(Translated by https://www.hiragana.jp/)
Supplemental Materials for: Synthesis of thin film infinite-layer nickelates by atomic hydrogen reduction: clarifying the role of the capping layer
License: CC BY-NC-SA 4.0
arXiv:2401.07129v1 [cond-mat.mtrl-sci] 13 Jan 2024

Supplemental Materials for: Synthesis of thin film infinite-layer nickelates by atomic hydrogen reduction: clarifying the role of the capping layer

C. T. Parzyck Laboratory of Atomic and Solid State Physics, Department of Physics, Cornell University, Ithaca, NY 14853, USA    V. Anil Laboratory of Atomic and Solid State Physics, Department of Physics, Cornell University, Ithaca, NY 14853, USA    Y. Wu Laboratory of Atomic and Solid State Physics, Department of Physics, Cornell University, Ithaca, NY 14853, USA    B. H. Goodge School of Applied and Engineering Physics, Cornell University, Ithaca, NY 14853, USA Kavli Institute at Cornell for Nanoscale Science, Cornell University, Ithaca, NY 14853, USA    M. Roddy Laboratory of Atomic and Solid State Physics, Department of Physics, Cornell University, Ithaca, NY 14853, USA    L. F. Kourkoutis School of Applied and Engineering Physics, Cornell University, Ithaca, NY 14853, USA Kavli Institute at Cornell for Nanoscale Science, Cornell University, Ithaca, NY 14853, USA    D. G. Schlom Kavli Institute at Cornell for Nanoscale Science, Cornell University, Ithaca, NY 14853, USA Department of Materials Science and Engineering, Cornell University, Ithaca, NY 14853, USA Leibniz-Institut für Kristallzüchtung, Max-Born-Straße 2, 12489 Berlin, Germany    K. M. Shen Laboratory of Atomic and Solid State Physics, Department of Physics, Cornell University, Ithaca, NY 14853, USA Kavli Institute at Cornell for Nanoscale Science, Cornell University, Ithaca, NY 14853, USA Institut de Ciència de Materials de Barcelona (ICMAB-CSIC), Campus UAB Bellaterra 08193, Spain
(January 13, 2024)

I Perovskite Nickelates

The stoichiometric calibration of nickelate films is of paramount importance to the growth of high-quality NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. As the binary oxide constituents, Nd22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT and NiO have negligible volatility, an absorption-controlled approach in molecular beam epitaxy (MBE) is not possible – excess deposited material either incorporates or forms precipitates–both of which negatively impact the film quality. Studies of the pulsed laser deposition (PLD) and MBE growth of nickelates indicate that deviation from a cation ratio of 1:1 has strong effects on the electrical properties of NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT films[1, 2, 3, 4] and a strong influence on the quality of subsequently reduced films – particularly the presence of Ruddlesden-Popper (RP) intergrowths [5, 6, 7]. In PLD growth, careful control of the film stoichiometry can be achieved through control of the laser fluence and source material (by repolishing the target surface to combat laser-induced changes [2], for instance). In MBE, on the other hand, the key is to establish the elemental source fluxes with a high degree of accuracy. We find that the typical accuracy of a quartz crystal microbalance (QCM), 10-15%, is not sufficient to reproducibly grow high-quality nickelate films; we follow instead a two-step calibration procedure detailed here.

Refer to caption
Figure S1: RHEED Oscillations of the binary oxides NiO/MgO (100) and Nd22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT/YSZ (111) used for calibration. (a) RHEED specular intensity of a NiO film during growth. (B) Fourier transform of the specular intensity showing a sharp peak at the monolayer frequency. (c-d) Post-growth RHEED images of the NiO film. (e-h) Same for a representative Nd22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT film.

I.1 RHEED Oscillation Calibration

Initial flux measurements are obtained by monitoring RHEED oscillations during the growths of Nd22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT on (ZrO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT)0.9050.905{}_{0.905}start_FLOATSUBSCRIPT 0.905 end_FLOATSUBSCRIPT(Y22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT)0.0950.095{}_{0.095}start_FLOATSUBSCRIPT 0.095 end_FLOATSUBSCRIPT (111) and NiO on MgO (100) using the parameters outlined in Sun et al. [8]. In Fig. S1 we detail the RHEED oscillation measurements of NiO and Nd22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT in (a-d) and (e-h), respectively. While both compounds grow over a very wide range of conditions (ozone flux, substrate temperature) we find the optimal conditions for RHEED oscillations of NiO are around 450 {}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTC  (following an 800 {}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTC  vacuum anneal to clean the MgO surface). Conversely, it is preferable to grow Nd22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT at higher temperatures, 970similar-toabsent970\sim 970∼ 970 {}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTC to stabilize the hexagonal structure; though often the first few layers will take on a bixbyite structure, evidenced by a noticeably different oscillation frequency as seen in (e). By monitoring the RHEED specular intensity in (a) and (e) and performing a Fourier transform, (b) and (f), the monolayer formation time can be easily extracted and converted into an elemental flux.

Refer to caption
Figure S2: Shutter timing diagram illustrating the sequential deposition of NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT thin films. The shutters for the neodymium and nickel sources are opened in sequence for time periods τたうNdsubscript𝜏Nd\tau_{\textrm{Nd}}italic_τたう start_POSTSUBSCRIPT Nd end_POSTSUBSCRIPT and τたうNisubscript𝜏Ni\tau_{\textrm{Ni}}italic_τたう start_POSTSUBSCRIPT Ni end_POSTSUBSCRIPT, respectively. Between each cycle the sample is ozone annealed with the shutters closed for a short period of τたうAnn.=5subscript𝜏Ann.5\tau_{\textrm{Ann.}}=5italic_τたう start_POSTSUBSCRIPT Ann. end_POSTSUBSCRIPT = 5 seconds to promote full oxidation of the nickel layer, as in Ref. [9].
Refer to caption
Figure S3: X-ray diffraction calibration data for 20 u.c. thick films of NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT on (001) SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. c𝑐citalic_c-axis lattice constant, determined from the 002 peak position, as a function of the nominal cation ratio Nd:Ni. Slopes of typical correction factors employed in the optimization procedure are labeled.

I.2 XRD and XRR Calibration

While the aforementioned calibration method greatly improves upon the basic calibration using a QCM, we find that its accuracy is typically limited to the the few percent scale for each source and as such requires further refinement. This is accomplished using the method described by Li et al. in Ref. 6: the c𝑐citalic_c-axis lattice constant (in our case the 002 peak position) is used as a proxy for non-stoichiometery with the minimal value of the lattice constant corresponding to a stoichiometric cation ratio of Nd:Ni = 1:1. Our repetition of this procedure produces much the same qualitative characteristics as in Ref. [6], although with a slightly lower minimal lattice constant of c=3.736𝑐3.736c=3.736italic_c = 3.736 Å, and is detailed in Fig. S3. To construct this diagram we start by growing a sequence of films with differing shutter times, τたうNd,Nisubscript𝜏Nd,Ni\tau_{\textrm{Nd,Ni}}italic_τたう start_POSTSUBSCRIPT Nd,Ni end_POSTSUBSCRIPT, and record the ratio τたうNd/τたうNisubscript𝜏Ndsubscript𝜏Ni\nicefrac{{\tau_{\textrm{Nd}}}}{{\tau_{\textrm{Ni}}}}/ start_ARG italic_τたう start_POSTSUBSCRIPT Nd end_POSTSUBSCRIPT end_ARG start_ARG italic_τたう start_POSTSUBSCRIPT Ni end_POSTSUBSCRIPT end_ARG. The shutter times τたうNdisubscript𝜏Ndi\tau_{\textrm{Ndi}}italic_τたう start_POSTSUBSCRIPT Ndi end_POSTSUBSCRIPT  and τたうNisubscript𝜏Ni\tau_{\textrm{Ni}}italic_τたう start_POSTSUBSCRIPT Ni end_POSTSUBSCRIPT  correspond to our best estimate of the time to deposit a single monolayer of NdO and NiO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT, respectively. A timing diagram showing the sequential opening of the Nd and Ni shutters in a constant background partial pressure of O33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT with a short, 5 second, anneal between cycles is provided in Fig. S2. While the flux ratio ΦふぁいNd/ΦふぁいNisubscriptΦふぁい𝑁𝑑subscriptΦふぁい𝑁𝑖\nicefrac{{\Phi_{Nd}}}{{\Phi_{Ni}}}/ start_ARG roman_Φふぁい start_POSTSUBSCRIPT italic_N italic_d end_POSTSUBSCRIPT end_ARG start_ARG roman_Φふぁい start_POSTSUBSCRIPT italic_N italic_i end_POSTSUBSCRIPT end_ARG is not known a priori, once the growth sequence converges to near the minimal value this ratio can be calculated from

1=ΦふぁいNdΦふぁいNi×τたうNd,fτたうNi,f.1subscriptΦふぁいNdsubscriptΦふぁいNisubscript𝜏Nd,fsubscript𝜏Ni,f1=\frac{\Phi_{\textrm{Nd}}}{\Phi_{\textrm{Ni}}}\times\frac{\tau_{\textrm{Nd,f}% }}{\tau_{\textrm{Ni,f}}}.1 = divide start_ARG roman_Φふぁい start_POSTSUBSCRIPT Nd end_POSTSUBSCRIPT end_ARG start_ARG roman_Φふぁい start_POSTSUBSCRIPT Ni end_POSTSUBSCRIPT end_ARG × divide start_ARG italic_τたう start_POSTSUBSCRIPT Nd,f end_POSTSUBSCRIPT end_ARG start_ARG italic_τたう start_POSTSUBSCRIPT Ni,f end_POSTSUBSCRIPT end_ARG .

and applied to the whole sequence to determine the Nd:Ni ratio for all films. Figure S3 is an aggregation of many such growth runs. The scatter increases as we depart from Nd:Ni = 1:1, which is a consequence of source drift on the above procedure. Though modern effusion cells provide excellent flux stability, often reducing flux drift to less then 1% per hour, the difference between τたうNd/τたうNisubscript𝜏Ndsubscript𝜏Ni\nicefrac{{\tau_{\textrm{Nd}}}}{{\tau_{\textrm{Ni}}}}/ start_ARG italic_τたう start_POSTSUBSCRIPT Nd end_POSTSUBSCRIPT end_ARG start_ARG italic_τたう start_POSTSUBSCRIPT Ni end_POSTSUBSCRIPT end_ARG at the start of the run and the end (when 1:1 is reached) means that estimation of Nd:Ni for the first samples in the sequence (Nd:Ni 1much-greater-thanabsent1\gg 1≫ 1) is less accurate than at the end (Nd:Ni 1similar-toabsent1\sim 1∼ 1). This estimated cone of uncertainty is illustrated by the shaded regions. Despite this uncertainty in the non-stoichiometry of films well away from 1:1, in the region near the minimum lattice constant discernible changes are still observed down to single percent changes in the shutter times, evidencing that this method can be used to hone the non-stoichiometry down to the percent level or less.

A typical optimization routine at the start of a growth campaign consists of starting with the RHEED oscillation calibration of one or both of Nd and Ni. Then the starting shutter times are calculated to begin with a few percent of excess Neodymium to bias the initial growth towards Nd:Ni >1absent1>1> 1; this is to ensure that the direction of the first step is known a priori (i.e., reducing the Nd shutter time) and to avoid venturing into the Ni-rich side of the diagram. We find that on the Ni-rich side the relationship between d002subscript𝑑002d_{002}italic_d start_POSTSUBSCRIPT 002 end_POSTSUBSCRIPT and the shutter times is less repeatable. We attribute this to the propensity of excess Ni to precipitate out of the film in the form of NiO, rather then readily incorporating like neodymium does. The flux ratio can then be ‘walked-in’ to 1:1 by progressive adjustment of the shutter times, growth, and XRD measurement. If the starting Nd excess is greater then 10% or so an ‘aggressive’ adjustment factor of 0.2 to 0.3 Å/% can be used in tuning the shutter times between subsequent films; as Nd excess is further reduced more conservative values of 0.3 to 0.47 Å/% are used to avoid accidentally jumping over onto the Ni-rich side. Once the Nd:Ni ratio is within 5-7% of the 1:1 stoichiometric ratio, the total film thickness can also be fine-tuned (by changing τたうNdsubscript𝜏Nd\tau_{\textrm{Nd}}italic_τたう start_POSTSUBSCRIPT Nd end_POSTSUBSCRIPT and τたうNisubscript𝜏Ni\tau_{\textrm{Ni}}italic_τたう start_POSTSUBSCRIPT Ni end_POSTSUBSCRIPT by the same fraction) to match the desired number of unit cells, typically 20, based on x-ray reflectivity (XRR) measurements of the NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT film thickness.

I.3 Effects of Non-Stoichiometry on Transport

Refer to caption
Figure S4: Effects of cation nhon-stoichiometry on 20 u.c. thick NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT thin films grown on (001)-oriented SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. Resistivity measurements of films of varying non-stoichiometry from slightly nickel rich to highly neodymium rich. Post-growth RHEED images, viewed along the [110]pcpc{}_{\textrm{pc}}start_FLOATSUBSCRIPT pc end_FLOATSUBSCRIPT azimuth, of each of the films are pictured on the right. All images use a linear intensity scale.

Here we describe the effects of non-stoichiometry on the transport properties of NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT films grown on (001) SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. The temperature dependent resistivity of a sequence of samples of varying lattice constants (and estimated non-stoichiometry based on Fig. S3) are presented in Fig. S4. The film with the lowest lattice constant has a low room-temperature resistivity of 200μみゅーΩおめがsimilar-toabsent200𝜇Ωおめが\sim 200\mu\Omega∼ 200 italic_μみゅー roman_Ωおめが-cm, and a sharp metal-to-insulator transition (MIT) with a jump of nearly two decades. The structural quality is also evident in the accompanying RHEED image, where sharp diffraction peaks are visible on the primary and half-order streaks. As the Nd:Ni ratio is increased a clear increase in the resistivity is visible (up to an order of magnitude), along with a broadening of the MIT and widening of the hysteresis loop. At the same time the sharp diffraction peaks fade and the RHEED streaks become ‘peanut shaped’ – likely due to an increase in the film roughness. We note a marked difference between Nd-rich films under tensile strain when compared to prior measurements under compressive strain [1, 2] where excess Nd drives the films metallic. As previously noted in studies of non-stoichiometry on SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT [3] the MIT temperature increases with increasing Nd content, and the MIT is not suppressed even up to very high levels (Nd:Ni 1.25similar-toabsent1.25\sim 1.25∼ 1.25). Finally, on the nickel-rich side of the diagram broadening of the MIT and an overall increase of the resistivity are observed. These results indicate that a stoichiometric film on SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT is the one with a minimum lattice constant, maximum transition sharpness, and minimum resistivity of the metallic state.

I.4 Additional Aging Effect Data

In Fig. S5 (unshifted) raw data corresponding to the sample aging study presented in Fig. 2 of the main text. Figs. S5(a) and S5(b) are reproduced from the main text for reference and Figs. S5(c) and S5(d) show data for the sample that was measured at much shorter intervals following growth. It is evident that in both cases a gradual increase of the film resistivity is present over time, with the room temperature values increasing by a factor of 4-5 over the course of 3.5similar-toabsent3.5\sim 3.5∼ 3.5  weeks while stored in a dry, N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT purged desiccator.

Refer to caption
Figure S5: Aging effects on the transport properties of stoichiometric NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT films grown on SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. (a) Resistivity of a single NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT film measured periodically following growth, reproduced from the main text. (b) Logarithmic derivative of the transport curves in panel (a). (c) Resistivity of another film measured at a higher frequency over the same period of time as the one pictured in panel (a). (d) Logarithmic derivative of the resistivity curves in (c).

II Infinite-Layer Nickelates

II.1 AFM and XRD of Optimized Films

Refer to caption
Figure S6: Measurements of the surface quality of NdNiO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT films reduced using atomic hydrogen. (a) RHEED images taken along the [110]pc𝑝𝑐{}_{pc}start_FLOATSUBSCRIPT italic_p italic_c end_FLOATSUBSCRIPT azimuth of an as grown 20 u.c. thick NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT film, the same film after capping with 2 u.c. of SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT, and after atomic hydrogen reduction. (b) The same as in (a) but with a 3 u.c. thick SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT cap. (c) AFM image of a 20 u.c. thick NdNiO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT film capped with 2 u.c. of SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT and reduced with atomic hydrogen; the extracted terrace-averaged RMS roughness is 124±10plus-or-minus12410124\pm 10124 ± 10 pm. (d) Zoom in on the stepped terraces. (e) Line profiles along (black) and across (red) the terraces showing preservation of the substrate’s atomic steps. RHEED images in (a) and (b) use a linear intensity scale.
Refer to caption
Figure S7: Diffraction intensity along the specular direction for an uncapped 20 u.c. thick NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT film and a 3 u.c. SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT capped NdNiO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT film grown on SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT showing the 00L00𝐿00L00 italic_L reflections. Nelson-Riley analysis of the infinite-layer film peaks gives a lattice constant of c=3.286𝑐3.286c=3.286italic_c = 3.286 Å. (*) indicates reflections from the (001) SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT substrate.

In this section we present additional measurements of optimized, reduced NdNiO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT films on SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. In Fig. S6 we show RHEED images and AFM topography of a reduced infinite-layer films. Figs. S6(a) and S6(b) present RHEED images of additional 20 u.c. thick NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT films after growth, after capping, and after atomic hydrogen reduction. In both reduced films, reduced using the optimized three-step procedure, the preservation of the RHEED streaks indicates the film surface remains ordered. The RHEED images of both films are qualitatively similar, though the 3 u.c. capped film appears to have a slightly lower incoherent background then that of the film with the thinner cap. Figures S6(c)-(e) show AFM measurements of a 20 u.c. thick NdNiO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT film capped with 2 u.c. of SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT and reduced with atomic hydrogen. The reduced film shows a stepped terrace structure analogous to the atomically flat substrate on which it was grown, with a low surface roughness of 124similar-toabsent124\sim 124∼ 124 pm. This is similar to films reduced using the solid-state aluminum reduction process introduced in Ref. 10. In Fig. S7 we show larger area θしーた2θしーた𝜃2𝜃\theta-2\thetaitalic_θしーた - 2 italic_θしーた scans of a perovskite, NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT, film and an optimized, reduced infinite-layer, NdNiO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT, film which cover all of the 00L00𝐿00L00 italic_L peaks observable with Cu Kαあるふぁ𝛼\alphaitalic_αあるふぁ x-rays. In the infinite-layer film, all four reflections are visible, with the 004004004004 peak sitting at the end of the range accessible with our goniometer.

Refer to caption
Figure S8: Conditions used for H and H22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT reduction of films during experiments exploring the effect of the SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT cap on reduction. (a) Replica of Fig. 7(a) from the main text with samples labeled. (b) Cartoon of a generic three-step (τたう20subscript𝜏20\tau_{2}\neq 0italic_τたう start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≠ 0) or two-step (τたう2=0subscript𝜏20\tau_{2}=0italic_τたう start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0) reduction. (c) Table of reduction conditions used in the preparation of samples in (a) as well as in Figs. S9, S10, and S11.

II.2 Reduction Effects on the Film Surface

In this section we present additional data concerning the effects of molecular and atomic hydrogen on the film surface. In Fig. S8 we present the details of the reduction program used for samples in Fig. 7 (samples A”-G”) of the main text as well as those presented later in this section (samples I”,J”). Of these, several samples were subjected to several sequential single-step reduction cycles (τたう2=0subscript𝜏20\tau_{2}=0italic_τたう start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0) with variable times, performing RHEED measurements between cycles. These include sample C”, which was reduced in four increments of 15, 15, 15, and 45 minutes (90 min total), sample I” in four increments of 15, 30, 30, and 15 minutes (90 min total), and J” in three increments of 15, 15, and 30 minutes (60 min total). Fluxes are quoted in atoms/cm22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT/sec for samples reduced with atomic hydrogen and in molecules/cm22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT/sec for those reduced with molecular hydrogen only. In all cases a consistent hydrogen flow rate of 0.72 sccm was used (giving a chamber pressure of 9.5×106similar-toabsent9.5superscript106\sim 9.5\times 10^{-6}∼ 9.5 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT  torr) and the atomic hydrogen flux was adjusted by tuning the cracker temperature.

Refer to caption
Figure S9: Molecular hydrogen reduction of a 20 u.c. thick NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT / SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT film capped with 2 u.c. of SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. (a) RHEED images along the [10] azimuth of a film exposed to H22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT, at a temperature of 300 {}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTC, for repeated intervals accumulating to 90 minutes total. (b) XRD measurements of the same film before and after the 90 minutes of total reduction time. (*) indicates reflections from the (001) oriented SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT substrate.
Refer to caption
Figure S10: Effects of atomic hydrogen reduction on an uncapped, 20 u.c. thick NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT / SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT film. (a) RHEED images along the [100]pcpc{}_{\textrm{pc}}start_FLOATSUBSCRIPT pc end_FLOATSUBSCRIPT and [110]pcpc{}_{\textrm{pc}}start_FLOATSUBSCRIPT pc end_FLOATSUBSCRIPT azimuths of a uncapped film; images are taken of the as grown film surface as well as after 15, 30, and 60 minutes of atomic hydrogen exposure time at a sample temperature of 300 {}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTC. (b) XRD measurements of the same film before and after the 60 minute exposure. (*) indicates reflections from the (001) oriented SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT substrate.

In Figs. S9 and S10 we detail two additional progressive annealing experiments. Figure S9 shows a sample with a 2 u.c. thick SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT cap annealed in molecular hydrogen. As can be seen in the RHEED images, the molecular hydrogen has little-to-no impact on the SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT surface, even for long exposure times. Similar to the uncapped sample shown in the main text, molecular hydrogen is insufficient to produce the infinite-layer phase (under these conditions) and instead an oxygen-deficient perovskite phase is visible in the XRD, with out-of-plane lattice paramter c3.69similar-to𝑐3.69c\sim 3.69italic_c ∼ 3.69 Å, following 90 minutes of exposure. Figure S10 shows a sample that was exposed to atomic hydrogen in progressive steps. Similar to the sample presented in the main text, the RHEED images indicate a rapid degradation of the film surface and the presence of other phases (faint rings and additional spots). Even after a relatively long exposure, 60 minutes, the majority of the film remains an oxygen-deficient perovskite (c3.69similar-to𝑐3.69c\sim 3.69italic_c ∼ 3.69 Å) without any indication of formation of the infinite-layer phase, evidenced by the XRD patterns in Fig. S10(b).

Refer to caption
Figure S11: Atomic hydrogen reduction of 20 u.c. thick NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT / SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT films capped with 1.2 nm of amorphous SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT. (a) XRD of nickelate films as grown (light) and after capping and reduction (dark); capping layers were grown using either simultaneous codeposition of Sr and Ti (blue) or sequential shuttered deposition (purple) at room temperature. 002 peaks of the reduced films are fit with an N𝑁Nitalic_N-slit interference pattern (teal, pink) to extract the coherent thickness. (b) Low-angle XRR of one of the samples before and after capping/reduction (dots) and corresponding best fits (lines) using a three slab model.

Finally, we discuss the effects of amorphous capping layers on the reduction of 20 u.c. NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT/SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT films using atomic hydrogen. X-ray diffraction measurements showing the 001 and 002 peak regions for two films with amorphous caps are provided in Fig. S11(a). For both films, the capping layer was deposited in a background pressure of 2×1062superscript1062\times 10^{-6}2 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT  torr of distilled O33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT  with the sample at room temperature. For sample H” the capping layer was deposited with Sr and Ti shutters opened simultaneously (codeposition) and for sample K” the shutters were opened sequentially (shuttered deposition); in both cases no streaks, spots, or rings were visible in RHEED after deposition of the SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT layers. The thickness of both capping layers was 1.2 nm – equivalent to 3 u.c. of SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT  – and the subsequent reductions, using the same conditions, appear to progress similarly based on the XRD measurements in Fig. S11(a). The obvious difference in XRD patterns between those samples reduced (using atomic hydrogen) with an amorphous cap (H” and K”) versus those reduced with no cap (D” and J”) illustrate the importance of the overlayer in facilitating the reduction even in the absence of epitaxial stabilization.

We do note, however, that the quality of the amorphous SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT  capped samples are somewhat reduced compared to those with a crystalline cap. Fitting of the XRD patterns of both the reduced films with an N𝑁Nitalic_N-slit interference pattern,

I1Nsin2(Naq/2)sin2(aq/2),proportional-to𝐼1𝑁superscript2𝑁𝑎𝑞2superscript2𝑎𝑞2I\propto\frac{1}{N}\frac{\sin^{2}(Naq/2)}{\sin^{2}(aq/2)},italic_I ∝ divide start_ARG 1 end_ARG start_ARG italic_N end_ARG divide start_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_N italic_a italic_q / 2 ) end_ARG start_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_a italic_q / 2 ) end_ARG ,

gives a Scherrer thickness of only 4.43 nm, i.e. N13.5similar-to𝑁13.5N\sim 13.5italic_N ∼ 13.5, indicating that the coherent crystalline region does not extend across the entirety of the film thickness. Additionally, the x-ray reflectivity, Fig. S11(b), of the reduced film cannot be accurately fit using a three-slab model (SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT/NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT/SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT) as was done for the crystalline film in Fig. 5(d) of the main text. The best-fit (green) for this reflectivity curve gives a roughness for the NdNiO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT layer, 3.2 nm (compared to that of the as grown film, 0.48 nm), and a thickness of 5.6 nm (intermediate to the Scherrer thickness of 4.4 nm and the expected thickness of 6.5 nm). These results are consistent with the capping layer providing some measure of epitaxial support during the reduction process, which is absent in those films with an amorphous cap, and a fraction of the film either forms a defect phase or decomposes – as reported previously in CaH22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT reduced films [5]. We cannot rule out, however, that the observed differences between the crystalline and amorphous SrTiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT capped films result from differences in transport of the reactants (e.g. H or O22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT) through the different matrices. We emphasize, however, that the role of the capping layer, at least in atomic hydrogen reduction experiments, clearly goes beyond that of epitaxial stabilization. In the XRD patterns of uncapped films, D” and J”, the effect of the reduction is not to cause decomposition of the entire thickness of the film, but rather to form a polycrystalline scale on the surface and to hamper further reduction – as evidenced by the minor 002 peak shifts observed in XRD. This is in contrast with the more modest observed differences between crystalline capped (Samples F”, G”) and amorphous capped (Samples H”, K”) films which all exhibit similar 002 peak positions.

References

  • Breckenfeld et al. [2014] E. Breckenfeld, Z. Chen, A. R. Damodaran, and L. W. Martin, Effects of nonequilibrium growth, nonstoichiometry, and film orientation on the metal-to-insulator transition in NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT thin films, ACS Applied Materials and Interfaces 6, 22436 (2014).
  • Preziosi et al. [2017] D. Preziosi, A. Sander, A. Barthélémy, and M. Bibes, Reproducibility and off-stoichiometry issues in nickelate thin films grown by pulsed laser deposition, AIP Advances 7, 015210 (2017).
  • Yamanaka et al. [2019] T. Yamanaka, A. N. Hattori, L. N. Pamasi, S. Takemoto, K. Hattori, H. Daimon, K. Sato, and H. Tanaka, Effects of Off-Stoichiometry in the Epitaxial NdNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT Film on the Suppression of Its Metal-Insulator-Transition Properties, ACS Applied Electronic Materials 1, 2678 (2019).
  • Pan et al. [2022] G. A. Pan, Q. Song, D. Ferenc Segedin, M. C. Jung, H. El-Sherif, E. E. Fleck, B. H. Goodge, S. Doyle, D. Córdova Carrizales, A. T. N’Diaye, P. Shafer, H. Paik, L. F. Kourkoutis, I. El Baggari, A. S. Botana, C. M. Brooks, and J. A. Mundy, Synthesis and electronic properties of Ndn+1𝑛1{}_{n+1}start_FLOATSUBSCRIPT italic_n + 1 end_FLOATSUBSCRIPTNin𝑛{}_{n}start_FLOATSUBSCRIPT italic_n end_FLOATSUBSCRIPTO3n+13𝑛1{}_{3n+1}start_FLOATSUBSCRIPT 3 italic_n + 1 end_FLOATSUBSCRIPT Ruddlesden-Popper nickelate thin films, Physical Review Materials 6, 055003 (2022).
  • Lee et al. [2020] 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 Materials 8, 041107 (2020).
  • Li et al. [2021] Y. Li, W. Sun, J. Yang, X. Cai, W. Guo, Z. Gu, Y. Zhu, and Y. Nie, Impact of Cation Stoichiometry on the Crystalline Structure and Superconductivity in Nickelates, Frontiers in Physics 9, 1 (2021).
  • Lee et al. [2023] K. Lee, B. Y. Wang, M. Osada, B. H. Goodge, T. C. Wang, Y. Lee, S. Harvey, W. J. Kim, Y. Yu, C. Murthy, S. Raghu, L. F. Kourkoutis, and H. Y. Hwang, Linear-in-temperature resistivity for optimally superconducting (Nd,Sr)NiO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTNature 619, 288 (2023).
  • Sun et al. [2022] J. Sun, C. T. Parzyck, J. H. Lee, C. M. Brooks, L. F. Kourkoutis, X. Ke, R. Misra, J. Schubert, F. V. Hensling, M. R. Barone, Z. Wang, M. E. Holtz, N. J. Schreiber, Q. Song, H. Paik, T. Heeg, D. A. Muller, K. M. Shen, and D. G. Schlom, Canonical approach to cation flux calibration in oxide molecular-beam epitaxy, Physical Review Materials 6, 033802 (2022).
  • King et al. [2014] P. D. C. King, H. I. Wei, Y. F. Nie, M. Uchida, C. Adamo, S. Zhu, X. He, I. Božović, D. G. Schlom, and K. M. Shen, Atomic-scale control of competing electronic phases in ultrathin LaNiO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPTNature Nanotechnology 9, 443 (2014).
  • Wei et al. [2023] W. Wei, K. Shin, H. Hong, Y. Shin, A. S. Thind, Y. Yang, R. F. Klie, F. J. Walker, and C. H. Ahn, Solid state reduction of nickelate thin films, Physical Review Materials 7, 013802 (2023).