The discovery of analog LixNbO2 memristors revealed a promising new memristive mechanism wherein the diffusion of Li+ rather than O2− ions enables precise control of the resistive states. However, directly correlating lithium concentration with changes to the electronic structure in active layers remains a challenge and is required to truly understand the underlying physics. Chemically delithiated single crystals of LiNbO2 present a model system for correlating lithium variation with spectroscopic signatures from operando soft x-ray spectroscopy studies of device active layers. Using electronic structure modeling of the x-ray spectroscopy of LixNbO2 single crystals, we demonstrate that the intrinsic memristive behavior in LixNbO2 active layers results from field-induced degenerate p-type doping. We show that electrical operation of LixNbO2-based memristors is viable even at marginal Li deficiency and that the analog memristive switching occurs well before the system is fully metallic. This study serves as a benchmark for material synthesis and characterization of future LixNbO2-based memristor devices and suggests that valence change switching is a scalable alternative that circumvents the electroforming typically required for filamentary-based memristors.

Functional oxide memristors have the potential to revolutionize neuromorphic computing, which aims to mimic the operation of biological brains using artificial circuits.1,2 While neuromorphic systems can be implemented with traditional CMOS circuitry, this requires a large number of conventional transistors, resulting in significant power consumption and scalability issues.3–5 By contrast, truly biomimetic circuits would lead to scalable, low-power processors capable of hardware-level autonomous learning, power-efficient image recognition, and other exciting possibilities.6,7 Neuristor circuits based on functional oxide memristors have the potential to enable these truly biomimetic circuits. However, the resistive switching of these memristors is typically attributed to a complex combination of processes (e.g., redox reactions, ionic transport, and phase changes.),8 which are highly dependent on device architecture and are not yet fully understood.9,10

Lithium niobite (LixNbO2), which has been previously studied due to properties such as superconductivity,11 has recently shown great potential for memristive applications.12,13 Whereas traditional filamentary devices typically require the migration of O2− ions to form narrow conductive channels and access discrete resistive states, the resistive states of LixNbO2 are analog in nature and are thought to be modulated by a more uniform diffusion of Li+ ions throughout the active layer.14 This more uniform ion diffusion mechanism may help to overcome some of the major disadvantages of traditional memristors, such as the need for an initial preforming/electroforming process to breakdown or otherwise alter the as-deposited material before the device can function.15 In addition, many sources of device failure, such as thermal stress from conduction through nanoscale pathways, nonuniform O2− ion clustering, or other failure modes,16 may be completely absent in these LiNbO2-based devices. Further investigation of these LiNbO2 memristors is therefore needed to fully characterize this promising memristive mechanism.

It has been proposed that the electronic properties of LixNbO2 are highly dependent on the Li+ content and that a bias-induced Li-gradient could enable precise control of the resistive state.17 This bias-induced Li gradient was previously investigated across an annular LiNbO2 device using spatially resolved in situ oxygen K-edge x-ray absorption spectroscopy (XAS),14 revealing gradual changes in the spectral lineshape across the active layer interpreted as lithium concentration variations. Recent advances in x-ray absorption simulations of the O K-edge18 have now presented an opportunity to confirm the predicted evolution of LixNbO2 within these devices and accurately describe their operational mechanism.

In this paper, we benchmark the intrinsic material properties responsible for the analog memristive behavior of LixNbO2. High-quality single crystals were grown using liquid phase electro-epitaxy (LPEE) and then chemically delithiated for direct comparison to atomistic modeling. LixNbO2 single crystals with well-defined lithium concentrations were preferred over active layer thin films in order to disentangle the effects of Li content variation from other phenomenon. Soft and hard x-ray spectroscopy techniques, as well as first-principles x-ray spectroscopic simulations, were used to monitor the niobium coordination and electronic structure evolution of LixNbO2. Using x-ray absorption spectroscopy, we were then able to confirm the depopulation of the Nb 4d  z2 orbital upon Li+ ion extraction, considered responsible for the memristive behavior. The excellent agreement between the theory and experiment allows us to explain the electronic structure evolution of LixNbO2 up to x = 0.5, i.e., degenerate p-type semiconductor regime (above which the LixNbO2 is metallic). After reexamining previous in situ oxygen K-edge x-ray absorption measurements of LixNbO2,14 we conclude that the analog memristive switching in the devices occurs within the degenerate p-doping regime rather than being associated with a full insulator-to-metal transition.

LiNbO2 single crystals were grown using a liquid phase electro-epitaxy (LPEE) method. The LPEE growth method makes use of Nb2O5 (99.9%) and LiBO2 (99.9%). The LiNbO2 crystals were grown over a 24 h period and nucleated on a niobium rod at 1.1 V with a 10:1 LiBO2 to Nb2O5 ratio. Some LiNbO2 crystals were delithiated in a 37% HCl aqueous bath at room temperature for 24 h followed by a rinse with DI water and then dried with nitrogen. Other crystals were set aside to act as pristine references, while some large crystals were reserved for device fabrication. To create the memristor devices, 100 nm of Ti and 500 nm of Au were deposited on the large crystals using evaporation and then patterned into device contacts using a lift-off process. These volatile ring dot devices were then tested with IV sweeps. Initial scans were performed to find minimum programming voltage starting at 0.1 V and increasing in steps of 0.1 V, followed by a collection of data at minimum programming voltage.19 

Some crystals were ball milled or ground manually using a mortar and pestle for powder measurements. X-ray diffraction (XRD) was performed on powderized LiNbO2 using a Bruker D8 Advance diffractometer with Bragg-Brentano geometry and a Cu Kα source at Georgia Institute of Technology. Phase identification was performed using Bruker DIFFRAC.EVA software coupled with the PDF-2016 database. X-ray absorption near edge structure (XANES) and extended x-ray absorption fine structure (EXAFS) measurements were also performed on powderized LiNbO2 to determine the effective Nb oxidation state in the bulk, as well as local electronic/atomic structure. The XANES and EXAFS were performed at beamline 20-BM of the Advanced Photon Source at Argonne National Laboratory in Lemont, IL (see Sec. I A of the supplementary material).

Soft and hard x-ray photoelectron spectroscopy (XPS and HAXPES), which have an effective probing depth of ∼4 nm (surface) and ∼15 nm (sub-surface), respectively, were performed on LiNbO2 crystals to determine the Li content, Nb oxidation state, and the valence band electronic structure. In addition, x-ray absorption spectroscopy (XAS) measurements of the oxygen K-edge were performed in total electron yield (TEY) mode, which can probe up to 5 nm deep.20 The HAXPES was performed at beamline I09 of the Diamond Light Source at the Harwell Science and Innovation Campus in Oxfordshire, UK, while variable photon energy XPS and XAS were performed at beamline 29-ID of the Advanced Photon Source at the Argonne National Laboratory in Lemont, IL. Some additional XPS was performed at the Analytical and Diagnostics Laboratory at Binghamton University, while additional XAS was performed at beamline 8 of the Advanced Light Source at the Lawrence Berkeley National Laboratory in Berkeley, CA (see Secs. I A and I B of the supplementary material).

Density functional theory (DFT) electronic ground state calculations at various Li concentrations were performed within the Vienna Ab Initio Simulation (VASP) package,21 (see Sec. I C of the supplementary material).The x-ray formalism of the core-hole approach for simulating the O K-edge in pristine and delithiated cases was performed within the ShirleyXAS + MBXASPY environment.18,22–24 The DFT information of the x-ray final state was obtained using full core-hole (FCH) approach in which an electron is removed from the inner shell of a designated excited atom within a supercell. Therefore, the interaction between the core hole and the electron was not explicitly included via many-body approaches, such as the Feynman diagram technique, but was instead accounted for using a modified oxygen pseudopotential with one electron removed from the 1s orbital for the O K-edge. The excited electron was then added to the occupied electronic structure. Next, the modified electronic system was relaxed to its ground state using DFT. Finally, the initial-state and final-state DFT orbitals and energies were provided as input for the MBXASPY software codes to produce the determinant spectra. We utilized a 1 × 1 × 1 supercell structure for the rutile and BCT NbO2 calculations, as well as the Li0.5NbO2. The fully lithiated LiNbO2 was calculated using a 3 × 3 × 1 supercell. All supercells were chosen such that their dimensions were large enough to avoid effects due to neighboring periodic images.

A representative pinched hysteresis IV curve taken on a high-quality, single crystal LiNbO2 ring dot memristor is shown in Fig. 1(a), consistent with previous reports of analog memristive behavior for LiNbO2.19,25 Since the crystals show similar IV curve responses to LiNbO2 devices grown via both molecular beam epitaxy12,14 and sputter deposition,26 we consider the underlying mechanism to be inherent to the material rather than the device processing. As such, chemically delithiated LiNbO2 single crystals should display the same electronic structure modifications as those electrochemically induced in the thin film devices.

FIG. 1.

(a) The IV Sweep (left) of a ring dot memristor device (right) fabricated using a LixNbO2 crystal grown via liquid phase electro-epitaxy. (b) A schematic band diagram depicting the niobium valence manifold reacting to Li deficiency. Oxygen 2p character forms only ≈10% of the bands at this energy and so are omitted for clarity.

FIG. 1.

(a) The IV Sweep (left) of a ring dot memristor device (right) fabricated using a LixNbO2 crystal grown via liquid phase electro-epitaxy. (b) A schematic band diagram depicting the niobium valence manifold reacting to Li deficiency. Oxygen 2p character forms only ≈10% of the bands at this energy and so are omitted for clarity.

Close modal

As shown in Fig. 1(b), the chemical potential (μ) is expected to lower into the Nb 4d  z2-derived valence band as lithium is extracted, thereby favoring hole formation (p-type). Although the Nb–O bond length and O K-edge XAS lineshape are reportedly sensitive to d  z2 occupancy,14,17,27 direct measurement of Li concentration combined with accurate simulations of the O K-edge XAS are required to fully verify the electronic structure evolution of LixNbO2 and understand the memristive mechanism.

To evaluate the quality of our pristine LiNbO2 crystals, powder x-ray diffraction (XRD) was performed. Figure 2(a) shows the hexagonal P63/mmc crystal structure (inset) and XRD pattern of our LiNbO2 powder with the reflections indexed with PDF 029-0815. The XRD pattern predominantly shows the expected structure with some negligible contribution from LiNbO3 impurities, which were likely introduced during the powder preparation process. XPS studies confirmed that these impurities were mainly limited to the crystal surface and could be reduced by exposing fresh crystal surfaces using a razor blade (see Sec. II A of the supplementary material). Henceforth, all results shown were taken on carefully cleaved samples and/or measured using bulk-sensitive techniques to avoid spectral contamination from over-oxidized surface species.

FIG. 2.

(a) X-ray diffraction and (b) EXAFS fitting of powderized LiNbO2 single crystals. The fitting range is denoted by the dashed vertical lines. (c) HAXPES (6 keV) of pristine LiNbO2 single crystals compared with NbO2 (Nb4+) and Nb2O5 (Nb5+) references, confirming the expected Nb3+ oxidation state of LiNbO2. Dashed vertical lines indicate energetic positions of the Nb 3d5/2 peak for each oxidation state.

FIG. 2.

(a) X-ray diffraction and (b) EXAFS fitting of powderized LiNbO2 single crystals. The fitting range is denoted by the dashed vertical lines. (c) HAXPES (6 keV) of pristine LiNbO2 single crystals compared with NbO2 (Nb4+) and Nb2O5 (Nb5+) references, confirming the expected Nb3+ oxidation state of LiNbO2. Dashed vertical lines indicate energetic positions of the Nb 3d5/2 peak for each oxidation state.

Close modal

Bulk-sensitive extended x-ray absorption fine structure (EXAFS) of lithium niobite ground pellets was performed to confirm the local atomic structure, as shown in Fig. 2(b). The data were fit using the LiNbO2 crystal structure from the materials project database (ICSD-451)28 (see Sec. II B of the supplementary material). The first peak around 1.654 Å corresponds to the Nb–O interaction in the first coordination shell, while the second peak around 2.596 Å is due to the Nb–Nb interaction in the second coordination shell. Excellent agreement between the experimental data and theory is achieved up to ∼5 Å.

HAXPES was then employed to confirm the bulk Nb oxidation states of our lithium niobite crystals, as shown in Fig. 2(c). NbO2 and Nb2O5 thin films are used as Nb4+ and Nb5+ oxidation state references. The Nb 3d spectra consists of two distinct spin-split peaks (3d5/2 and 3d3/2), with the energetic positions of the primary 3d5/2 peak at 204.1 eV, 206.7, and 207.3 for Nb3+, Nb4+, and Nb5+, respectively.15,29–34 Our pristine LiNbO2 displays no evidence of Nb4+; however, it does display a weak Nb5+-like component consistent with there being an over-oxidized LiNbO3 surface. Taken together, the XRD, EXAFS, and HAXPES results confirm the overall phase purity of our LiNbO2 crystals.

The effect of chemical delithiation on the electronic structure of LiNbO2 was studied via more surface sensitive XPS and XAS, as shown in Fig. 3. The lower photon energy (hν = 700 eV) used for the XPS results in a higher Li 1s photo-ionization cross section than is possible with a traditional laboratory-based XPS system, thus improving our sensitivity to variations in Li content associated with delithiation (see Sec. II C of the supplementary material). After the crystal was soaked for 24 h in an HCl bath, a drop in the Li 1s peak intensity is observed in Fig. 3(a), associated with a reduction in bulk Li content. This delithiation is accompanied by a compensating change to a higher Nb oxidation state observed in both the Nb 4s and 3d core regions, as well as a reduced Nb–O bond length in EXAFS fitting and a shifted Nb K-edge XANES spectra (see Sec. II B of the supplementary material). This delithiation also results in the depopulation of Nb states near the Fermi level, as shown in Fig. 3(b). Importantly, the lack of a clear Fermi edge upon delithiation indicates that the material stays within p-type semiconductor regime, i.e., not fully metallic.

FIG. 3.

Soft XPS of the (a) Li 1s core level and (b) valence band region, as well as (c) TEY mode O K-edge XAS compared with x-ray determinant approach simulations for pristine and 24 hour HCL soaked LiNbO2 crystals.

FIG. 3.

Soft XPS of the (a) Li 1s core level and (b) valence band region, as well as (c) TEY mode O K-edge XAS compared with x-ray determinant approach simulations for pristine and 24 hour HCL soaked LiNbO2 crystals.

Close modal

Figure 3(c) shows experimental O K-edge spectra taken using TEY mode XAS and simulations of the O K-edge calculated using the x-ray formalism of the core-hole approach.22 One feature that clearly changes with lithium content is a pre-edge feature located around 1–2 eV, which is found to increase in intensity with delithiation. While simulations show a complete lack of this pre-edge feature at 100% lithiation, our pristine experimental spectra still displays some weight at this energy, indicating that the pristine single crystals may not possess a 100% lithium concentration. This is somewhat expected, considering our LPEE growth requires dissolution and cleaning in a DI water bath, and DI water has been previously reported to remove 10%–11% of the lithium from LiNbO2.35 It is also important to note that our O K-edge lineshapes are in agreement with those previously reported by Greenlee et al. from spatially resolved in situ O K-edge XAS of LiNbO2 devices. This indicates that perfectly stoichiometric LiNbO2 is not necessary to achieve a memristive response.14 Additionally, the trends Greenlee et al. observed across their device from the positive to negative electrode match our observed changes with delithiation, confirming that changing Li content produces the observed analog memristive response.

Additional computational methods were used to qualitatively estimate the regime of p-type doping observed in our samples. Figure 4 shows the projected density of states of LixNbO2 as a function of Li content. Starting with the stoichiometric x = 100% case, the calculations predict a ∼1.5 eV semiconducting gap generated by the Nb d manifold splitting. Although the experimentally observed gap is closer to 2 eV,17 these results are consistent with other theoretical predictions27,36 and the general tendency of LDA methods to underestimate gap strengths.

FIG. 4.

Evolution of the density of states as a function of x in LixNbO2. A continuously increasing hole population was observed with decreasing Li.

FIG. 4.

Evolution of the density of states as a function of x in LixNbO2. A continuously increasing hole population was observed with decreasing Li.

Close modal

Looking at the first principles simulations for delithiated cases (x < 100%), we can now clarify the electronic response without assuming a rigid band model. As Li is driven out of the system, there is a clear autodoping effect where the top of the valence band is heavily p-doped as previously reported;17 however, the band curvature is still high at marginal delithiation and only flattens out upon deeper delithiation. The onset of a divergent dielectric function, meaning the onset of metallicity, is not predicted to occur until nearly 50% delithiation (see Sec. II D of the supplementary material). This indicates that the formation of light hole states, namely, the depopulation of d  z2 observed in operando O K-edge XAS,14 can give rise to the low field analog memristive response before truly metallic behavior sets in.

By comparing the measured valence band spectra [Fig. 3(b)] to simulations and operando XAS studies of active layers, we conclude that the memristive response is due to lithium variation within the p-type semiconductor (marginal delithiation) regime. We also conclude that the memristive response does not require the metallic phase (Fig. 4), suggesting the dominant turn-on effect is instead the formation of light p-carriers. In that regard, although the migration of lithium ions within LiNbO2 is not confined within preformed filaments, the memristive mechanism is similar to that of some filamentary memristors investigated using XAS wherein the migration of oxygen ions modulates the conductivity of the semiconducting active layer.16,37 However, in the case of LiNbO2, the more uniform Li diffusion throughout the active layer results in a much more gradual (analog) conductivity response than is observed in traditional filamentary memristors.

In conclusion, the analog memristive behavior observed in LixNbO2 single crystals has been investigated using a variety of experimental and theoretical techniques. We show intrinsic memristive behavior in lithium niobite phase-pure single crystals and use chemical delithiation of these crystals to benchmark the observed bias-induced electronic structure changes. We show that removal of Li+ oxidizes the Nb and thus depopulates the top Nb d-states within the valence band, facilitating the onset of p-type conduction prior to metallicity. This work clarifies that the field driven Li-ion motion inherent to LixNbO2 is a viable analog switching mechanism that does not require any complex interfacial effects or preforming/electroforming to function.

See supplementary material for further details on the experimental and theoretical methodologies, as well as additional supporting data including EXAFS fitting results, XPS taken at multiple photon energies, and calculations of the dielectric tensor as a function of Li content.

This material is based on the work supported by the Air Force Office of Scientific Research under Award No. FA9550–18–1–0024. Galo J. Paez acknowledges doctoral degree grant support from the Fulbright Foreign Student Program (Grant No. E0565514), and Keith Tirpak was supported by a NSF-REU (Grant No. NSF DMR-1658990). We acknowledge Diamond Light Source for time on Beamline I09 under Proposals SI20647. This research used resources of the Center for Functional Nanomaterials and the National Synchrotron Light Source II, which are U.S. Department of Energy (DOE) Office of Science facilities at Brookhaven National Laboratory, under Contract No. DE-SC0012704. This research used resources of the Advanced Light Source and the Molecular Foundry, which are U.S. DOE Office of Science facilities at Lawrence Berkeley National Laboratory under Contract No. DE-AC02-05CH11231. This research used resources of the Advanced Photon Source, a U.S. DOE Office of Science User Facility operated for the DOE Office of Science by the Argonne National Laboratory under Contract No. DE-AC02-06CH11357; additional support by the National Science Foundation under Grant No. DMR-0703406 is also acknowledged.

1.
T.
Prodromakis
and
C.
Toumazou
, “
A review on memristive devices and applications
,” in
2010 17th IEEE International Conference on Electronics, Circuits and Systems
(
IEEE
,
2010
), pp.
934
937
.
2.
C.
Sung
,
H.
Hwang
, and
I. K.
Yoo
, “
Perspective: A review on memristive hardware for neuromorphic computation
,”
J. Appl. Phys.
124
,
151903
(
2018
).
3.
E.
Chicca
,
G.
Indiveri
, and
R.
Douglas
, “
An adaptive silicon synapse
,” in
Proceedings of the 2003 International Symposium on Circuits and Systems, 2003 (ISCAS ’03)
(
IEEE
,
2003
), Vol. 1, pp.
I
I
.
4.
W. M.
Holt
, “
1.1 Moore’s law: A path going forward
,” in
2016 IEEE International Solid-State Circuits Conference (ISSCC)
(
IEEE
,
2016)
, pp.
8
13
.
5.
Y.
Li
,
Z.
Wang
,
R.
Midya
,
Q.
Xia
, and
J. J.
Yang
, “
Review of memristor devices in neuromorphic computing: Materials sciences and device challenges
,”
J. Phys. D: Appl. Phys.
51
,
503002
(
2018
).
6.
D.
Kuzum
,
S.
Yu
, and
H. P.
Wong
, “
Synaptic electronics: Materials, devices and applications
,”
Nanotechnology
24
,
382001
(
2013
).
7.
I. A.
Young
and
D. E.
Nikonov
, “
Principles and trends in quantum nano-electronics and nano-magnetics for beyond-CMOS computing
,” in
2017 47th European Solid-State Device Research Conference (ESSDERC)
(
IEEE
,
2017
), pp.
1
5
.
8.
J. J.
Yang
,
M. D.
Pickett
,
X.
Li
,
D. A.
Ohlberg
,
D. R.
Stewart
, and
R. S.
Williams
, “
Memristive switching mechanism for metal/oxide/metal nanodevices
,”
Nat. Nanotechnol.
3
,
429
(
2008
).
9.
T.
Prodromakis
,
C.
Toumazou
, and
L.
Chua
, “
Two centuries of memristors
,”
Nat. Mater.
11
,
478
481
(
2012
).
10.
J. J.
Yang
,
D. B.
Strukov
, and
D. R.
Stewart
, “
Memristive devices for computing
,”
Nat. Nanotechnol.
8
,
13
(
2013
).
11.
M. J.
Geselbracht
,
T. J.
Richardson
, and
A. M.
Stacy
, “
Superconductivity in the layered compound LixNbO2
,”
Nature
345
,
324
(
1990
).
12.
J. D.
Greenlee
,
W. L.
Calley
,
W.
Henderson
, and
W. A.
Doolittle
, “
Halide based MBE of crystalline metals and oxides
,”
Phys. Status Solidi C
9
,
155
160
(
2012
).
13.
J. D.
Greenlee
,
W. L.
Calley
,
M. W.
Moseley
, and
W. A.
Doolittle
, “
Comparison of interfacial and bulk ionic motion in analog memristors
,”
IEEE Trans. Electron Devices
60
,
427
432
(
2013
).
14.
J. D.
Greenlee
,
C. F.
Petersburg
,
W.
Laws Calley
,
C.
Jaye
,
D. A.
Fischer
,
F. M.
Alamgir
, and
W.
Alan Doolittle
, “
In-situ oxygen x-ray absorption spectroscopy investigation of the resistance modulation mechanism in LiNbO2 memristors
,”
Appl. Phys. Lett.
100
,
182106
(
2012
).
15.
J. C.
Shank
,
M. B.
Tellekamp
,
M. J.
Wahila
,
S.
Howard
,
A. S.
Weidenbach
,
B.
Zivasatienraj
,
L. F.
Piper
, and
W. A.
Doolittle
, “
Scalable memdiodes exhibiting rectification and hysteresis for neuromorphic computing
,”
Sci. Rep.
8
,
12935
(
2018
).
16.
S.
Kumar
,
Z.
Wang
,
X.
Huang
,
N.
Kumari
,
N.
Davila
,
J. P.
Strachan
,
D.
Vine
,
A. L. D.
Kilcoyne
,
Y.
Nishi
, and
R. S.
Williams
, “
Oxygen migration during resistance switching and failure of hafnium oxide memristors
,”
Appl. Phys. Lett.
110
,
103503
(
2017
).
17.
J. C.
Shank
,
M. B.
Tellekamp
, and
W. A.
Doolittle
, “
Evidence of ion intercalation mediated band structure modification and opto-ionic coupling in lithium niobite
,”
J. Appl. Phys.
117
,
035704
(
2015
).
18.
Y.
Liang
and
D.
Prendergast
, “
Quantum many-body effects in x-ray spectra efficiently computed using a basic graph algorithm
,”
Phys. Rev. B
97
,
205127
(
2018
).
19.
J. D.
Greenlee
,
J. C.
Shank
,
M. B.
Tellekamp
,
B. P.
Gunning
,
C. A.
Fabien
, and
W. A.
Doolittle
, “
Liquid phase electro-epitaxy of memristive LiNbO2 crystals
,”
Cryst. Growth Des.
14
,
2218
2222
(
2014
).
20.
F.
Lin
,
I. M.
Markus
,
D.
Nordlund
,
T.-C.
Weng
,
M. D.
Asta
,
H. L.
Xin
, and
M. M.
Doeff
, “
Surface reconstruction and chemical evolution of stoichiometric layered cathode materials for lithium-ion batteries
,”
Nat. Commun.
5
,
3529
(
2014
).
21.
G.
Kresse
and
J.
Furthmüller
, “
Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set
,”
Phys. Rev. B
54
,
11169
11186
(
1996
).
22.
Y.
Liang
,
J.
Vinson
,
S.
Pemmaraju
,
W. S.
Drisdell
,
E. L.
Shirley
, and
D.
Prendergast
, “
Accurate x-ray spectral predictions: An advanced self-consistent-field approach inspired by many-body perturbation theory
,”
Phys. Rev. Lett.
118
,
096402
(
2017
).
23.
P.
Giannozzi
,
S.
Baroni
,
N.
Bonini
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
G. L.
Chiarotti
,
M.
Cococcioni
,
I.
Dabo
,
A.
Dal Corso
,
S.
de Gironcoli
,
S.
Fabris
,
G.
Fratesi
,
R.
Gebauer
,
U.
Gerstmann
,
C.
Gougoussis
,
A.
Kokalj
,
M.
Lazzeri
,
L.
Martin-Samos
,
N.
Marzari
,
F.
Mauri
,
R.
Mazzarello
,
S.
Paolini
,
A.
Pasquarello
,
L.
Paulatto
,
C.
Sbraccia
,
S.
Scandolo
,
G.
Sclauzero
,
A. P.
Seitsonen
,
A.
Smogunov
,
P.
Umari
, and
R. M.
Wentzcovitch
, “
Quantum ESPRESSO: A modular and open-source software project for quantum simulations of materials
,”
J. Phys.: Condens. Matter
21
,
395502
(
2009
).
24.
P.
Giannozzi
,
O.
Andreussi
,
T.
Brumme
,
O.
Bunau
,
M. B.
Nardelli
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
M.
Cococcioni
,
N.
Colonna
,
I.
Carnimeo
,
A. D.
Corso
,
S.
de Gironcoli
,
P.
Delugas
,
R. A.
DiStasio
, Jr.
,
A.
Ferretti
,
A.
Floris
,
G.
Fratesi
,
G.
Fugallo
,
R.
Gebauer
,
U.
Gerstmann
,
F.
Giustino
,
T.
Gorni
,
J.
Jia
,
M.
Kawamura
,
H.-Y.
Ko
,
A.
Kokalj
,
E.
Küçükbenli
,
M.
Lazzeri
,
M.
Marsili
,
N.
Marzari
,
F.
Mauri
,
N. L.
Nguyen
,
H.-V.
Nguyen
,
A. O.
de-la Roza
,
L.
Paulatto
,
S.
Poncé
,
D.
Rocca
,
R.
Sabatini
,
B.
Santra
,
M.
Schlipf
,
A. P.
Seitsonen
,
A.
Smogunov
,
I.
Timrov
,
T.
Thonhauser
,
P.
Umari
,
N.
Vast
,
X.
Wu
, and
S.
Baroni
, “
Advanced capabilities for materials modelling with quantum ESPRESSO
,”
J. Phys.: Condens. Matter
29
,
465901
(
2017
).
25.
J. D.
Greenlee
,
J. C.
Shank
,
M. B.
Tellekamp
,
E. X.
Zhang
,
J.
Bi
,
D. M.
Fleetwood
,
M. L.
Alles
,
R. D.
Schrimpf
, and
W. A.
Doolittle
, “
Radiation effects on LiNbO2 memristors for neuromorphic computing applications
,”
IEEE Trans. Nucl. Sci.
60
,
4555
4562
(
2013
).
26.
J. C.
Shank
,
M. B.
Tellekamp
, and
W. A.
Doolittle
, “
The crystallization and properties of sputter deposited lithium niobite
,”
Thin Solid Films
609
,
6
11
(
2016
).
27.
E. R.
Ylvisaker
and
W. E.
Pickett
, “
First-principles study of the electronic and vibrational properties of LiNbO2
,”
Phys. Rev. B
74
,
075104
(
2006
).
28.
A.
Jain
,
S. P.
Ong
,
G.
Hautier
,
W.
Chen
,
W. D.
Richards
,
S.
Dacek
,
S.
Cholia
,
D.
Gunter
,
D.
Skinner
,
G.
Ceder
, and
K. A.
Persson
, “
The materials project: A materials genome approach to accelerating materials innovation
,”
APL Mater.
1
,
011002
(
2013
).
29.
I.
Goldfarb
,
D.
Ohlberg
,
J.
Strachan
,
M.
Pickett
,
J. J.
Yang
,
G.
Medeiros-Ribeiro
, and
R.
Williams
, “
Band offsets in transition-metal oxide heterostructures
,”
J. Phys. D: Appl. Phys.
46
,
295303
(
2013
).
30.
X.
Jia
,
L.
Kang
,
X.
Liu
,
Z.
Wang
,
B.
Jin
,
S.
Mi
,
J.
Chen
,
W.
Xu
,
P.
Wu
 et al., “
High performance ultra-thin niobium films for superconducting hot-electron devices
,”
IEEE Trans. Appl. Supercond.
23
,
2300704
(
2013
).
31.
Y.
Wang
,
R. B.
Comes
,
S.
Kittiwatanakul
,
S. A.
Wolf
, and
J.
Lu
, “
Epitaxial niobium dioxide thin films by reactive-biased target ion beam deposition
,”
J. Vac. Sci. Technol., A
33
,
021516
(
2015
).
32.
Q.-Y.
Hao
,
B.-C.
Zhu
,
D.-K.
Wang
,
S.-Y.
Zeng
,
Z.
Gao
,
Y.-W.
Hu
,
Y.
Wang
,
Y.-K.
Wang
, and
K.-B.
Tang
, “
A new potassium intercalation compound of 3R-Nb1.1S2 and its superconducting hydrated derivative synthesized via soft chemistry strategy
,”
ChemistrySelect
1
,
2610
2616
(
2016
).
33.
C.
Dong
,
X.
Wang
,
X.
Liu
,
X.
Yuan
,
W.
Dong
,
H.
Cui
,
Y.
Duan
, and
F.
Huang
, “
In situ grown Nb4N5 nanocrystal on nitrogen-doped graphene as a novel anode for lithium ion battery
,”
RSC Adv.
6
,
81290
81295
(
2016
).
34.
S. B.
Basuvalingam
,
B.
Macco
,
H. C.
Knoops
,
J.
Melskens
,
W. M.
Kessels
, and
A. A.
Bol
, “
Comparison of thermal and plasma-enhanced atomic layer deposition of niobium oxide thin films
,”
J. Vac. Sci. Technol., A
36
,
041503
(
2018
).
35.
E. G.
Moshopoulou
,
P.
Bordet
, and
J. J.
Capponi
, “
Superstructure and superconductivity in Li1−xNbO2 (x ≈ 0.7) single crystals
,”
Phys. Rev. B
59
,
9590
9599
(
1999
).
36.
D. L.
Novikov
,
V. A.
Gubanov
,
V. G.
Zubkov
, and
A. J.
Freeman
, “
Electronic structure and electron-phonon interactions in layered LixNbO2 and NaxNbO2
,”
Phys. Rev. B
49
,
15830
15835
(
1994
).
37.
C.
Baeumer
,
C.
Schmitz
,
A.
Marchewka
,
D. N.
Mueller
,
R.
Valenta
,
J.
Hackl
,
N.
Raab
,
S. P.
Rogers
,
M. I.
Khan
,
S.
Nemsak
,
M.
Shim
,
S.
Menzel
,
C. M.
Schneider
,
R.
Waser
, and
R.
Dittmann
, “
Quantifying redox-induced Schottky barrier variations in memristive devices via in operando spectromicroscopy with graphene electrodes
,”
Nat. Commun.
7
,
12938
(
2016
).

Supplementary Material