limit of standard laboratory PXRD. A method that allows us to detect even subtle amounts of NiO inclusions is BSE imaging. In particular, unlike SE images, which reflect the sample topography, BSE images reveal the spatial distribution of the elements. Figure 2(d) shows the simultaneously acquired SE and BSE images of a broken surface from the PrNiO3 boule. In the latter image, NiO rich regions are clearly observed in the dark contrast.

In principle, the emergence of NiO rich regions should be alleviated by growth from a precursor with a Pr excess. However, we find that when varying the stoichiometry to a subtle Pr rich content, Pr4Ni3O10 forms an intergrown phase, especially in regions in proximity to the surface of the boule [see light contrast in the right panel of Fig. 2(e)]. This occurs because impurity phases are usually pushed to the surface in an OFZ growth as nucleation occurs in the core of the boule. Due to our external heating source, the heat is only indirectly transported to the center resulting in a small gradient. Nonetheless, this observation is in contrast to the pressure stability, as the center also sees the least pressure, and Pr4Ni3O10 is more stable at lower pressures.

For PrNiO3, the growth direction corresponds to the orthorhombic (100) direction. Similarly to LaNiO3, under the application of a relatively strong force, the PrNiO3 boule can be cleaved. The exposed surfaces are rough with small facets that are mostly aligned in parallel to the (100) plane. The boule can also be cleaved along the orthogonal direction, where the (010) direction points out of the plane and the (001) direction is the horizontal one. Notably, for both La- and Pr-compounds, these cleaving patterns emerge even for the doped cases that will be discussed below.

Next, we turn to the synthesis of the n = 5 Ruddlesden–Popper phase Pr6Ni5O16 under 300 bars of oxygen partial pressure. If successful, the compound could provide a perspective for topotactic reductions to the Pr6Ni5O12 phase, in analogy to superconducting Nd6Ni5O12 films,41 and seems ideally suited for the high-pressure OFZ growth, as the pressure stability range should fall into the region accessible with the HKZ. However, instead of Pr6Ni5O16, our PXRD characterization reveals a phase mixture of the perovskite and the n = 3 phase Pr4Ni3O10 [Fig. 2(b)], which are the phases that encompass the n = 5 compound in the growth phase diagram. Similarly, we find in an attempted synthesis (not shown here) that the n = 5 variant La3Eu3Ni5O16 cannot be stabilized under 300 bars of oxygen (see Table I), despite the different ionic radii of La and Eu. This is distinct from other layered materials such as cuprates, where the mixing of ionic radii of the smaller rare earth ion with the large Ba facilitates the stabilization of complex layered structures. Here, we find that instead of n = 5, the n = 1 and 3 Ruddlesden–Popper phases form. Hence, we conclude that at a given pressure of 300 bars, the synthesis of La- and Pr-based Ruddlesden–Popper phases higher than n = 3 is likely unfeasible, as the existence of these phases is probably restricted to a very narrow range in pressure and composition space.

Electron-doping of PrNiO3 can be achieved through the substitution of the trivalent Pr ions by nominally tetravalent Ce ions. Here we carry out the OFZ growth at 280 and 300 bars of oxygen partial pressure, respectively, aiming for 5% Ce doping of the perovskite phase. The upper panel in Fig. 3(a) shows the as-grown boules for 280 (top) and 300 bars (bottom). In comparison to the growth of the undoped compounds, a more homogeneous and viscous liquid forms during this growth, facilitating the growth of a long and homogeneous boule. In our characterization, the boules grown under 280 and 300 bars show similar phase formations, and hence we only focus on the latter in the following. Already for the small amount of 5% doping, a phase mixture between Ce-substituted PrNiO3, (Pr,Ce)O2, and NiO forms [see PXRD in Fig. 3(a)], presumably because the solubility limit of Ce is very low. The phase formation in different regions of the boule is best seen in BSE imaging, with Fig. 3(b) giving a wide-scale overview of the cross section of the boule along the growth direction.

FIG. 3.

(a) Images of the boules grown at 280 and 300 bars and the PXRD pattern from a growth of the nominal composition Pr0.95Ce0.05NiO3. The calculated Bragg peak positions of the Pr1−xCexNiO3 majority [41.3(3) wt. %] and NiO [27.7(4) wt. %] and (Pr,Ce)O2 [31.0(2) wt. %] minority phases are indicated as orange, purple, and green vertical bars, respectively. (b) BSE image of a broken surface from the grown boule, with the growth direction running from bottom left to top right. The blue arrow indicates lines of eutectic freezing, which form when the solubility limit is reached. The regions with darker and lighter contrast correspond to NiO and (Pr,Ce)O2, respectively. (c) Phase diagram for the phase formations of the OFZ melt under 300 bars of oxygen partial pressure for (Pr,Ce)O2 and NiO. (d) Magnified BSE image of the same broken surface as in (b), but zoomed into an eutectic separation region.

FIG. 3.

(a) Images of the boules grown at 280 and 300 bars and the PXRD pattern from a growth of the nominal composition Pr0.95Ce0.05NiO3. The calculated Bragg peak positions of the Pr1−xCexNiO3 majority [41.3(3) wt. %] and NiO [27.7(4) wt. %] and (Pr,Ce)O2 [31.0(2) wt. %] minority phases are indicated as orange, purple, and green vertical bars, respectively. (b) BSE image of a broken surface from the grown boule, with the growth direction running from bottom left to top right. The blue arrow indicates lines of eutectic freezing, which form when the solubility limit is reached. The regions with darker and lighter contrast correspond to NiO and (Pr,Ce)O2, respectively. (c) Phase diagram for the phase formations of the OFZ melt under 300 bars of oxygen partial pressure for (Pr,Ce)O2 and NiO. (d) Magnified BSE image of the same broken surface as in (b), but zoomed into an eutectic separation region.

Close modal

In OFZ growth, we are constantly feeding the weighed-in stoichiometry; therefore, when a solubility limit in growth is reached, the seed crystal incorporates too little of the corresponding element. Hence, this accumulates in the growth until an eutectic point [the minimum in the phase diagram, e.g., 0.2 and 0.75 shown in Fig. 3(c)] in the phase diagram is reached. Here, the eutectic part freezes out and forms a layer of the eutectic stoichiometry (which contains a major part of the impurity phase). This eutectic can be seen in Fig. 3(b), which shows a large cross section of the boule via BSE with elemental resolution. Here, black lines are visible, revealing these eutectic rings. Figure 3(d) shows a magnified version of these eutectic mixes, revealing three phases intergrown with each other. Note that for all subsequent BSE images in the other subchapters, we only show the magnified versions of the eutectic parts in our boule, and XRD focuses on these sections as well, but in all cases, similar eutectic rings are formed and completely homogeneous parts exist as well, similar to what is visible in Fig. 3(b) on the top right.

Nonetheless, the matrix is a perovskite phase as the Laue diffraction images from broken surfaces of the boule (not shown here) can be indexed in space group Pbnm, which suggests that large grains of Ce-substituted PrNiO3 have crystallized in our boule, although BSE imaging reveals that the other two phases are intergrown in these grains [Figs. 3(b) and 3(d)]. Importantly, we do not detect any traces of higher order Ruddlesden–Popper phases, such as Pr4Ni3O10, which is in contrast to the Pr-excess growth of the undoped compound shown in Fig. 2(e). Instead, we already observe the formation of (Pr,Ce)O2, which is an end member of the growth phase diagram. Hence, we propose the modified growth phase diagram shown in Fig. 3(c) for a Ce-mixed pseudo binary composition.

Next, we focus on the electron-doping of LaNiO3. As a first attempt, we attempted to synthesize crystals of composition La0.8Ce0.2NiO3, which according to prior work on infinite-layer thin films is optimal for superconductivity, but observed that the growth was not very stable, as can be seen in the image of the grown boule in Fig. 4(a). Large crystals could not be extracted from the growth, and we find a low wt. % of the perovskite phase (see Table II). Therefore, we go for lower substitutions and obtain large single-crystals with millimeter dimensions [as shown in Fig. 4(a)] for the nominal compositions La0.95Ce0.05NiO3 and La0.91Ce0.09NiO3. For the La-based compounds, we observe that already small Ce-substitution levels lead to the emergence of a competing phase, which in this case is the pyrochlore phase La2Ce2O7. The presence of this secondary phase is detected in the PXRD patterns [Figs. 4(b) and 4(c)] and the BSE image [Fig. 4(e)], revealing that La2Ce2O7 is embedded in the single crystalline perovskite matrix as eutectic rings together with NiO. Hence, the Ce-substitution of the perovskite phase is challenging, although small crystals of (La,Pr)1−xCexNiO3 surrounded by regions of secondary phases can be realized in the boule.

FIG. 4.

Images of the boules and the PXRD patterns from growths of the nominal compositions (a) La0.8Ce0.2NiO3, (b) La0.95Ce0.05NiO3, and (c) La0.91Ce0.09NiO3. The calculated Bragg peak positions of the La1−xCexNiO3 majority [85.8(5) wt. %] and NiO [6.2(2) wt. %] and La2Ce2NiO7 [8.00(7) wt. %] minority phases are indicated as green, red, and blue vertical bars, respectively. (d) Image of a cleaved single crystal of La0.95Ce0.05NiO3, which was utilized in the XPS study. The growth direction runs from the bottom to the top. (e) BSE image of a surface of the La0.91Ce0.09NiO3 boule cleaved perpendicular to the growth direction. Regions with darker and lighter contrast correspond to NiO and La2Ce2NiO7, respectively.

FIG. 4.

Images of the boules and the PXRD patterns from growths of the nominal compositions (a) La0.8Ce0.2NiO3, (b) La0.95Ce0.05NiO3, and (c) La0.91Ce0.09NiO3. The calculated Bragg peak positions of the La1−xCexNiO3 majority [85.8(5) wt. %] and NiO [6.2(2) wt. %] and La2Ce2NiO7 [8.00(7) wt. %] minority phases are indicated as green, red, and blue vertical bars, respectively. (d) Image of a cleaved single crystal of La0.95Ce0.05NiO3, which was utilized in the XPS study. The growth direction runs from the bottom to the top. (e) BSE image of a surface of the La0.91Ce0.09NiO3 boule cleaved perpendicular to the growth direction. Regions with darker and lighter contrast correspond to NiO and La2Ce2NiO7, respectively.

Close modal

A different route to achieve electron-doping is the substitution of tetravalent Zr ions. Such doping was recently proposed theoretically for the topotactically reduced n = 2 compound La2.4Zr0.6Ni2O6,20 where superconducting behavior is expected according to the nominal valence of Ni. As a first step, we attempt to synthesize this material at a moderate pressure of 14 bars of oxygen. While such pressure was sufficient to synthesize the non-substituted n = 2 compound La3Ni2O7,55 we find for the Zr-doped variant that a phase mixture between the perovskite and the double-perovskite phase La2NiZrO6 forms (see Table II). As the n = ∞ phase is typically stable starting with 30 bars,52 we reduce the pressure and attempt a synthesis at 2.5–4.5 bars (see Table II). In this case, however, we obtain a mixture of the n = 0 and n = ∞ phases, finally proving the destabilization of higher order Ruddlesden–Popper phases with electron doping.

As a next step, we focus on the synthesis of the perovskite phase and test the maximum possible Zr substitution content. To this end, we prepare a “gradient rod” that is composed of sections with different substitution levels of Zr. The prepared sections correspond to 5%, 8.75%, 12.5%, 16.75%, and 20% Zr [Fig. 5(a)]. During the growth, we observe that all substitutions help stabilize the melt, while secondary phases emerge for 8.75% and higher substitutions. Notably, the growth conditions are stable even in the case of evolving phase mixtures. However, no large single-crystals form in these attempts due to the shrinking sizes of the perovskite domains, as evidenced in the BSE images by the decrease of a single contrast matrix [see Figs. 5(f) and 5(g)]. A rudimentary proposal for the pseudo-binary phase diagram is shown in Fig. 5(h), which notably only provides points in the center, where our PXRD characterization [exemplary PXRD patterns in Figs. 5(b) and 5(c)] reveals the formation of the pure perovskite and perovskite double-perovskite mixtures at higher doping levels, respectively.

FIG. 5.

(a) Image of the boule with a nominal La1−xZrxNiO3 composition, where the white lines separate sections with nominal Zr-substitutions of 5%, 8.75%, 12.5%, 16.75%, and 20% (from bottom to top), respectively. (b) and (c) PXRD patterns of pulverized pieces from the sections with the nominal La0.95Zr0.05NiO3 (b) and La0.8Zr0.2NiO3 (c) compositions. The calculated Bragg positions of the La1−xZrxNiO3 majority and NiO minority phases are indicated by purple and green vertical bars, respectively. In (b) and (c), the majority phase contributes 99.5(2) and 57.8(5) wt. %, respectively, while the NiO minority phase corresponds to 0.5(2) and 2.8(2) wt. %, respectively. In addition, the La2ZrNiO6 double-perovskite phase is present [39.4(4) wt. %] in the PXRD in (c), as indicated by purple vertical bars. (d)–(g) SE (left) and BSE (right) images of broken surfaces from the sections with nominal Zr-substitutions of 5% (d), 8.75% (e), 12.5% (f), and 16.75% (g). The inclusions with a dark contrast in the BSE images correspond to NiO or dirt on the surface, while a lighter contrast (indicated by the arrows) in (f) and (g) corresponds to La2ZrNiO6 (highlighted by blue arrows). (h) Pseudo binary phase diagram of (Zr,La)Ox and NiO at 85 bars of oxygen partial pressure.

FIG. 5.

(a) Image of the boule with a nominal La1−xZrxNiO3 composition, where the white lines separate sections with nominal Zr-substitutions of 5%, 8.75%, 12.5%, 16.75%, and 20% (from bottom to top), respectively. (b) and (c) PXRD patterns of pulverized pieces from the sections with the nominal La0.95Zr0.05NiO3 (b) and La0.8Zr0.2NiO3 (c) compositions. The calculated Bragg positions of the La1−xZrxNiO3 majority and NiO minority phases are indicated by purple and green vertical bars, respectively. In (b) and (c), the majority phase contributes 99.5(2) and 57.8(5) wt. %, respectively, while the NiO minority phase corresponds to 0.5(2) and 2.8(2) wt. %, respectively. In addition, the La2ZrNiO6 double-perovskite phase is present [39.4(4) wt. %] in the PXRD in (c), as indicated by purple vertical bars. (d)–(g) SE (left) and BSE (right) images of broken surfaces from the sections with nominal Zr-substitutions of 5% (d), 8.75% (e), 12.5% (f), and 16.75% (g). The inclusions with a dark contrast in the BSE images correspond to NiO or dirt on the surface, while a lighter contrast (indicated by the arrows) in (f) and (g) corresponds to La2ZrNiO6 (highlighted by blue arrows). (h) Pseudo binary phase diagram of (Zr,La)Ox and NiO at 85 bars of oxygen partial pressure.

Close modal

Additionally, we explore the phase formation for the growth of La0.8Zr0.2NiO3 at various oxygen partial pressures ranging from 15, 25, 35 to 150 bars (not shown here). We find that for stable growth of the perovskite phase, at least 35 bars of pressure are required. Above this pressure, however, we do not detect significant changes in the phase formation or the incorporated Zr-substitution content. These results indicate that the solubility limit of Zr in nickelates lies around 8%, which is also confirmed by a detailed EDS analysis (see  Appendix B), where even for crystals from a growth with a nominally higher Zr substitution, the detected content in the perovskite matrix does not exceed 8%. The formed double-perovskite phase can be considered an impurity phase where Zr substitutes half of the Ni site instead of the desired substitution of the La sites. Hence, one might suspect that Zr generally substitutes for the Ni site, even for low Zr admixtures. However, our Rietveld refinement of the perovskite phase reveals that in this phase, Zr solely occupies the La site. Moreover, we observe a solubility limit of ∼8% for substitutions with elements other than Zr (see below).

Motivated by these insights, we next carried out the growth of crystals with a much lower Zr substitution content and optimized growth parameters. Pictures of the corresponding La0.98Zr0.02NiO3 and La0.95Zr0.05NiO3 boules are shown in Fig. 6. From the boules, we extract large and high-quality perovskite crystals, as confirmed by PXRD [Figs. 6(a) and 6(b)] and BSE imaging [Fig. 6(d)] as well as EDS analysis (see  Appendix B).

FIG. 6.

(a) and (b) Images of the boules and the PXRD patterns from the growths of the nominal compositions La0.98Zr0.02NiO3 (a) and La0.95Zr0.05NiO3 (b). The calculated Bragg peak positions of the majority La1−xZrxNiO3 [97.5(4) and 99.5(6) wt. %, respectively] and minority NiO phases [3.5(5) and 0.5(2) wt. %, respectively] are indicated by green and brown vertical bars, respectively. (c) Image of a cleaved single crystal from the La0.95Zr0.05NiO3 boule. (d) SE (left) and BSE (right) images of the polished surface of the La0.95Zr0.05NiO3 crystal analyzed in Fig. 9. The dark spots correspond to NiO.

FIG. 6.

(a) and (b) Images of the boules and the PXRD patterns from the growths of the nominal compositions La0.98Zr0.02NiO3 (a) and La0.95Zr0.05NiO3 (b). The calculated Bragg peak positions of the majority La1−xZrxNiO3 [97.5(4) and 99.5(6) wt. %, respectively] and minority NiO phases [3.5(5) and 0.5(2) wt. %, respectively] are indicated by green and brown vertical bars, respectively. (c) Image of a cleaved single crystal from the La0.95Zr0.05NiO3 boule. (d) SE (left) and BSE (right) images of the polished surface of the La0.95Zr0.05NiO3 crystal analyzed in Fig. 9. The dark spots correspond to NiO.

Close modal

Hole-doping of perovskite nickelates can be achieved through substitution with divalent ions such as Ca and Sr. In a previous high-pressure flux growth, the synthesis of small Ca-substituted LaNiO3 single-crystals was achieved,48 with 8% Ca content determined by XRD refinement, whereas EDS indicated a higher level of 12% in cleaved pieces. In an attempt to realize larger Ca-substituted crystals with the floating zone technique under oxygen pressures ranging from 200 to 300 bars, we perform growths with nominal compositions of La0.8Ca0.2NiO3 and Pr0.8Ca0.2NiO3. However, we find that only the n = 1 Ruddlesden–Popper phases La1.6Ca0.4NiO4 and Pr1.6Ca0.4NiO4 as well as NiO form at such high substitution levels (not shown here). Since our floating zone growth with Zr revealed a solubility limit around 8% [Fig. 5(e)], we next attempt the growth of La0.92Sr0.08NiO3. However, we find that connecting the corresponding feed and seed rods at 300 bars of pressure is unfeasible due to heavy oxidization and freezing of the melt. To circumvent these issues, we start the growth at a lower pressure, and by sequentially increasing it up to 300 bars, stable growth is achieved. The results of the growth are presented in Table II.

To confirm the anticipated solubility limit of Sr of 8%, we subsequently attempted the growth of La0.84Sr0.16NiO3. Unlike in the 8% case, a direct connection at 300 bars of pressure is possible, highlighting that an increase in pressure is necessary to stabilize the higher nominal Ni3.16+ oxidation state compared to Ni3.08+ in La0.92Sr0.08NiO3. Figure 7(a) shows the PXRD of pulverized pieces broken off from different parts of the boule. As expected for the growth of hole-doped compounds where extreme oxygen pressures are required, a concentration gradient is observed toward the center of the boule [see inset in Fig. 7(a)]. In particular, we find that the fraction of the desired perovskite phase decreases toward the center of the boule, which is in accord with an increasing oxygen deficiency reported in previous floating zone growths of LaNiO3.16 The wt. % and lattice constants from three different regions of the boule are summarized in Table II.

FIG. 7.

Image of the cross section of the boule (inset) and PXRD patterns from a growth of the nominal composition La0.84Sr0.16NiO3. The three PXRD patterns are taken from pieces from different positions in the boule, as indicated by the blue arrows. The calculated Bragg peak positions (vertical bars) and the difference between the experimental and calculated (blue line) are shown for the bottom PXRD pattern. The corresponding wt. % of the phases are summarized in Table II. (b) and (c) BSE images from different parts of the cross section reveal a perovskite matrix with NiO (dark) and (La,Sr)2NiO4 (light) eutectic inclusions.

FIG. 7.

Image of the cross section of the boule (inset) and PXRD patterns from a growth of the nominal composition La0.84Sr0.16NiO3. The three PXRD patterns are taken from pieces from different positions in the boule, as indicated by the blue arrows. The calculated Bragg peak positions (vertical bars) and the difference between the experimental and calculated (blue line) are shown for the bottom PXRD pattern. The corresponding wt. % of the phases are summarized in Table II. (b) and (c) BSE images from different parts of the cross section reveal a perovskite matrix with NiO (dark) and (La,Sr)2NiO4 (light) eutectic inclusions.

Close modal

As a secondary phase in the PXRD, we detect the n = 1 Ruddlesden–Popper phase, which is also evidenced in the BSE imaging as a light color [Figs. 7(b) and 7(c)]. As can be seen especially in Figs. 7(b) and 7(c), in the center of the boule, the grains of the perovskite phase are immersed in a matrix of the n = 1 Ruddlesden–Popper phase and NiO [Fig. 7(b)], whereas the situation is reversed in the outer regions, where the Ruddlesden–Popper phase and NiO form small inclusions in the perovskite phase [Fig. 7(c)]. Such intergrowth behavior appears to be a general characteristic for all investigated substitutions above 8% and points toward the existence of a general solubility limit in perovskite nickelates independent of the dopant species. In contrast, substitution levels far above 8% have been reported for thin film perovskite nickelates,33 which might be facilitated by the epitaxial strain provided by the substrate.

Nevertheless, our EDS analysis (see  Appendix B) performed locally on grains with the perovskite phase indicates that the Zr and Sr contents can reach values between 12% and 13% (see Table II), which is comparable to the highest value detected by EDS in high-pressure flux grown Ca-substituted LaNiO3 single-crystals.48 These values are substantially higher (8%), but it is possible that the EDS analysis includes a systematic overestimation of these dopants. In contrast, for the rare-earth element Ce, where the EDS analysis is based on the M instead of the K-line, we find a maximum of 7(1)% substitution (Table II), which is compatible with a solubility limit of 8%. In conclusion, complementary studies with analysis methods other than EDS are desirable to determine the maximum substitution content of Zr and Sr realizable in optical floating zone grown perovskite nickelates.

Recent investigations into the effects of hole- and electron-doping on perovskite nickelates have primarily utilized thin film samples.22,24 In electron-doped Nd1−xCexNiO3 films, a suppression of the metal-to-insulator transition of NdNiO3 was reported, favoring a shift toward a metal-to-metal transition for Ce doping levels as subtle as 2.5%.24 In the following, we explore similar effects in our (Pr,Ce)NiO3 crystals.

Figure 8 presents a comparative analysis of the electronic and magnetic properties of our undoped and Ce-doped PrNiO3 crystals. Note that the crystals have been cut into oriented pieces and annealed in a high gas pressure autoclave at 600 °C under an oxygen partial pressure of 600 bars for 5 days, followed by a rapid cooling down to room temperature. This methodical approach ensures the exclusion of the influence of oxygen deficiency on the measured properties. Figure 8(a) displays the expected sharp metal-to-insulator transition over several orders of magnitude in the resistivity in the undoped PrNiO3 single-crystal around 128 K, consistent with previous literature.54 For our single crystals with a nominal composition of Pr0.95Ce0.05NiO3, EDS indicates a lower realized Ce substitution level of around 2(1)%. We observe a metal-to-metal transition around 116 K in Fig. 8(b). This change from a metal-to-insulator to a metal-to-metal transition is qualitatively similar to that in Nd1−xCexNiO3 films.24 

FIG. 8.

(a) and (b) Resistivity as a function of temperature for a PrNiO3 (a) and a Pr0.95Ce0.05NiO3 crystal (b). The current direction coincides with the c-axis direction. The upper inset in (b) shows the cut Pr0.95Ce0.05NiO3 crystal, and the lower inset shows the specific heat measured upon cooling. (c) and (d) Magnetization as a function of temperature for the PrNiO3 (c) and Pr0.95Ce0.05NiO3 (d) crystals. The insets display a magnification of the hysteresis loops observed around the magnetic transitions.

FIG. 8.

(a) and (b) Resistivity as a function of temperature for a PrNiO3 (a) and a Pr0.95Ce0.05NiO3 crystal (b). The current direction coincides with the c-axis direction. The upper inset in (b) shows the cut Pr0.95Ce0.05NiO3 crystal, and the lower inset shows the specific heat measured upon cooling. (c) and (d) Magnetization as a function of temperature for the PrNiO3 (c) and Pr0.95Ce0.05NiO3 (d) crystals. The insets display a magnification of the hysteresis loops observed around the magnetic transitions.

Close modal

In the magnetic susceptibility, anomalies with hysteretic behavior occur at similar temperatures as the metal-to-insulator and metal-to-metal transitions in PrNiO3 and (Pr,Ce)NiO3, respectively [Figs. 8(c) and 8(d)]. These anomalies indicate the presence of an antiferromagnetic transition.54 Above the transition, the susceptibility can be well described by a Curie-Weiß law χ = CTθW, which yields the Curie constant C = 2.20(6) emu K/mol, Oe−1 = 27.6(7) · 10−6 m3 K/mol, and θW = −91(1) K for (Pr,Ce)NiO3. This is well comparable to undoped PrNiO3 with a Curie constant C = 2.38(1) emu K/mol, Oe−1 = 29.9(1) · 10−6 m3 K/mol, and θW = −99(1) K. We note that the magnetic signal of our (Pr,Ce)NiO3 sample contains a contribution from the intergrown (Pr,Ce)O2 phase, which exhibits an antiferromagnetic transition at 12 K whose signature can be seen both in the resistivity and susceptibility [Figs. 8(b) and 8(d)].

The susceptibility curves in Figs. 8(c) and 8(d) are expected to be dominated by the paramagnetic signal of the Pr3+ cations. Nevertheless, the effective moment is extracted using μeff=3kC/NμB2=2.828C. Assuming the effective paramagnetic moment of non-disproportionated Ni3+ LS (μeff=1.73μB=3kC/NμB2), CNi3+ is estimated to be 0.375 emu K/mol. Subsequently, if we consider C = CNi + CPr, the μeff of Pr is calculated to be 3.82 μB, which is only slightly larger than the expected moment of 3.5 μB and the reported one for PrNiO3.60 

Next, we characterize the physical properties of a selection of La-based compounds. To this end, we analyze large crystals of LaNiO3, La0.95Zr0.05NiO3, and La0.91Ce0.09NiO3 by Laue diffraction and cut them into rectangular pieces with the (11̄0) direction pointing out of the plane, while the c-axis and (110) direction lie in the horizontal plane [Fig. 9(a)]. Subsequently, the surfaces of the shaped and oriented crystals are polished using a 1 μm grain size. The crystals are then annealed in a high gas pressure autoclave at 600 °C under an oxygen partial pressure of 600 bars for 5 days, followed by quick cooling to room temperature. We find that oxygen annealing is required to alleviate the oxygen deficiency of the as-grown crystals, which induces antiferromagnetic order.32, Figure 9(b) shows the resistivity of the annealed crystals measured along the c axis. The high quality of our single-crystals is reflected in the large residual resistivity ratio (RRR) of LaNiO3, with a value of 19, exceeding previously reported values.16,50,52,53,61 The RRR value of the Zr-substituted crystal is found to be 3, and that of the Ce-substituted crystal is 43. The slope of the resistivity curve of all three crystals is similar, with Fermi-liquid behavior at temperatures below 100 K and a T1.5 scaling behavior at higher temperatures. This is consistent with earlier reports of OFZ grown perovskite nickelates.51,52 While the Ce substitution lowers the overall resistivity, as expected for electron doping, the Zr substitution appears to have the opposite effect.

FIG. 9.

(a) Images of the shaped and polished LaNiO3, La0.91Ce0.09NiO3, and La0.95Zr0.05NiO3 single-crystals (left to right), with horizontal c-axis orientation. (b) Resistivity as a function of temperature for the three crystals. The current direction coincides with the c-axis direction. (c) Magnetization as a function of temperature for the three crystals. A field of 0.1 T was applied along the c-axis direction.

FIG. 9.

(a) Images of the shaped and polished LaNiO3, La0.91Ce0.09NiO3, and La0.95Zr0.05NiO3 single-crystals (left to right), with horizontal c-axis orientation. (b) Resistivity as a function of temperature for the three crystals. The current direction coincides with the c-axis direction. (c) Magnetization as a function of temperature for the three crystals. A field of 0.1 T was applied along the c-axis direction.

Close modal

Due to its metallic nature, the magnetic properties of LaNiO3 are not as simple as following the expected Curie laws, unlike PrNiO3. Pure LaNiO3 has complex magnetic correlations that cannot be described by simple Pauli paramagnetism.15 A Stoner enhanced maximum (highlighted by an arrow) is observed in LaNiO3 centered around 220 K [Fig. 9(c)], which cannot be described by simple Pauli paramagnetism. Zr doping leads to a complete suppression of this observed maximum in the magnetic susceptibility. As a result, the full temperature range of the susceptibility of the Zr doped compound can be captured by a Curie-Weiß fit, with C = 1.504(3) · 10−3 emu K/mol, Oe−1 = 1.89(4) · 10−8 m3 K/mol, and θW = −2.2(1) K. As discussed earlier, we take CZr to be C − 0.375. This results in μeff = 3 μB, which is comparable to an expected moment μZr2+=2.8μB of a J = 2 state, suggesting a magnetic Zr contribution. Notably, on the other hand, we observe a similar Stoner enhanced maximum and a somewhat lower Curie tail for Ce substitution. In La0.91Ce0.09NiO3, however, a considerable mass fraction is La2Ce2O7. La2Ce2O7 is solely diamagnetic and results in a small contribution to susceptibility. Therefore, for the Ce doped crystal, the real mass of the perovskite phase was considered, determined from PXRD via Rietveld refinement.

To further characterize our samples and determine the oxidation states of the substituted ions, we investigated the freshly cleaved surfaces of selected single crystals using XPS. Specifically, this technique offers insights into the effectiveness of the intended charge carrier doping. Due to the significant overlap and complex multi-component structures of the La 3d and Ni 2p core-level spectra, we will concentrate on a qualitative analysis in the following.

An overview of all XPS spectra across a wide binding energy range is displayed in Fig. 10(a). In Fig. 10(b), the peaks at 854.8 and 851.2 eV correspond to La 3d3/2, while the doublet at 837.9 and 834.6 eV can be identified as La 3d5/2, indicating the La3+ oxidation state. The La 3d3/2 and Ni 2p1/2 peaks overlap,66 and the Ni 2p3/2 peak is accompanied by a satellite line, which is positioned ∼6–7 eV higher in binding energies. The Ni 2p1/2 peak follows at 872.3 eV.

FIG. 10.

(a) XPS spectra of LaNiO3 (black),50 La0.92Sr0.08NiO3 (green), La0.95Zr0.05NiO3 (blue), and La0.91Ce0.09NiO3 (red) across a wide binding energy range from 100 to 1200 eV. (b) Zoom into the region between 875 and 825 eV, which includes the La 3d and Ni 2p edges. (c) Zoom into the region between 916.5 and 881.5 eV, which includes the Ce 3d edge. Data from a CeO2 film from Ref. 62 are shown for comparison. (d) Zoom into the region between 187 and 176 eV, corresponding to the Zr 3d edge. Data from a ZrO2 film from Ref. 63 and a La2(Zr,Cu)O4 pellet from Ref. 64 are shown for comparison. (e) Zoom into the region between 140 and 125 eV, corresponding to the Sr 3d edge. Data of a (La,Sr)CoO3 from Ref. 65 are shown for comparison.

FIG. 10.

(a) XPS spectra of LaNiO3 (black),50 La0.92Sr0.08NiO3 (green), La0.95Zr0.05NiO3 (blue), and La0.91Ce0.09NiO3 (red) across a wide binding energy range from 100 to 1200 eV. (b) Zoom into the region between 875 and 825 eV, which includes the La 3d and Ni 2p edges. (c) Zoom into the region between 916.5 and 881.5 eV, which includes the Ce 3d edge. Data from a CeO2 film from Ref. 62 are shown for comparison. (d) Zoom into the region between 187 and 176 eV, corresponding to the Zr 3d edge. Data from a ZrO2 film from Ref. 63 and a La2(Zr,Cu)O4 pellet from Ref. 64 are shown for comparison. (e) Zoom into the region between 140 and 125 eV, corresponding to the Sr 3d edge. Data of a (La,Sr)CoO3 from Ref. 65 are shown for comparison.

Close modal

Figure 10(c) focuses on the XPS of La0.91Ce0.09NiO3. Ce 3d is known to exhibit very complex multiplet splitting, resulting in several peaks. Notably, the multiplet structure observed in our La0.91Ce0.09NiO3 spectrum matches the characteristic spectrum of Ce4+ in CeO2,62 thereby confirming the presence of carrier doping in this system since Ce4+ is replacing La3+ sites. In contrast, Ce3+ would give rise to two doublets at 880.9/885.2 and 899.1/903.4 eV,67 which we do not observe.

In the case of the presence of Zr4+ in electron-doped La1−xZrxNiO3, the corresponding peak of the Zr 3d5/2 binding energy is expected to be around 182 eV.63 However, we observe two Zr 3d5/2 doublets, corresponding to binding energies of 181.1 and 180 eV, clearly indicating a lower Zr valency and likely even a mixed valence state [see reference data from ZrO2 in Fig. 10(d)]. This suggests that the anticipated electron doping might not have occurred in this system, which is consistent with the increase in resistivity observed in the Zr-substituted crystal as opposed to the decrease in resistivity in the Ce-substituted crystal [Fig. 9(b)]. Nevertheless, we cannot rule out that the shape of the spectrum in Fig. 10(d) is strongly influenced by surface effects, as in the case of Sr-doping, which is described below.

Finally, we explore the possibility of achieving hole-doping by incorporating Sr2+ ions in La0.92Sr0.08NiO3. Previous works have suggested that hole doping is realized in Sr-substituted nickelate thin films,31–35 but no direct evidence of the Sr2+ valence state has been reported so far, to the best of our knowledge. In the XPS of our La0.92Sr0.08NiO3 crystal, we find that the Sr 3d multiplet exhibits several overlapping peaks [Fig. 10(e)], in particular the 3d5/2 peaks at 131.8 and 133.8 eV, and the 3d3/2 peaks at 133.6 and 135.6 eV. These peaks differ from the simpler 3d doublet at 134.3 and 132.5 eV of Sr2+ in SrTiO3.68 Yet, our observed peak structure is closely similar to that of the Sr ions in La0.4Sr0.6CoO3.65 In the cobaltate, the two doublets were ascribed to distinct surface and lattice Sr sites due to the formation of SrO and Sr(OH)2 on the surface. Such a scenario could apply similarly to the Sr ions in our nickelate sample.

In summary, we have successfully grown La0.98Zr0.02NiO3, La0.95Zr0.05NiO3, and La0.92Sr0.08NiO3. The realization of these doped perovskite nickelates signifies a prominent advancement compared to earlier works that were focused on undoped LaNiO3 and PrNiO3 single-crystals. On the other hand, the (La,Pr,Eu)6Ni5O16 and La2.4Zr0.6Ni2O7 phases did not form under the employed growth conditions. The initial synthesis of Pr0.95Ce0.05NiO3, La0.95Ce0.05NiO3, La0.91Ce0.09NiO3, and La0.84Sr0.16NiO3 was accomplished, albeit with the drawback of substantial amounts of impurity inclusions in the matrix. As a key aspect of our study, we identify that substitutions higher than 8% or any electron doping of the Ruddlesden–Poppers phases (n = 2, 3) lead to phase separation for OFZ growth under 300 bars of oxygen pressure.

Our physical properties and XPS characterizations reveal that electron-doping is realized for Ce-substituted PrNiO3 and LaNiO3, whereas Zr-substitution likely results in mixed-valent states, potentially affecting the metallic ground state and the Stoner enhanced maximum in the magnetic susceptibility.

Our results underscore the pressing need for new, cutting-edge strategies that transcend the conventional single dopant route to realize bulk crystals of nickelate superconductors. Therefore, co-doping with more than one ionic species or self doping via the oxygen stoichiometry, as previously established in cuprates,69 might offer a path to realize superconductivity in topotactically reduced nickelate crystals.

We acknowledge F. Predel for acquiring BSE images at an early stage of this project and C. Busch for technical support. We acknowledge helpful discussions with S. Hayashida and are grateful for access to the Merlin SEM at the Scientific Facility Nanostructuring Lab (NSL) at MPI FKF.

The authors have no conflicts to disclose.

P. Puphal: Conceptualization (equal); Formal analysis (lead); Investigation (lead); Writing – original draft (lead); Writing – review & editing (equal). V. Sundaramurthy: Formal analysis (equal); Investigation (equal). V. Zimmermann: Investigation (equal). K. Küster: Investigation (equal). U. Starke: Conceptualization (equal); Writing – review & editing (equal). M. Isobe: Conceptualization (equal); Writing – review & editing (equal). B. Keimer: Conceptualization (equal); Writing – review & editing (equal). M. Hepting: Conceptualization (equal); Writing – review & editing (lead).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Before the OFZ growth of the La1−xZrxNiO3 substitution series, we performed a study on the growth pressure during a single growth of La0.8Zr0.2NiO3. Our initial attempt employed 15 bars of oxygen pressure, as used in the nominally grown La2.4Zr0.6Ni2O7, and we gradually increased the pressure up to 150 bars, where the growth remained stable. We note that the application of 300 bars of pressure led to growth instabilities.

In Fig. 11, we present a summary of the pressure change effects, revealing that phase formation remains unaffected from 35 bars onwards. Consequently, we chose an intermediate pressure of 85 bars for our final growth. Although the BSE images display a significant phase mixture (Fig. 5), we can still isolate small pieces of major phase accumulations from the n = 1 Ruddlesden–Popper phases and perovskite phases. Interestingly, the perovskite parts appear to contain an increased substitutional content of Zr (Fig. 5), which coexists alongside tiny domains of NiO and the n = 1 Ruddlesden–Popper phase.

FIG. 11.

(a) Image of the boule of a growth of the nominal composition La0.8Zr0.2NiO3 using various oxygen pressures, together with the PXRD patterns from the edge region of the 35 bar part (black), the middle region (red), and the middle region of the 150 bar part (blue). The green bars indicate the peak positions of the perovskite phase from the refinement, while the gray bars indicate the double-perovskite phase. (b) The SE images feature broken pieces of the two separate phases, primarily consisting of n = 1 Ruddlesden–Popper (left) and perovskite phase (right), and an exemplary EDS result (center). The resulting average stoichiometry is given in the legend, with La1.1(2)Zr0.5(1)Ni1.1(2)O4.0(3) and La0.81(1)Zr0.15(7)Ni1.4(3)O3.09(3).

FIG. 11.

(a) Image of the boule of a growth of the nominal composition La0.8Zr0.2NiO3 using various oxygen pressures, together with the PXRD patterns from the edge region of the 35 bar part (black), the middle region (red), and the middle region of the 150 bar part (blue). The green bars indicate the peak positions of the perovskite phase from the refinement, while the gray bars indicate the double-perovskite phase. (b) The SE images feature broken pieces of the two separate phases, primarily consisting of n = 1 Ruddlesden–Popper (left) and perovskite phase (right), and an exemplary EDS result (center). The resulting average stoichiometry is given in the legend, with La1.1(2)Zr0.5(1)Ni1.1(2)O4.0(3) and La0.81(1)Zr0.15(7)Ni1.4(3)O3.09(3).

Close modal

In addition to the extensive XRD and BSE analysis, we also performed EDS characterization, with representative results summarized in Fig. 12. We focused the EDS analysis on homogeneous pieces with lower doping levels, confirming the successful substitution through a systematic examination of at least 20 points across various regions of the grown boule on the phase-pure part of the perovskite matrix. An example SEM image is provided in Fig. 12.

FIG. 12.

EDS scans of crystals, with an exemplary SE image of a broken piece shown in the inset, for nominally grown (a) Pr0.95Ce0.05NiO3, (b) La0.91Ce0.09NiO3, (c) (La,Zr)NiO3, (d) La0.95Zr0.05NiO3, and (e) La0.92Sr0.08NiO3. (f) EDS map of La0.84Sr0.16NiO3 of a “cleaved” cross section of the boule, with all elements overlaid in one picture on the SE picture, using the color code: La (red), Sr (purple), O (blue), and Ni (green).

FIG. 12.

EDS scans of crystals, with an exemplary SE image of a broken piece shown in the inset, for nominally grown (a) Pr0.95Ce0.05NiO3, (b) La0.91Ce0.09NiO3, (c) (La,Zr)NiO3, (d) La0.95Zr0.05NiO3, and (e) La0.92Sr0.08NiO3. (f) EDS map of La0.84Sr0.16NiO3 of a “cleaved” cross section of the boule, with all elements overlaid in one picture on the SE picture, using the color code: La (red), Sr (purple), O (blue), and Ni (green).

Close modal

By averaging over all measured points and considering the standard deviation as error bars, we obtained the following substitutional contents: (a) Pr0.93(3)Ce0.02(1)NiO3, (b) La0.96(2)Ce0.046(3)NiO3, (c) 5%: La0.99(9)Zr0.071(4)NiO3, 8.75%: La0.96(5)Zr0.11(1)NiO3, 12.5%: La0.9(2)Zr0.12(2)NiO3, and 16.75%: La0.91(6)Zr0.12(1)NiO3, (d) La0.9(2)Zr0.07(2)NiO3, and (e) La0.8(1)Sr0.12(5)NiO3.

The phase separation, evident in both BSE and XRD data for doping levels of 8% and higher, is due to the solubility limit of the substituent. This phenomenon is also reflected in the EDS data, which reveal a limited substitution content in the bulk matrix. It is important to note that, in comparison to the rare-earth elements, the Zr and Sr content is systematically overestimated in our EDS analysis by a few percent. The radial distribution of phases in the boule, resulting from the varying oxygen partial pressure within the melt, is best illustrated in our EDS map displayed in Fig. 12(f). In this figure, all elements are color coded and superimposed onto the same SE image, revealing a red perovskite region at the exterior and a blue n = 1 Ruddlesden–Popper phase region at the interior of the grown boule.

1.
M.
Greenblatt
, “
Ruddlesden-Popper Lnn+1NinO3n+1 nickelates: Structure and properties
,”
Curr. Opin. Solid State Mater. Sci.
2
,
174
(
1997
).
2.
S.
Sugai
,
M.
Sato
,
T.
Kobayashi
,
J.
Akimitsu
,
T.
Ito
,
H.
Takagi
,
S.
Uchida
,
S.
Hosoya
,
T.
Kajitani
, and
T.
Fukuda
, “
High-energy spin excitations in the insulating phases of high-Tc superconducting cuprates and La2NiO4
,”
Phys. Rev. B
42
,
1045(R)
(
1990
).
3.
Y.
Kobayashi
,
S.
Taniguchi
,
M.
Kasai
,
M.
Sato
,
T.
Nishioka
, and
M.
Kontani
, “
Transport and magnetic properties of La3Ni2O7−δ and La4Ni3O10−δ
,”
J. Phys. Soc. Jpn.
65
,
3978
(
1996
).
4.
I. I.
Mazin
,
D. I.
Khomskii
,
R.
Lengsdorf
,
J. A.
Alonso
,
W. G.
Marshall
,
R. M.
Ibberson
,
A.
Podlesnyak
,
M. J.
Martínez-Lope
, and
M. M.
Abd-Elmeguid
, “
Charge ordering as alternative to Jahn-Teller distortion
,”
Phys. Rev. Lett.
98
,
176406
(
2007
).
5.
Z.
Li
,
W.
Guo
,
T. T.
Zhang
,
J. H.
Song
,
T. Y.
Gao
,
Z. B.
Gu
, and
Y. F.
Nie
, “
Epitaxial growth and electronic structure of Ruddlesden–Popper nickelates (Lan+1NinO3n+1, n = 1–5)
,”
APL Mater.
8
,
091112
(
2020
).
6.
S.
Catalano
,
M.
Gibert
,
J.
Fowlie
,
J.
Íñiguez
,
J.-M.
Triscone
, and
J.
Kreisel
, “
Rare-earth nickelates RNiO3: Thin films and heterostructures
,”
Rep. Prog. Phys.
81
,
046501
(
2018
).
7.
A.
Petrov
,
V.
Cherepanov
,
A.
Zuyev
, and
V.
Zhukovsky
, “
Thermodynamic stability of ternary oxides in Ln M O (Ln = La, Pr, Nd; M = Co, Ni, Cu) systems
,”
J. Solid State Chem.
77
,
1
(
1988
).
8.
A.
Wold
,
B.
Post
, and
E.
Banks
, “
Rare earth nickel oxides
,”
J. Am. Chem. Soc.
79
,
4911
(
1957
).
9.
P.
Lacorre
,
J.
Torrance
,
J.
Pannetier
,
A.
Nazzal
,
P.
Wang
, and
T.
Huang
, “
Synthesis, crystal structure, and properties of metallic PrNiO3: Comparison with metallic NdNiO3 and semiconducting SmNiO3
,”
J. Solid State Chem.
91
,
225
(
1991
).
10.
J. B.
Torrance
,
P.
Lacorre
,
A. I.
Nazzal
,
E. J.
Ansaldo
, and
C.
Niedermayer
, “
Systematic study of insulator-metal transitions in perovskites RNiO3 (R=Pr,Nd,Sm,Eu) due to closing of charge-transfer gap
,”
Phys. Rev. B
45
,
8209(R)
(
1992
).
11.
M.
Medarde
,
A.
Fontaine
,
J. L.
García-Muñoz
,
J.
Rodríguez-Carvajal
,
M.
de Santis
,
M.
Sacchi
,
G.
Rossi
, and
P.
Lacorre
, “
RNiO3 perovskites (R=Pr,Nd): Nickel valence and the metal-insulator transition investigated by x-ray-absorption spectroscopy
,”
Phys. Rev. B
46
,
14975
(
1992
).
12.
G.
Catalan
, “
Progress in perovskite nickelate research
,”
Phase Transitions
81
,
729
(
2008
).
13.
V.
Scagnoli
,
U.
Staub
,
A. M.
Mulders
,
M.
Janousch
,
G. I.
Meijer
,
G.
Hammerl
,
J. M.
Tonnerre
, and
N.
Stojic
, “
Role of magnetic and orbital ordering at the metal-insulator transition in NdNiO3
,”
Phys. Rev. B
73
,
100409(R)
(
2006
).
14.
M.
Hepting
,
R. J.
Green
,
Z.
Zhong
,
M.
Bluschke
,
Y. E.
Suyolcu
,
S.
Macke
,
A.
Frano
,
S.
Catalano
,
M.
Gibert
,
R.
Sutarto
,
F.
He
,
G.
Cristiani
,
G.
Logvenov
,
Y.
Wang
,
P. A.
van Aken
,
P.
Hansmann
,
M.
Le Tacon
,
J.-M.
Triscone
,
G. A.
Sawatzky
,
B.
Keimer
, and
E.
Benckiser
, “
Complex magnetic order in nickelate slabs
,”
Nat. Phys.
14
,
1097
(
2018
).
15.
B.-X.
Wang
,
S.
Rosenkranz
,
X.
Rui
,
J.
Zhang
,
F.
Ye
,
H.
Zheng
,
R. F.
Klie
,
J. F.
Mitchell
, and
D.
Phelan
, “
Antiferromagnetic defect structure in LaNiO3−δ single crystals
,”
Phys. Rev. Mater.
2
,
064404
(
2018
).
16.
H.
Zheng
,
B.-X.
Wang
,
D.
Phelan
,
J.
Zhang
,
Y.
Ren
,
M.
Krogstad
,
S.
Rosenkranz
,
R.
Osborn
, and
J.
Mitchell
, “
Oxygen inhomogeneity and reversibility in single crystal LaNiO3−δ
,”
Crystals
10
,
557
(
2020
).
17.
L.
Iglesias
,
M.
Bibes
, and
J.
Varignon
, “
First-principles study of electron and hole doping effects in perovskite nickelates
,”
Phys. Rev. B
104
,
035123
(
2021
).
18.
A. S.
Botana
,
V.
Pardo
, and
M. R.
Norman
, “
Electron doped layered nickelates: Spanning the phase diagram of the cuprates
,”
Phys. Rev. Mater.
1
,
021801(R)
(
2017
).
19.
M.
Kotiuga
and
K. M.
Rabe
, “
High-density electron doping of SmNiO3 from first principles
,”
Phys. Rev. Mater.
3
,
115002
(
2019
).
20.
P.
Worm
,
L.
Si
,
M.
Kitatani
,
R.
Arita
,
J. M.
Tomczak
, and
K.
Held
, “
Correlations tune the electronic structure of pentalayer nickelates into the superconducting regime
,”
Phys. Rev. Mater.
6
,
L091801
(
2022
).
21.
X.
Chen
,
P.
Jiang
, and
Z.
Zhong
, “
Electron-doping Ruddlesden-Popper nickelate
,”
Europhys. Lett.
142
,
26005
(
2023
).
22.
R. K.
Patel
,
K.
Patra
,
S. K.
Ojha
,
S.
Kumar
,
S.
Sarkar
,
A.
Saha
,
N.
Bhattacharya
,
J. W.
Freeland
,
J.-W.
Kim
,
P. J.
Ryan
,
P.
Mahadevan
, and
S.
Middey
, “
Hole doping in a negative charge transfer insulator
,”
Commun. Phys.
5
,
216
(
2022
).
23.
J.
Li
,
S.
Ramanathan
, and
R.
Comin
, “
Carrier doping physics of rare earth perovskite nickelates RENiO3
,”
Front. Phys.
10
,
834882
(
2022
).
24.
Q.
Song
,
S.
Doyle
,
G. A.
Pan
,
I.
El Baggari
,
D.
Ferenc Segedin
,
D.
Córdova Carrizales
,
J.
Nordlander
,
C.
Tzschaschel
,
J. R.
Ehrets
,
Z.
Hasan
,
H.
El-Sherif
,
J.
Krishna
,
C.
Hanson
,
H.
LaBollita
,
A.
Bostwick
,
C.
Jozwiak
,
E.
Rotenberg
,
S.-Y.
Xu
,
A.
Lanzara
,
A. T.
N’Diaye
,
C. A.
Heikes
,
Y.
Liu
,
H.
Paik
,
C. M.
Brooks
,
B.
Pamuk
,
J. T.
Heron
,
P.
Shafer
,
W. D.
Ratcliff
,
A. S.
Botana
,
L.
Moreschini
, and
J. A.
Mundy
, “
Antiferromagnetic metal phase in an electron-doped rare-earth nickelate
,”
Nat. Phys.
19
,
522
(
2023
).
25.
S.-W.
Cheong
,
H.
Hwang
,
B.
Batlogg
,
A.
Cooper
, and
P.
Canfield
, “
Electron-hole doping of the metal-insulator transition compound RENiO3
,”
Physica B
194-196
,
1087
(
1994
).
26.
J.
Alonso
,
M.
Martínez-Lope
, and
M.
Hidalgo
, “
Hole and electron doping of RNiO3 (R = La, Nd)
,”
J. Solid State Chem.
116
,
146
(
1995
).
27.
J. L.
García-Muñoz
,
M.
Suaaidi
,
M. J.
Martínez-Lope
, and
J. A.
Alonso
, “
Influence of carrier injection on the metal-insulator transition in electron- and hole-doped R1−xAxNiO3 perovskites
,”
Phys. Rev. B
52
,
13563
(
1995
).
28.
K.
Ramadoss
,
N.
Mandal
,
X.
Dai
,
Z.
Wan
,
Y.
Zhou
,
L.
Rokhinson
,
Y. P.
Chen
,
J.
Hu
, and
S.
Ramanathan
, “
Sign reversal of magnetoresistance in a perovskite nickelate by electron doping
,”
Phys. Rev. B
94
,
235124
(
2016
).
29.
Q.
Li
,
C.
He
,
J.
Si
,
X.
Zhu
,
Y.
Zhang
, and
H.-H.
Wen
, “
Absence of superconductivity in bulk Nd1−xSrxNiO2
,”
Commun. Mater.
1
,
16
(
2020
).
30.
J. F.
Mitchell
, “
A nickelate renaissance
,”
Front. Phys.
9
,
813483
(
2021
).
31.
D.
Li
,
K.
Lee
,
B. Y.
Wang
,
M.
Osada
,
S.
Crossley
,
H. R.
Lee
,
Y.
Cui
,
Y.
Hikita
, and
H. Y.
Hwang
, “
Superconductivity in an infinite-layer nickelate
,”
Nature
572
,
624
(
2019
).
32.
S.
Zeng
,
C. S.
Tang
,
X.
Yin
,
C.
Li
,
M.
Li
,
Z.
Huang
,
J.
Hu
,
W.
Liu
,
G. J.
Omar
,
H.
Jani
,
Z. S.
Lim
,
K.
Han
,
D.
Wan
,
P.
Yang
,
S. J.
Pennycook
,
A. T. S.
Wee
, and
A.
Ariando
, “
Phase diagram and superconducting dome of infinite-layer Nd1−xSrxNiO2 thin films
,”
Phys. Rev. Lett.
125
,
147003
(
2020
).
33.
K.
Lee
,
B. H.
Goodge
,
D.
Li
,
M.
Osada
,
B. Y.
Wang
,
Y.
Cui
,
L. F.
Kourkoutis
, and
H. Y.
Hwang
, “
Aspects of the synthesis of thin film superconducting infinite-layer nickelates
,”
APL Mater.
8
,
041107
(
2020
).
34.
M.
Osada
,
B. Y.
Wang
,
B. H.
Goodge
,
S. P.
Harvey
,
K.
Lee
,
D.
Li
,
L. F.
Kourkoutis
, and
H. Y.
Hwang
, “
Nickelate superconductivity without rare-earth magnetism: (La,Sr)NiO2
,”
Adv. Mater.
33
,
2104083
(
2021
).
35.
Q.
Gao
,
Y.
Zhao
,
X.-J.
Zhou
, and
Z.
Zhu
, “
Preparation of superconducting thin films of infinite-layer nickelate Nd0.8Sr0.2NiO2
,”
Chin. Phys. Lett.
38
,
077401
(
2021
).
36.
S.
Zeng
,
C.
Li
,
L. E.
Chow
,
Y.
Cao
,
Z.
Zhang
,
C. S.
Tang
,
X.
Yin
,
Z. S.
Lim
,
J.
Hu
,
P.
Yang
, and
A.
Ariando
, “
Superconductivity in infinite-layer nickelate La1−xCaxNiO2 thin films
,”
Sci. Adv.
8
,
eabl9927
(
2022
).
37.
M.
Crespin
,
P.
Levitz
, and
L.
Gatineau
, “
Reduced forms of LaNiO3 perovskite. Part 1.—Evidence for new phases: La2Ni2O5 and LaNiO2
,”
J. Chem. Soc., Faraday Trans. 2
79
,
1181
(
1983
).
38.
M. A.
Hayward
,
M. A.
Green
,
M. J.
Rosseinsky
, and
J.
Sloan
, “
Sodium hydride as a powerful reducing agent for topotactic oxide deintercalation synthesis and characterization of the nickel(I) oxide LaNiO2
,”
J. Am. Chem. Soc.
121
,
8843
(
1999
).
39.
M. A.
Hayward
and
M. J.
Rosseinsky
, “
Synthesis of the infinite layer Ni(I) phase NdNiO2+x by low temperature reduction of NdNiO3 with sodium hydride
,”
Solid State Sci.
5
,
839
(
2003
).
40.
M.
Hepting
,
M. P. M.
Dean
, and
W.-S.
Lee
, “
Soft x-ray spectroscopy of low-valence nickelates
,”
Front. Phys.
9
,
808683
(
2021
).
41.
G. A.
Pan
,
D.
Ferenc Segedin
,
H.
LaBollita
,
Q.
Song
,
E. M.
Nica
,
B. H.
Goodge
,
A. T.
Pierce
,
S.
Doyle
,
S.
Novakov
,
D.
Córdova Carrizales
,
A. T.
N’Diaye
,
P.
Shafer
,
H.
Paik
,
J. T.
Heron
,
J. A.
Mason
,
A.
Yacoby
,
L. F.
Kourkoutis
,
O.
Erten
,
C. M.
Brooks
,
A. S.
Botana
, and
J. A.
Mundy
, “
Superconductivity in a quintuple-layer square-planar nickelate
,”
Nat. Mater.
21
,
160
(
2021
).
42.
A. S.
Botana
,
K.-W.
Lee
,
M. R.
Norman
,
V.
Pardo
, and
W. E.
Pickett
, “
Low valence nickelates: Launching the nickel age of superconductivity
,”
Front. Phys.
9
,
813532
(
2022
).
43.
B.-X.
Wang
,
H.
Zheng
,
E.
Krivyakina
,
O.
Chmaissem
,
P. P.
Lopes
,
J. W.
Lynn
,
L. C.
Gallington
,
Y.
Ren
,
S.
Rosenkranz
,
J. F.
Mitchell
, and
D.
Phelan
, “
Synthesis and characterization of bulk Nd1−xSrxNiO2 and Nd1−xSrxNiO3
,”
Phys. Rev. Mater.
4
,
084409
(
2020
).
44.
B. H.
Goodge
,
B.
Geisler
,
K.
Lee
,
M.
Osada
,
B. Y.
Wang
,
D.
Li
,
H. Y.
Hwang
,
R.
Pentcheva
, and
L. F.
Kourkoutis
, “
Resolving the polar interface of infinite-layer nickelate thin films
,”
Nat. Mater.
22
,
466
(
2023
).
45.
J.
Zhang
,
A.
Botana
,
J.
Freeland
,
D.
Phelan
,
H.
Zheng
,
V.
Pardo
,
M.
Norman
, and
J.
Mitchell
, “
Large orbital polarization in a metallic square-planar nickelate
,”
Nat. Phys.
13
,
864
(
2017
).
46.
P.
Puphal
,
V.
Pomjakushin
,
R. A.
Ortiz
,
S.
Hammoud
,
M.
Isobe
,
B.
Keimer
, and
M.
Hepting
, “
Investigation of hydrogen incorporations in bulk infinite-layer nickelates
,”
Front. Phys.
10
,
842578
(
2022
).
47.
R. A.
Ortiz
,
P.
Puphal
,
M.
Klett
,
F.
Hotz
,
R. K.
Kremer
,
H.
Trepka
,
M.
Hemmida
,
H.-A. K.
von Nidda
,
M.
Isobe
,
R.
Khasanov
,
H.
Luetkens
,
P.
Hansmann
,
B.
Keimer
,
T.
Schäfer
, and
M.
Hepting
, “
Magnetic correlations in infinite-layer nickelates: An experimental and theoretical multimethod study
,”
Phys. Rev. Res.
4
,
023093
(
2022
).
48.
P.
Puphal
,
Y.-M.
Wu
,
K.
Fürsich
,
H.
Lee
,
M.
Pakdaman
,
J. A. N.
Bruin
,
J.
Nuss
,
Y. E.
Suyolcu
,
P. A.
van Aken
,
B.
Keimer
,
M.
Isobe
, and
M.
Hepting
, “
Topotactic transformation of single crystals: From perovskite to infinite-layer nickelates
,”
Sci. Adv.
7
,
eabl8091
(
2021
).
49.
Y.-M.
Wu
,
P.
Puphal
,
H.
Lee
,
J.
Nuss
,
M.
Isobe
,
B.
Keimer
,
M.
Hepting
,
Y. E.
Suyolcu
, and
P. A.
van Aken
, “
Topotactically induced oxygen vacancy order in nickelate single crystals
,”
Phys. Rev. Mater.
7
,
053609
(
2023
).
50.
P.
Puphal
,
B.
Wehinger
,
J.
Nuss
,
K.
Küster
,
U.
Starke
,
G.
Garbarino
,
B.
Keimer
,
M.
Isobe
, and
M.
Hepting
, “
Synthesis and physical properties of LaNiO2 crystals
,”
Phys. Rev. Mater.
7
,
014804
(
2023
).
51.
H.
Guo
,
Z. W.
Li
,
L.
Zhao
,
Z.
Hu
,
C. F.
Chang
,
C.-Y.
Kuo
,
W.
Schmidt
,
A.
Piovano
,
T. W.
Pi
,
O.
Sobolev
,
D. I.
Khomskii
,
L. H.
Tjeng
, and
A. C.
Komarek
, “
Antiferromagnetic correlations in the metallic strongly correlated transition metal oxide LaNiO3
,”
Nat. Commun.
9
,
43
(
2018
).
52.
J.
Zhang
,
H.
Zheng
,
Y.
Ren
, and
J. F.
Mitchell
, “
High-pressure floating-zone growth of perovskite nickelate LaNiO3 single crystals
,”
Cryst. Growth Des.
17
,
2730
(
2017
).
53.
K.
Dey
,
W.
Hergett
,
P.
Telang
,
M. M.
Abdel-Hafiez
, and
R.
Klingeler
, “
Magnetic properties of high-pressure optical floating-zone grown LaNiO3 single crystals
,”
J. Cryst. Growth
524
,
125157
(
2019
).
54.
H.
Zheng
,
J.
Zhang
,
B.
Wang
,
D.
Phelan
,
M. J.
Krogstad
,
Y.
Ren
,
W. A.
Phelan
,
O.
Chmaissem
,
B.
Poudel
, and
J. F.
Mitchell
, “
High pO2 floating zone crystal growth of the perovskite nickelate PrNiO3
,”
Crystals
9
,
324
(
2019
).
55.
Z.
Liu
,
H.
Sun
,
M.
Huo
,
X.
Ma
,
Y.
Ji
,
E.
Yi
,
L.
Li
,
H.
Liu
,
J.
Yu
,
Z.
Zhang
,
Z.
Chen
,
F.
Liang
,
H.
Dong
,
H.
Guo
,
D.
Zhong
,
B.
Shen
,
S.
Li
, and
M.
Wang
, “
Evidence for charge and spin density waves in single crystals of La3Ni2O7 and La3Ni2O6
,”
Sci. China: Phys., Mech. Astron.
66
,
217411
(
2022
).
56.
J.
Zhang
,
H.
Zheng
,
Y.-S.
Chen
,
Y.
Ren
,
M.
Yonemura
,
A.
Huq
, and
J. F.
Mitchell
, “
High oxygen pressure floating zone growth and crystal structure of the metallic nickelates R4Ni3O10 (R = La, Pr)
,”
Phys. Rev. Mater.
4
,
083402
(
2020
).
57.
S.
Huangfu
,
G. D.
Jakub
,
X.
Zhang
,
O.
Blacque
,
P.
Puphal
,
E.
Pomjakushina
,
F. O.
von Rohr
, and
A.
Schilling
, “
Anisotropic character of the metal-to-metal transition in Pr4Ni3O10
,”
Phys. Rev. B
101
,
104104
(
2020
).
58.
J.
Rodríguez-Carvajal
, “
Recent advances in magnetic structure determination by neutron powder diffraction
,”
Physica B
192
,
55
(
1993
).
59.
N.
Fairley
,
V.
Fernandez
,
M.
Richard‐Plouet
,
C.
Guillot-Deudon
,
J.
Walton
,
E.
Smith
,
D.
Flahaut
,
M.
Greiner
,
M.
Biesinger
,
S.
Tougaard
,
D.
Morgan
, and
J.
Baltrusaitis
, “
Systematic and collaborative approach to problem solving using X-ray photoelectron spectroscopy
,”
Appl. Surf. Sci.
5
,
100112
(
2021
).
60.
Y. M.
Klein
,
M.
Kozłowski
,
A.
Linden
,
P.
Lacorre
,
M.
Medarde
, and
D. J.
Gawryluk
, “
RENiO3 single crystals (RE = Nd, Sm, Gd, Dy, Y, Ho, Er, Lu) grown from molten salts under 2000 bar of oxygen gas pressure
,”
Cryst. Growth Des.
21
,
4230
(
2021
).
61.
Y.
Tomioka
,
T.
Ito
,
E.
Maruyama
,
S.
Kimura
, and
I.
Shindo
, “
Magnetic and electronic properties of single crystals of perovskite nickelate oxide LaNiO3 prepared by the laser diode floating zone method
,”
J. Phys. Soc. Jpn.
90
,
034704
(
2021
).
62.
A.
Pfau
,
K. D.
Schierbaum
, and
W.
Göpel
, “
Comparison spectra of (111) textured CeO2 thin films by XPS
,”
Surf. Sci. Spectra
4
,
288
(
1996
).
63.
P.
Lackner
,
Z.
Zou
,
S.
Mayr
,
U.
Diebold
, and
M.
Schmid
, “
Using photoelectron spectroscopy to observe oxygen spillover to zirconia
,”
Phys. Chem. Chem. Phys.
21
,
17613
(
2019
).
64.
J. A.
Anderson
and
J. L. G.
Fierro
, “
Bulk and surface properties of copper-containing oxides of the general formula LaZr1−xCuxO3
,”
J. Solid State Chem.
108
,
305
(
1994
).
65.
X.
Wang
,
K.
Huang
,
L.
Yuan
,
S.
Xi
,
W.
Yan
,
Z.
Geng
,
Y.
Cong
,
Y.
Sun
,
H.
Tan
,
X.
Wu
,
L.
Li
, and
S.
Feng
, “
Activation of surface oxygen sites in a cobalt-based perovskite model catalyst for CO oxidation
,”
J. Phys. Chem. Lett.
9
,
4146
(
2018
).
66.
S.
Mickevičius
,
S.
Grebinskij
,
V.
Bondarenka
,
B.
Vengalis
,
K.
Šliužienė
,
B.
Orlowski
,
V.
Osinniy
, and
W.
Drube
, “
Investigation of epitaxial LaNiO3−x thin films by high-energy XPS
,”
J. Alloys Compd.
423
,
107
(
2006
).
67.
E.
Bêche
,
P.
Charvin
,
D.
Perarnau
,
S.
Abanades
, and
G.
Flamant
, “
Ce 3d XPS investigation of cerium oxides and mixed cerium oxide (CexTiyOz)
,”
Surf. Interface Anal.
40
,
264
(
2008
).
68.
R. P.
Vasquez
, “
SrTiO3 by XPS
,”
Surf. Sci. Spectra
1
,
129
(
1992
).
69.
A.
Frano
,
M.
Bluschke
,
Z.
Xu
,
B.
Frandsen
,
Y.
Lu
,
M.
Yi
,
R.
Marks
,
A.
Mehta
,
V.
Borzenets
,
D.
Meyers
,
M. P. M.
Dean
,
F.
Baiutti
,
J.
Maier
,
G.
Kim
,
G.
Christiani
,
G.
Logvenov
,
E.
Benckiser
,
B.
Keimer
, and
R. J.
Birgeneau
, “
Control of dopant crystallinity in electrochemically treated cuprate thin films
,”
Phys. Rev. Mater.
3
,
063803
(
2019
).
Published open access through an agreement with Max-Planck-Gesellschaft