1 Introduction
Ever since the discovery of copper-based high-temperature superconductivity, there has been a strong motivation to identify non-cuprates superconductor with similar cuprates-like electronic structures. Considering Ni is next to Cu in the periodic table, there are high expectations to find superconductivity in nickel oxides [
1]. Initial attempts in the past three decades turned out to be unsuccessful, until recently, Li
et al. [
2] discovered superconductivity in infinite-layer nickelate Nd
0.8Sr
0.2NiO
2 epitaxial films, leading to a new era for the research of superconductivity. In addition, understanding of the nickel-based superconductivity might provide useful clue to elucidate the superconductivity mechanisms for cuprates and pnictides.
Despite sharing many similarities (i.e., 3d
9 electronic configuration, hole-doping superconducting phase-diagrams [
3,
4], small coherent length [
5], weak electron−phonon coupling [
6], superconducting dimensionality [
5,
7]), nickelates still exhibit some essential differences compared with cuprates. For example, cuprates are mainly charge-transfer insulators with doped holes residing in O 2p orbitals forming Zhang−Rice singlets, while nickelates show Mott−Hubbard-like behavior with doped holes staying in Ni 3d bands. Cuprates show ideal single-band feature, while nickelates show multi-band feature (Ni 3d
x2−y2, Ni 3d
z2, Nd 5d and interstitial orbitals) [
8,
9]. Cuprates show higher transition temperature (
Tc) with long-range antiferromagnetic ordering, while nickelates show lower
Tc and local antiferromagnetic fluctuations. Recently, charge density wave, spin excitation and strong Pauli-limit violation have been identified in infinite-layer nickelates [
5,
8,
10-
13].
Currently, there are few debates on the physics of infinite-layer nickelates, e.g., intrinsic magnetism (whether long-range or short-range antiferromagnetic ordering) [
8,
14-
16] and pair mechanism (s-wave, d-wave, or s/d wave) [
6,
17-
20]. Moreover, nickelates superconductivity is only observed in epitaxial films, not in bulk polycrystalline or single crystals [
21]. These issues reflect our understanding of the fundamental properties of infinite-layer nickelates is still limited, which is essentially restricted by the challenge in material synthesis.
The chemical synthesis of infinite-layer nickelate can be traced back to early 1980s. Since then, scientists have been studying to find effective methods for chemical reduction until metal hydrides are proposed as reducing agent [Fig.1(a)], which can work at relatively low temperatures [
22,
23]. This technique is then applied to perform reduction for epitaxial perovskite films (e.g., LaNiO
3, BaTiO
3, SrFeO
3, SrVO
3) [
24-
28] and layered materials (e.g., LaFeAsO, Sr
3Co
2O
4, Sr
3Fe
2O
4Cl
2, LaBaCo
2O
5) [
29-
32]. Compared with other perovskites, the difficulty of chemical reduction for doped nickelates is twofold: (i) the most stable valence state for Ni is +2, while high valence states (e.g., +3 for LaNiO
3 and +3.2 for Nd
0.8Sr
0.2NiO
3) and ultra-low valence state (+1 for LaNiO
2) are thermodynamically unstable [
23]. (ii) Sr doping decreases the stability for perovskite nickelate, thus secondary phases or structure defects (Ruddlesden−Popper domains) often coexist [
33,
34], which is detrimental for achieving superconductivity. Therefore, the synthesis of superconducting nickelate is a great challenge and currently only a few groups worldwide can obtain high-quality superconducting infinite-layer film with zero-resistance [
2,
3,
15,
20,
35-
38]. This problem becomes the major barrier for the development of this field, thus requires further investigation.
Our group has recently made progress in synthesizing superconducting nickelate epitaxial films [
11,
36]. In this work, using 20% Sr-doped neodymium nickelate as an example, we systematically investigate the thin-film growth process during pulsed laser deposition (PLD) from plume dynamics perspective. Through thorough materials characterizations, we show the challenges for obtaining superconducting nickelates involve two key aspects. One is to obtain infinite-layer crystal structure and the other is to achieve ideal superconducting state. In particular, the laser energy density and growth temperature are demonstrated as critical parameters to obtain phase-pure infinite-layer Nd
0.8Sr
0.2NiO
2, which is important to achieve superconducting nickelates. Under optimized growth conditions, Nd
0.8Sr
0.2NiO
2 film exhibits a
Tc of 13 K and zero-resistance at 7 K. We also demonstrate that the superconducting Nd
0.8Sr
0.2NiO
2 film is very stable in ambient conditions up to 512 days.
2 Results and discussion
PLD is a well-established technique to grow epitaxial high-quality crystalline materials, particularly for high crystallographic symmetry oxides such as perovskites and spinels [
39,
40], for both scientific research and industry applications. A distinguished feature of PLD is the congruent material transfer under ultrafast laser pulses, which enables to grow high-quality films with complex cation composition [
41,
42]. Technically, many instrumental parameters can affect the film growth process, such as substrate temperature, oxygen pressure, laser energy density, laser spot size/uniformity, laser frequency, deposition rate, growth time/film thickness, and substrate-target distances. Practically, the first three parameters are the most important, which deserve systematical study.
For nickelates, synergetic Pauli exclusion principle and Hund coupling enables the d
8 electronic configuration (t
2g6e
g2) to be more energetically stable than others such as d
7 (t
2g6e
g1), d
6 (t
2g6e
g0), or d
9 (t
2g6e
g3), making Ni
+2 the most favorable valence state for nickel oxides. This is the fundamental reason that nickel monoxide NiO is naturally more common than nickel sesquioxide Ni
2O
3. As a result, to achieve higher Ni valence states, e.g., Ni
3+ for LaNiO
3, high oxygen pressure or ozone is needed to compensate oxygen loss during materials synthesis [
43,
44]. Otherwise, oxygen vacancy still dominates, leading to Ni
2+/Ni
3+ mixtures, which masks the intrinsic material properties or physical phenomena [
45]. Over the past decades, it is well-established that to ensure high-quality epitaxial LaNiO
3 films with minimal oxygen vacancies, high oxygen pressure during film deposition is critical [
46,
47]. Considering Nd
0.8Sr
0.2NiO
3 shows even higher Ni valence states (Ni
3.2+) than LaNiO
3 (Ni
3+) due to 20% divalent Sr
2+ doping, we use the highest oxygen pressure of 200 mTorr, which is the limit that can be tolerated by our high-pressure RHEED system. As a result, we narrow down the aforementioned three critical growth parameters into two, i.e., laser energy density and substrate temperature, which will be the focus in the following study.
Although two early studies have considered the film processing conditions for superconducting nickelates [
33,
38], the experiments and discussions are not rigorous, i.e., they either did not use the same sample that has been subjected to consecutive growth and reduction processes, or the considered samples are not under the same conditions (i.e., there exists extra variation on film thickness, or cover layer, or reduction parameters), thus these studies are not based on single variation of parameters. In this work, to avoid the similar problem of experiment design and to ensure the accuracy of single variation of parameters, precise film growth rates (under different laser energy density and substrate temperature) are determined by RHEED oscillations and XRR, respectively. Then, all film depositions are performed with calibrated growth rate to ensure identical film thickness for all samples.
Typical RHEED diffraction patterns and intensity oscillations for Nd
0.8Sr
0.2NiO
3 films are first studied. Fig.1(b) and (c) show the representative RHEED spots before and after Nd
0.8Sr
0.2NiO
3 film growth along [110] orientation. The sharp RHEED pattern indicates the film surface after growth is flat. Fig.1(d) shows the RHEED intensity oscillation monitored for the films grown at different laser fluences. The periodic oscillation indicates a typical layer-by-layer growth mode. Note RHEED oscillation during initial growth stage shows nonuniform periodicity, indicating possible existence of surface reconstruction and roughing, which may reflects the competition between the Ruddlesden−Popper phase and the perovskite phase due to comparable formation energy [
33].
Fig.2(a) shows XRD scans of 15-nm-thick epitaxial Nd
0.8Sr
0.2NiO
3 films under different laser energy densities (1.0−1.3 J·cm
−2). All growth was performed at substrate temperature of 600 °C, oxygen pressure of 200 mTorr and laser frequency of 4 Hz. Generally, all films show well-defined diffraction peaks with clear Kiessig fringes, suggesting good film quality. The left inset in Fig.2(a) shows the reciprocal space mapping (RSM) of Nd
0.8Sr
0.2NiO
3 film near SrTiO
3 (103) reflection, where the film is clamped to SrTiO
3 in-plane lattice, indicating a coherent strain state with good epitaxial relationship. The right inset in Fig.2(a) shows a representative AFM image, indicating the film surface exhibits an atomically smooth step-like structure, further confirming the high-quality perovskite film growth. Note at laser energy density of 1.0 and 1.3 J·cm
−2, the film (001) peak is strongly suppressed, indicating Ruddlesden−Popper defective phase [
33]. Fig.2(b) shows a zoomed-in XRD near the (002) peak with the inset showing the extracted
c lattice parameters. The (002) peak positions show clear laser energy density dependency, e.g., samples at the intermediate energy densities (1.1 and 1.2 J·cm
−2) show higher (002) peak positions exceeding 48.4° (
c lattice parameter is close to bulk value of 3.75 Å), while samples at the large and small energy densities (1.0 and 1.3 J·cm
−2) show low (002) peak positions below 48° with suppressed peak intensity, reflecting dominancy of Ruddlesden−Popper phase [
33]. The persistent XRR oscillations in Fig.2(c) also confirm the flat film surface, consistent with AFM results. The film thickness determined by XRR is also consistent with RHEED oscillations.
The as-grown Nd
0.8Sr
0.2NiO
3 films are then subjected to CaH
2 topotactic reduction at the optimized conditions (300 °C and 2 hours) [
11,
36]. The XRD spectra for the reduced films are shown in Fig.2(d). Note for nonoptimized laser energy densities (1.0 and 1.3 J·cm
−2), the reduced film does not show well-defined (002) peak with weak (001) peak, suggesting the Ruddlesden−Popper phase dominated precursor film cannot be reduced to ideal infinite-layer structure. While for intermediate laser energy densities (1.1 and 1.2 J·cm
−2), the samples show well-defined diffraction peak and correct lattice parameters [inset of Fig.2(e)], indicating the formation of infinite-layer Nd
0.8Sr
0.2NiO
2 films. In addition, under the optimized laser energy density (1.1 J·cm
−2), the (001)/(002) peaks show stronger intensity, indicating better crystallinity. The RSM and AFM [inset of Fig.2(d)] indicate coherent strain and step-like surface structure, suggesting the obtained Nd
0.8Sr
0.2NiO
2 still maintains well-developed epitaxial relationship after CaH
2 reduction.
To understand the correlation between laser energy density and electrical properties, temperature dependent resistivity was measured down to 1.6 K for all the reduced films [Fig.3(a)]. Generally, all samples show a similar resistivity trend above 100 K, consistent with literature results [
4]. At non-optimized conditions (1.0 and 1.3 J·cm
−2), the films after CaH
2 reduction are dominated by the reduced Ruddlesden−Popper phase, thus exhibit metal−insulator transition with large resistance below 100 K. For intermediate condition (1.2 J·cm
−2), the reduced film shows a metallic-like behavior, but resistivity upturn still appears below 25 K, suggesting Kondo scattering [
4]. In contrast, the Nd
0.8Sr
0.2NiO
2 film grown at optimized condition (1.1 J·cm
−2) demonstrates clear superconducting transition below 13 K and achieves zero-resistance near 7 K. Fig.3(b) illustrates the temperature dependent Hall coefficient (
RH) for the superconducting Nd
0.8Sr
0.2NiO
2 film. At room temperature, the
RH is negative and is large in magnitude, while at low temperatures (~50 K),
RH shows a negative-to-positive sign conversion. These features are generally consistent with hole-doping of an electronic system with multiorbital contributions, suggesting mixed carrier contributions from electrons and holes [
3,
4].
Since the electrical data in Fig.3(a) are from four samples fabricated with identical film growth and reduction parameters (i.e., substrate temperature, oxygen pressure, film thickness, CaH2 reduction temperature and reduction time) except the only change in laser energy density, these data directly reveal the critical role of laser energy density for the formation of superconducting nickelates. Such a critical dependency leads to a super narrow “growth window” for perovskite precursor Nd0.8Sr0.2NiO3 film, which can be understood from fundamental laser ablation perspective. PLD is not a simple laser-induced thermal melting and vaporization process, but rather complex involving ultrafast ablation and subsequent laser-solids interaction. Typically, the ultrafast (30 ns) excimer laser pulse with a certain energy will be strongly absorbed by oxide target surface (within a depth of few tens of nm) in a confined area of few mm2 (laser spot). Rapid local heating (~fs) far exceeds thermal relaxation (~μs) leading to an extreme environment with strong deviation from thermodynamic equilibrium limit. As a result, dense plume of plasma condensate will be created, which will be further heated and accelerated by the incoming photons.
This plasma condensate is very important for the formation of epitaxial films with correct phase [
48-
50]. Although the shape and content of the plasma condensate can be affected by other growth parameters (e.g., oxygen pressure, growth temperature), the laser energy plays the most critical role, as it can directly affect the angular distribution of the elements in the plasma condensate, which is critical for the formation of the materials with correct crystallographic phase [
42,
51]. In the past decades, it is found that inequivalent angular distribution of the constitutes (in targets) can lead to significant deviation from the ideal film by forming secondary phase or crystal defects, as demonstrated in complex oxides LaAlO
3 [
52,
53], YBa
2Cu
3O
7−x [
54], SrTiO
3 [
55], SrZrO
3, [
56] (Pr,Sr)MnO
3 [
57], and Pb(Zr,Ti)O
3 [
58]. Likewise, the large atomic mass mismatches between Nd and Ni as well as the natural volatility of nickel oxide turn (Nd,Sr)NiO
3 to be a complex system with inequivalent angular distribution of plume condensate during PLD. This hidden plume dynamics issue, together with low formation energy of Ruddlesden−Popper secondary phase, makes the formation of perovskite (Nd,Sr)NiO
3 to be very challenging with rather narrow “growth window” and laser energy density is the most critical growth parameter. Furthermore, the nominal laser energy density might be slightly different from the actual laser energy density due to variation of uniformity/focus of the laser beam and laser-induced morphological modification at target surface. Thus, the optimized laser energy density for (Nd,Sr)NiO
3 film might show some variation for groups with different PLD systems. We propose that the details of the plume dynamics for (Nd,Sr)NiO
3 condensate and laser-target interaction during PLD process deserve further in-depth and systematical study by state-of-art
in-situ ultrafast imaging and spectroscopic techniques [
59,
60].
To further verify the aforementioned mechanism, we also change another growth condition, i.e., the substrate growth temperature, to explore its impact on the film growth. We chose the temperature variation of 570−650 °C (wide enough range to cover most of perovskite oxide epitaxy), while keeping laser energy density of 1.1 J·cm−2 and other parameters (oxygen partial pressure, film thickness, reduction temperature and time, etc.) constant. Fig.4(a) and (b) show the XRD scans for the Nd0.8Sr0.2NiO3 precursor films. Note the growth temperature window for perovskite film is relatively wide as (001)/(002) peaks for all the films remain at the same position. The XRD patterns of films after CaH2 reduction are shown in Fig.4(c) and (d). All films after reduction show the Nd0.8Sr0.2NiO2 infinite-layer structure. Note that the film grown at 650 °C shows weaker diffraction peaks after reduction, indicating poor film quality due to higher growth temperature. These data illustrate that Nd0.8Sr0.2NiO3 has a wide growth window on substrate temperature, but high temperature growth leads to poor quality for the reduced infinite-layer films. This can also be rationalized from the aforementioned plume dynamics perspective. After laser ablation, the laser-induced plume condensate escapes from the target surface and travels through towards substrate. The momentum and energy of the constitutes in the plume condensate are mainly determined by laser ablation and oxygen molecular scatter, and the substrate surface temperature only plays minimal effect. Therefore, the angular distribution of the plume condensate and the perovskite film growth will not be significantly impact by the substrate temperature.
Although the formation of perovskite and infinite-layer structure seems to be insensitive to the substrate temperature, it does significantly impact the electrical properties. Fig.5(a) shows the temperature dependent resistivity for the Nd
0.8Sr
0.2NiO
2 films (growth temperature of 570−650 °C). Except the highest growth temperature, all other films show superconducting transition with clear resistance drop. At the optimized temperature (600 °C), the Nd
0.8Sr
0.2NiO
2 film shows ideal zero-resistance while at higher or low temperatures (570 and 620 °C) it cannot achieve the zero-resistance state but still shows distinctive superconducting transition. However, at the highest temperature (650 °C), the film does not show any sign of superconducting transition, instead it shows a Kondo-like resistance upturn at 20 K, indicating metal-insulator transition. Since defective secondary phase, i.e., reduced Ruddlesden−Popper inclusions, can still exist in the infinite-layer Nd
0.8Sr
0.2NiO
2 film after reduction [
7,
33], the observed nonideal electrical properties at the non-optimized growth temperature can be attributed to these defect phase that compete with the ideal superconducting phase. This is also consistent with the XRD data in Fig.4(c) that only the film grown at the optimized temperature of 600 °C shows the highest diffraction intensity, while other films show suppressed diffraction intensity, indicating the existence of reduced Ruddlesden−Popper defective phase.
Finally, we show the superconducting Nd
0.8Sr
0.2NiO
2 film grown at the optimized laser energy density and substrate temperature demonstrate excellent environmental stability even without a SrTiO
3 caping layer. As we had reported that the superconducting state of Nd
0.8Sr
0.2NiO
2 film can be destroyed by water [
36], so it is important that storage condition should be water-free. Once the superconducting Nd
0.8Sr
0.2NiO
2 film is synthesized, it is just stored in a simple Teflon sample box in conventional dry air condition (at room temperature and 35% relative humidity) without glove box protection. Fig.5(b) shows temperature dependent resistivity of the Nd
0.8Sr
0.2NiO
2 film as a function of storage time. While the film shows a slight increase of the normal-state resistance above 14 K (due to the surface absorbance caused by air exposure), the superconducting state and
Tc are not affected for 512 days of continuous air exposure. This result clear demonstrates the superconducting Nd
0.8Sr
0.2NiO
2 film can be very robust which is critical for future electronics and device applications.
3 Conclusion
We have systematically investigated the film growth process towards achieving superconducting infinite-layer nickelate Nd0.8Sr0.2NiO2 epitaxial film by PLD. We find the challenges for obtaining superconducting nickelates involve two key aspects. One is to obtain infinite-layer crystal structure and the other is to achieve ideal superconducting state. Due to relatively low formation energy of the Ruddlesden−Popper defective phase, it competes the perovskite phase Nd0.8Sr0.2NiO3. Inequivalent angular distributions of plume condensate for Nd0.8Sr0.2NiO3 target during laser ablation making laser energy density to be very critical to stabilize phase-pure perovskite precursor, which is prerequisite for the formation of infinite-layer structure. We also found both laser energy density and substrate temperature are important for obtaining ideal superconducting state. At the optimized growth conditions, the superconducting Nd0.8Sr0.2NiO2 film can be obtained sustainably. We also demonstrate that superconducting Nd0.8Sr0.2NiO2 film can be very stable in dry air conditions up to 512 days. These results provide important information for the development of the field and for promoting further research on fundamental science and device applications of nickelate superconductors.
4 Methods
Film synthesis. The perovskite Nd
0.8Sr
0.2NiO
3 epitaxial films were grown on single-crystal SrTiO
3 (001) substrates using pulsed laser deposition (PLD, Demcon TSST) with a KrF excimer laser (
λ = 248 nm). As-received SrTiO
3 (001) substrates (from HeFeiKeJing Mater. Ltd.) were etched by HF and annealed at 1050 °C for 2 h to achieve atomic-flat TiO
2-termination [
61,
62]. The obtained 5 mm × 5 mm size TiO
2-terminated SrTiO
3 (001) substrates were then pre-annealed in PLD chamber for 1 h at 640 °C and oxygen partial pressure about 5 × 10
−6 Torr to obtain sharp steps before film growth. In this study, we performed ablation using uniform rectangular laser spots size of 1.0 mm × 3 mm, which were formed by aperture imaging. The polycrystalline target Nd
0.8Sr
0.2NiO
3 (synthesized by solid-state reaction method from commercial oxide powders of Nd
2O
3, SrCO
3, Ni
2O
3 with stoichiometric ratios) were ablated using laser energy density in the range of 1.0−1.3 J·cm
−2 with a frequency of 4 Hz. High-pressure
in-situ reflective high-energy electron diffraction (RHEED, Staib Instruments) was used to monitor the film growth process. In addition, the growth temperature of the perovskite precursor films was set to 550−650 °C and the oxygen partial pressure was kept at 200 mTorr. The temperature rises and falls rate of films is strictly controlled at 10 K/min. All samples mentioned in the text were grown without SrTiO
3 capping layer [
36].
Topotactic reduction. The as-grown film is loosely wrapped in aluminum foil to avoid direct contact with the reducing agent, placed in a Pyrex glass tube containing 0.1 g CaH
2 powder and vacuum sealed (pressure < 0.1 mTorr). The topotactic reduction was performed at the optimized conditions based on our previous studies [
36], i.e., at the temperature of 300 °C and reduction time period of 2 h with the rate of temperature rise and fall fixed at 10 °C·min
−1. For air stability test, the as-synthesis superconducting film was stored in vacuum condition right after the structural and electrical measurements.
Structural characterization and transport measurements. The crystal structures of both the as-grown and reduced films were analyzed by high-resolution X-ray diffraction (XRD) and reciprocal space mapping (RSM) using a Bruker D8 Discover diffractometer with a LYNXEYE-2 line detector. A four-bounce double-crystal Ge (220) and a Cu X-ray mirror were placed in the incident beam path to generate monochromatic Cu-Kα X-rays with a beam divergence of 12 arc seconds and the diffracted beam was filtered through a 0.09° parallel plate collimator. Film thickness was determined from X-ray reflectivity (XRR) measurement with an air/film/substrate model simulation. The surface morphology and step-terrace structure of the grown nickelate films were characterized by atomic force microscopy (AFM, Park system, NX10). The temperature dependent (down to 1.6 K) electrical resistivity and Hall effect (up to 9 T magnetic field) of the films were measured by the cryogenic magnet-free system (CFMS, Cryogenics Ltd.) using the four-probe method and the Van der Pauw geometry. Electrical contact of Al wires was bond on deposited Au pads on the film surface by ultrasonic wire bonder.