Rare Earth Nickelates RNiO3 Thin Films A
Rare Earth Nickelates RNiO3 Thin Films A
Rare Earth Nickelates RNiO3 Thin Films A
Abstract
Reference
CATALANO, Sara, et al. Rare-Earth Nickelates RNiO3: Thin films and heterostructures.
Reports on Progress in Physics, 2018, vol. 81, no. 046501
Available at:
http://archive-ouverte.unige.ch/unige:103464
Disclaimer: layout of this document may differ from the published version.
Rare-Earth Nickelates RNiO3: Thin films and heterostructures
1
DQMP, Université de Genève, 24 Quai Ernest-Ansermet, 1211 Geneva, Switzerland
2
Materials Research and Technology Department, Luxembourg Institute of Science and
Technology, 41 Rue du Brill, 4422 Belvaux, Luxembourg
3
Physics and Materials Science Research Unit, University of Luxembourg, 41 Rue du Brill, 4422
Belvaux, Luxembourg
* jens.kreisel@list.lu
Abstract
This review stands in the larger framework of functional materials by focussing on heterostructures
of Rare-Earth Nickelates, described by the chemical formula RNiO3 where R is a trivalent rare-earth
R = La, Pr, Nd, Sm, …, Lu. Nickelates are characterized by a rich phase diagram of structural and
physical properties and serve as a benchmark for the physics of phase transitions in correlated
oxides where electron-lattice coupling plays a key role. Much of the recent interest in nickelates
concerns heterostructures, that is single layers of thin film, multilayers or superlattices, with the
general objective of modulating their physical properties through strain control, confinement or
interface effects. We will discuss the extensive studies on nickelate heterostructures as well as
outline different approaches to tuning and controlling their physical properties and, finally, review
application concepts for future devices.
1
Outline
1. Introduction
2. ABO3 perovskites
2.1 Structure, symmetry and oxygen octahedra rotations
2.2 d-orbital physics
2.3 Metal-to-Insulator Transition: Mott insulator and charge transfer insulator
3. RNiO3 Nickelates
3.1 Phase diagram
3.2 Basic electronic configuration
3.3 Metal-to-Insulator Transition (MIT)
3.4 Magnetic Néel transition
3.5 Behavior under pressure
3.6 Behavior under magnetic fields, multiferroic effects
3.7 Theory
6. Scope of applications
6.0 Context
6.1 Exploiting the conductivity of LaNiO3
6.2 Exploiting the metal-to-insulator transition and associated properties
6.2.1 Resistance Switching
6.2.2 Opto-electronic properties
6.3 Magnetic properties
2
1. INTRODUCTION
This review stands in the larger framework of research on functional materials. A material can be
considered functional if it possesses a property that can be tuned, making it particularly appealing
for application.
Among functional materials, correlated transition metal oxides (TMOs) have attracted particular
attention due to their remarkable structural, electronic, magnetic and optical properties. Such a
range of behavior expands the potential for the development of novel electronic architectures. A
distinctive feature of complex TMOs is the competition between electronic phases close to phase
boundaries, which are often highly sensitive to subtle structural changes and, in this way, a method
of fine control is provided. In conjunction with the increasing interest in the understanding and
implementation of TMOs, recent years have seen great advances in the growth of high quality
epitaxial TMO heterostructures. This includes the fabrication of digital oxide heterostructures,
making possible the manipulation of the the electronic states of TMOs by exploiting interface-
induced phenomena and reduced dimensionality. Such epitaxial heterostructures have allowed for
the discovery of new, and elucidation of existing, physical phenomena to the point where it is now
realistic to envisage oxide electronic devices with the potential to compete with semiconductor
architectures 1-4.
In this review, we focus on the family of perovskite nickelates, which is described by the
chemical formula RNiO3, where R is a trivalent rare-earth R = La, Pr, Nd, Sm, …, Lu. These
nickelates are characterized by a rich landscape of structural and physical properties. The phase
diagram of the RNiO3 family stands as a canonical example of the physics of phase transitions in
correlated oxides to which the electron-lattice coupling contributes significantly. One of the main
characteristics of nickelates, with the exception of R = La, is a sharp metal-to-insulator transition
(MIT), which is mainly determined by the Ni-O-Ni bond angle. As a result, the physical properties
of perovskite nickelates can be modulated by a large variety of parameters ranging from
temperature, pressure and R-size, to epitaxial strain in heterostructures, and also stoichiometry or
electrostatic doping.
Much of the recent interest in nickelates is focused on thin films and their integration into
superlattices with the general objective of modulating their physical properties through strain
control, confinement and interface effects. Moreover, due to the lack of sizeable nickelate single
crystals, epitaxial thin films also constitute the best system in which to study the fascinating physics
of these compounds.
3
The nickelate family has previously been reviewed by Medarde5, Catalan6 and, more recently,
Middey et al.7. Here we focus on perovskite nickelates and their properties in heterostructure form –
both single layers and multilayers, or superlattices. We also detail various strategies that allow their
properties to be tuned and controlled and the applicability of these materials in future devices.
2. ABO3 PEROVSKITES
2.1 Structure, symmetry and oxygen octahedral rotations
Perovskite oxides are characterized by the chemical formula ABO3, where A and B are two
cations, with A larger than B, and O is an oxygen anion. The ideal perovskite structure is cubic with
space group Pm-3m, as illustrated in Fig. 1a. The B cations (represented in orange) are 6-fold
coordinated with the O anions (red), forming the BO6 octahedra that constitute the fundamental
structural and functional unit of perovskites. The octahedral units form a three-dimensional corner-
sharing network, with the A cations (purple) arranged in the cavities and 12-fold coordinated with
the oxygens. The resulting crystal lattice structure is highly stable and the A and B sites can each
support a wide variety of elements, enriching this family of materials with a plethora of properties 8.
Ideally, for the cubic unit cell shown in Fig. 1a, the lattice parameter a and the ionic radii ri
should satisfy the relation a = √2(rA+ rO) = 2(rB+ rO), where rA, rB and rO denote the radii of the A, B
and O ions, respectively. In reality, for most of the perovskite compounds this relation does not hold
and structural distortions occur in order to optimize the ionic packing, bringing the system away
from the cubic ideal. The tendency to distort can be estimated by the Goldschmidt tolerance factor t,
defined as:
(𝑟! + 𝑟! )
𝑡=
2(𝑟! + 𝑟! )
For 0.9 < t < 1, the cubic perovskite structure is stable in most cases, although, some compounds
relax to a rhombohedral symmetry. For t < 0.9, the crystal tends to adopt either a rhombohedral or
an orthorhombic structure (see Fig. 1b). For t > 1, perovskites usually develop a ferroelectric
instability and, when t is increased even further, they crystallize as a hexagonal polymorph. The
most typical structural distortions of the cubic lattice can be described by displacements of the A
cations in combination with rotations, or tilts, and distortions of the corner-sharing BO6 units. Fig.
1e illustrates the octahedral tilts leading to an orthorhombic structure from the cubic one (Figure
1d). The symmetry-allowed rigid rotations of BO6 octahedra in ABO3 compounds have been
9,10
classified by the pioneering studies of Glazer , where he showed that the most common
4
distortions of the BO6 network can be described as simple octahedral rotations α, β, γ about the x, y,
z principal crystallographic axes of the perovskite lattice, respectively, as in Fig. 1c.
Figure 1: Perovskite ABO3 structure. (a) cubic and (b) orthorhombic unit cells. In (b) the relationship
between the pseudocubic and the orthorhombic unit cells is sketched. (c) Illustration of the tilt axes identified
in Glazer notation 10. Top views of the corner-sharing BO6 units in the (d) cubic and (e) orthorhombic
structure. (f) Schematic of the oxygen octahedral network with the apical <B-Oapical-B> angle Θ indicated. In
all panels, the A cation is the orange central sphere, the B cations are the purple spheres arranged at the corner
of the cubic unit cell and 6-fold coordinated with the O anions (red).
Often, the degree of distortion of perovskite compounds is also quantified by means of the
weighted average B-O-B bond angle Θ, illustrated in Fig. 1f, which scales proportionally with t 11.
Notice that there are two distinct B-O-B angles within the Pbnm structure, corresponding to the two
distinct oxygen positions: 2 out-of-plane B-Oapical-B and 4 in-plane B-Obasal-B angles.
Regardless of the specific structure adopted, it is convenient to describe the actual material with
respect to the five atom pseudocubic (pc) unit cell that is common to all the perovskite compounds.
For example, it is useful to compare the pc-lattice parameter, apc, of different compounds, in order
to assess their structural similarity. The relationship between the pseudocubic and orthorhombic
unit cells is sketched in Fig.1b.
5
2.2 d orbital physics
As previously mentioned, a characteristic property of ABO3 compounds is that they are
structurally stable for a wide variety of cations on the A and B sites. In particular, perovskites with a
transition metal on the B site display a remarkable range of electronic phases arising from the rich
physics of the d electrons and the fact that their overlap, either direct or through the oxygen 2p
orbitals, can be tuned by structural changes.
Figure 2: 3d orbitals. (a) Illustration of the 3d wave functions, with respect to the ligand oxygen positions.
The colors correspond to the different sign, or phase, of the wave functions. (b) Octahedral crystal field
splitting ΔCF of the 3d orbitals.
3d orbitals form in five distinct wave functions, illustrated in Fig. 2a. In the ABO3 compounds,
the six oxygen anions surrounding the B cation produce an octahedral crystal field, which lifts the
degeneracy of the d orbitals. As a result, the wave functions that point toward the oxygen sites (eg)
are lifted in energy above those directed towards the space between oxygen sites (t2g), as shown in
Fig. 2b. Furthermore, as d states are spatially confined, TMOs are characterized by strong electronic
correlations, U, with respect to the material bandwidth, W. Such a large U:W ratio can eventually
bring about a Mott transition in open-shell systems, as discussed later.
Inter-site hopping and magnetic interactions are established through the overlap between the B
site d wave functions and the p wave functions of the O2- anions. The properties of ABO3
compounds are, therefore, easily manipulated by tilting and deforming of the BO6 units. In fact, as
6
both oxygen p and transition metal d orbitals are highly directional, even subtle changes in the
ABO3 structure will influence the charge, orbital and spin degrees of freedom of the system, and
vice versa. This high level of interplay between so many parameters is at the origin of the variety of
electronic phases observed in perovskite compounds12. In the following section, we will focus on a
particular example of electronic phase transition: the metal-to-insulator transition.
7
band of the transition metal ions lies close to the p band of the ligand oxygen ions, which mediate
the inter-site hopping of the electrons. Within this context, the lowest energy charge excitation is
the transfer of an electron from the p band to the d orbitals, leaving one ligand hole L in the oxygen
band and n+1 electrons at the transition metal site:
𝑑 ! → 𝑑 !!! 𝐿.
The energy required for this kind of excitation constitutes the charge transfer (CT) gap, Δ, which
can be viewed as the energy separation between the centers of mass of the oxygen p band and the
transition metal upper Hubbard band, see Fig. 3b. Materials complying with this picture are known
as charge-transfer insulators.
Figure 3: Schematic band structure of (a) a Mott-Hubbard insulator and (b) a charge-transfer insulator.
The difference between the Mott-Hubbard-type and CT-type of insulators is sketched in Fig. 3. In
the former, the MIT is controlled by the Coulomb interaction, U, which is the energy required for
the process 𝑑!! + 𝑑!! → 𝑑!!!! + 𝑑!!!! , while in the latter the MIT is controlled by the CT energy, Δ,
corresponding to the process 𝑑 ! → 𝑑 !!! 𝐿. These energy scales are to be compared to the system
bandwidth, W. As W decreases, the gap opens and the material tends to become an insulator. For a
Mott-Hubbard insulator, U < Δ and the gap energy 𝐸!"# is directly proportional to U; in contrast,
8
Finally, some TMOs are found to be insulating even though they are predicted to be metallic
according to the ZSA scheme. In fact, the CT energy decreases systematically as the valence of the
TM ion increases 17. TM compounds with unusually high valence, such as Cu3+, Ni3+ and Fe4+, are
expected to be characterized by low, or even negative, Δ and, therefore, should display metallic
behavior. Instead, some of these compounds, such as NaCuO2, LiNiO2,CaFeO3 and all the rare earth
18
nickelates except LaNiO3, are insulators at low temperature . The nature of these insulators was
first unveiled by Mizokawa et al. who, through a detailed analysis of the x-ray photoemission and
18,19
absorption spectra of NaCuO2 , concluded that the CT energy should be close to zero and
comparable to the p-d hybridization strength. It was shown that, for this compound, the ground state
is characterized by a dominant contribution of the dn+1L electronic configuration, strongly
hybridized with the dn state. While the dn+1L configuration is typically described as a continuum of
states, the strong p-d hybridization causes a localized state with dn+1L character to be split off from
the continuum. Then, the insulating gap is found to be proportional to the hybridization strength and
the lowest energy charge excitation corresponds to the process:
(𝑑 !!! 𝐿)! + (𝑑 !!! 𝐿)! → (𝑑 !!! 𝐿! )! + (𝑑 !!! )! ,
meaning that the gap has a p-p character. As a consequence, Mizokawa et al. proposed an extension
of the ZSA diagram including the negative CT insulators, identified by a small or negative Δ value,
as compared to the p-d hybridization strength and the Coulomb interaction U. As discussed later,
the perovskite nickelates belong to this region of the phase diagram and display an insulating
ground state, making them a notable example of a negative CT system.
3. RNiO3 NICKELATES
3.1 Phase diagram
The perovskite nickelate family is characterized by the chemical formula RNiO3 where R is a
trivalent rare earth R = La, Pr, Nd, Sm, …, Lu. The physical and structural properties of these
compounds are summarized in the phase diagram displayed in Fig. 4, parameterized by temperature
and R size. One of the main characteristics of RNiO3s, with the exception of R = La, is a sharp
metal-to-insulator transition (MIT) as the temperature is reduced, with the transition temperature
5,6,20
decreasing as the size of the R cation increases . Equivalently, the evolution of the MIT can be
correlated to the Goldschmidt tolerance factor t, as described in section 2.1. The decrease of the
tolerance factor with R size tends to reduce the Ni-O-Ni angle, Θ, with a consequent reduction of
the overlap between the Ni-3d and O-p orbitals. Therefore, the Ni-O-Ni bond angle is also often
taken as the control parameter in the nickelate phase diagram.
9
For a long time, the crystal structure of the RNiO3 family was thought to be orthorhombic (space
group Pbnm) in both the metallic and the insulating regimes, except for the case of LaNiO3, which
crystalizes in a rhombohedral structure (R-3c space group). It was recognized early on that the MIT
is concomitant with an expansion of the unit cell volume of about 0.2 %, and that sharp MITs are
often accompanied by a structural phase transition. This phase transition was observed by neutron
diffraction and high-resolution x-ray diffraction experiments, which revealed two distinguishable
NiO6 octahedral groups in the insulating phase, leading to a monoclinic distortion and a space group
P21/n 21-24. This structural transition can be seen as a cooperative breathing distortion of the oxygen
octahedra, which alternately contract and expand in a three-dimensional checkerboard pattern.
Consequently, the insulating phase is characterized by a Ni-O bond length disproportionation (BD)
with two inequivalent Ni sites, as illustrated in Fig. 5. These distinct Ni-O bonds are crucial for the
occurrence of the MIT according to the most recent theoretical models (see section 3.7) 25-32.
At high temperature, RNiO3 compounds are metallic and paramagnetic, whereas the insulating
ground state is antiferromagnetic. For R = Nd, Pr, the MIT occurs simultaneously with a Néel
transition from the paramagnetic to the antiferromagnetic state; in contrast, for rA ≤ rSm the MIT and
spin-ordering transitions take place at different temperatures, with TNéel < TMI. Furthermore,
measurements of resistivity versus temperature of both NdNiO3 and PrNiO3 exhibit a clear
hysteresis of the MIT, indicative of a first-order transition, while barely any hysteretic behavior is
5,6
observed for the members of the family with rA ≤ rSm . The absence of a visible hysteresis loop,
initially interpreted as a signature of a second-order transition, is more likely due to increased
thermal fluctuations reducing the size of the hysteretic effect as TMI is shifted to higher
temperatures. Indeed, measurements performed on YNiO321 and SmNiO333 detected a small but
visible hysteresis of the resistivity versus temperature curve close to the critical temperature.21 On
the other hand, the Néel transition is observed to be first-order for the compounds displaying TNéel =
33-35
TMI and second-order in those for which TNéel < TMI . A paramagnetic metallic behavior has
5,6,36
always been observed for bulk polycrystalline LaNiO3 samples, at all probed temperatures .
Nevertheless, measurements on recently synthesized LaNiO3 single-crystals have suggested a
surprising antiferromagnetic transition at TNéel = 157 K, while the material remains metallic and
rhombohedral 37,38.
10
Figure 4: Phase diagram of the RNiO3 family summarizing the evolution of the structure as well as the
metal-insulator and Néel transition temperatures as a function of the tolerance factor t (bottom x axis) and the
<Ni-O-Ni> angle Θ (top x axis). BD indicates the bond-disproportionated state. R3c, Pbnm and P21/n indicate
space groups corresponding to rhombohedral, orthorhombic and monoclinic crystal structures, respectively. A
sketch of the high and low temperature structure is overlaid on the figure. Notice that the scale of the <Ni-O-
Ni> angle is non-linear. The t values are calculated from experimental measurements of the R-O and Ni-O
distances and therefore differ from the theoretical ones, which assume a purely ionic bonding. Adapted from
ref 6,20,22,39.
11
calculations and experimental results confirm that Δ can be small or even negative for the RNiO3
compounds, with a dominant contribution of the d8L electronic configuration 26,46
. Recent
experiments combining x-ray absorption (XAS) and resonant inelastic x-ray scattering (RIXS) show
that the oxygen p-band contributes significantly to the RNiO3 electronic structure, confirming the
negative charge transfer scenario 46. This spontaneous occurrence of oxygen holes in the nickelates
is sometimes referred to as self-doping.
(a) (b)
Figure 5: Illustration of the BD characterizing the RNiO3 ground-state. (a) The short-bond NiO6 units
surrounded by long bond NiO6 units, forming a checkerboard arrangement. (b) Sketch of the electronic
configuration corresponding to the two inequivalent Ni sites, highlighting the charge and spin distribution.
Adapted from ref. 32
Concerning the metallic phase, the Fermi surface of the nickelates is formed by an electron pocket
at the Γ point and hole pockets centered at the corners of the Brillouin zone, giving two distinct
47-50
conduction channels . In fact, the Hall coefficient of the nickelates is positive, indicating the
51-
presence of p-carrier, whereas measurements of the Seebeck coefficient reveal mainly n-carriers
58
. Thus, a correct model of the metallic phase should take into account at least two conduction
bands.
12
therefore, the hopping strength tpd are reduced. Consequently, the bandwidth W of the system
decreases and TMI rises from 0 K (LaNiO3) up to 600 K (LuNiO3) 60.
16
The observation of a significant O-18O isotope effect on TMI highlighted the presence of a
marked electron-lattice coupling in the system and its substantial role in the MIT.61 In fact, as
described in section 3.1, the MIT is concomitant with a symmetry reduction to a monoclinic
structure, accompanied by a breathing distortion of the NiO6 octahedra, creating two inequivalent
Ni sites in the insulating phase. Such a breathing distortion, which leads to a bond-disproportionate
state with long and short Ni—O bonds (see Fig. 5), lowers the symmetry and is a key ingredient for
the stabilization of the RNiO3 insulating ground state.
Initially, the bond disproportionation was associated with a charge ordering on the Ni sites,
21,44,62
alternating 𝑑 !!! and 𝑑 !!! electronic configurations . Such a scenario was reinforced by a
62
variety of experimental measurements: resonant x-ray diffraction at the Ni K-edge , high
23,24
resolution neutron and non-resonant x-ray diffraction , high resolution x-ray absorption
63
spectroscopy (XAS) and, more recently, 57Fe probe Mössbauer spectroscopy 64. The presence of
the charge disproportionation was inferred either from analysis of the diffraction patterns, which
revealed bond lengths corresponding to the two inequivalent Ni sites (Brown bond-valence model
24
) or, alternatively, a valence splitting of the Ni cations was extracted from high resolution XAS at
63
the Ni K-edge . Interestingly, these studies also show that the amount of charge
24,63
disproportionation, δ, increases with the degree of orthorhombic distortion along the family .
Most recent calculations and experiments, however, support a more subtle picture in which a
uniform Ni d8 occupation is more likely to occur. According to this picture, the configuration in the
high temperature, metallic phase would be close to d8L for all Ni atoms and then, as the bond
disproportionation distortion takes place, there would be a splitting 𝑑 ! 𝐿 + 𝑑 ! 𝐿 → 𝑑 ! 𝐿! + 𝑑 ! into
25-28,41
two different Ni sites . The Ni-d8 state corresponds to the Ni atoms in the large O6 octahedral
cages and can be described as a Ni2+ cation in a high-spin state (t2g6eg↑2eg↓0 configuration with S =
1). Regarding the Ni-d8L2 ion, located within the small O6 octahedron, it also resembles a Ni2+
ionization state and a direct estimate of charge transfer between the two Ni sites, using DFT+U
25,28,65
calculations, renders a negligible result . Despite this, however, the two Ni states are in fact
very different: the Ni-d L ion has a fully occupied t2g6 shell (as it would correspond to a Ni4+ low-
8 2
spin configuration) and, additionally, it shares about two electrons (ligand holes) from the
surrounding oxygen octahedron. These additional electrons are not (or, at most, weakly) spin
polarized, so that the total magnetic moment of this Ni is nearly null (in good approximation, we
have an S = 0 spin configuration). Thus, the insulating gap splits the dispersive d8L band of the high
temperature phase into two localized bands 18,25-27.
13
It is interesting to note that the above picture for the ground state of the RNiO3 compounds has
now been verified by various model-Hamiltonian, DMFT+DFT and DFT+U approaches, all of
which render compatible results 25-28. We will discuss the theoretical efforts further in Section 3.7.
14
As previously described, looking at the bulk RNiO3 phase diagram (Fig. 4) one can distinguish two
main regimes, one for which TNéel = TMI and a second where TNéel < TMI. Originally, it was
suggested that such a difference in the critical temperature of the system could be related to the
existence of distinct electronic correlations for the members with rA < rNd, changing the intrinsic
67,72
nature of the insulating ground state . Furthermore, Lee et al. proposed a Landau model for R =
Nd, Pr where both the MIT and Néel transitions result from the nesting of a spin density wave at
qBragg = ( ¼ ¼ ¼ )pc, suggesting that the antiferromagnetic interaction drives the MIT, in agreement
50,61
with Medarde et al. . In spite of these interesting proposals, for the case of bulk nickelates with
R ≠ La the experimental results confirm that the antiferromagnetic phase can be stabilized only
below (TNéel < TMI) or concurrently with (TNéel = TMI) the MIT, suggesting that the magnetic
interaction is suppressed by charge fluctuations in the presence of delocalized electrons 5,6.
Figure 6: (a) Bragg peak corresponding to the (¼ ¼ ¼)pc wavevector of a SmNiO3 film (10 nm), measured
with resonant soft x-ray scattering at the Ni L3 edge, with π (red curve) and σ (black curve) polarized light, at
T = 4 K. The linear dichroism is consistent with the antiferromagnetic nature of the diffracted intensity. (b)
Temperature dependence of the intensity of the (¼ ¼ ¼)pc Bragg peak, measured at the Ni L3 and Sm M5
resonances of the same SmNiO3 film, showing the induced behavior of the magnetization at the rare earth
edge. (c) Sketch of a proposed non-collinear magnetic structure for the nickelates. The black arrows indicate
15
the axes of the orthorhombic unit cell. The red arrow highlights the Bragg vector (½ 0 ½)o in orthorhombic
notation, which corresponds to q0 = (¼ ¼ ¼)pc in pseudocubic notation.
Within this picture, the case where TNéel = TMI would not indicate a particular coupling between
magnetic and charge interactions, but would simply result from the quenching of the Néel transition
down to TMI. Such a picture is also supported by recent theoretical calculations73 and optical
spectroscopy experiments 74. Lu et al. have studied the charge and spin response functions of RNiO3
with multiorbital random phase approximation (RPA) in a two-band Hubbard model, showing that a
charge instability at the wave vector qc = (½ ½ ½)pc drives the system into the bond-
disproportionated phase via strong electron-phonon coupling, whereas the spin instability at qBragg =
(¼ ¼ ¼)pc plays only a minor role in the MIT 73, contrary to the proposal of Lee et al. Moreover,
Ruppen et al. have measured the temperature dependence of the optical conductivity spectra of
74
epitaxial films within a regime that reproduces the RNiO3 bulk behavior well . In this work, the
authors demonstrate that the onset of the antiferromagnetic order only reinforces the bond-
disproportionation, while the MIT remains bound to the onset of breathing distortion, which is
74
identified as a vital ingredient for the MIT to occur . Similar conclusions are reached in recent
DFT+U works 27,73.
While these conclusions are currently well-established for the case of polycrystalline bulk
samples and sufficiently thick (above 10 u.c.) films of nickelates, ultrathin films and nickelate-
based superlattices reveal an even richer behavior, as described in detail in the following sections.
Also, as mentioned in section 3.1, the growth of high quality LaNiO3 single crystals has been
recently reported, together with the observation of static magnetic ordering within the metallic
37,38
phase . Such results may mean that, in the coming years, the so-far well-established nickelate
phase diagram may need to be modified to reflect these new perspectives,.
16
increasing the applied pressure results in a decrease of the transition temperature TMI, whereas the
room temperature resistivity of the samples remains unchanged. For the case of PrNiO3, the
insulating phase is suppressed for sufficiently high pressure (P > 11.6 kbar), metallic transport
behavior being observed throughout the whole temperature range probed (4 K – 300 K). For the
case of NdNiO3, the suppression of the metallic phase was predicted to occur for P > 31 kbar, which
75
exceeded the maximum pressure attainable during the experiments . Notably, such a behavior
distinguishes the nickelates from other charge transfer oxides, e.g. the layered cuprates, for which
the insulating phase is stabilized by the application of increasing hydrostatic pressure 75,80.
The closing of the insulating gap Egap with pressure can be explained by an increase of the system
28,81
bandwidth (W) associated with the imposed structural constraints . Roughly speaking, Egap
depends on the difference between the charge transfer energy Δ and the system bandwidth W: Egap =
Δ – W. In the RNiO3 family, the corner-sharing NiO6 octahedral units tilt in order to adapt to the
difference between the R-O and Ni-O bond lengths, as quantified by the tolerance factor, t. The
application of pressure helps to reduce such a mismatch by diminishing the empty space in the unit
cell, so the system tends to have smaller orthorhombic tilts, or, equivalently, straighter Ni-O-Ni
82
bond angles, as experimentally observed by Medarde et al. in bulk PrNiO3 compounds . As a
consequence, the overlap of the oxygen band with the Ni d-shell increases and the bandwidth
broadens, favoring conductivity. On the other hand, the charge transfer energy Δ seems to be
essentially unaffected by pressure75,83, in contrast with the case of other charge transfer compounds
such as the layered cuprates. Finally, x-ray absorption spectroscopy (XAS) measurements
performed by Ramos et al. on YNiO3 indicate that, within the resolution of the experiment, the local
geometry of the NiO6 units remains unaffected by pressure 79.
Few studies have been published concerning the effect of magnetic field on the RNiO3 properties.
Yet, Canfield et al. reported that the resistivity curves and TMI of NdNiO3 are, in essence,
insensitive to the application of fields of up to 4 T76, suggesting that the RNiO3 behavior is
extremely robust against magnetic field, which, in turn, suggests magnetoresistive effects should be
small.
In this context, it is worth noting that an improper ferroelectric order in the insulating phase of
the nickelates, driven by the onset of the Ni spin order, which would break the P21/n
centrosymmetry to render the material polar, has been predicted from DFT+U simulations 84. Such a
magnetically-induced electric polarization should thus be highly responsive to changes in the spin
17
configuration, as is typical of the so-called “type II” multiferroics (such as e.g. TbMnO3), which
85,86
show similar behavior . If such a polar order were experimentally confirmed, its control by
means of magnetic fields would open yet another interesting field of possibilities for RNiO3
compounds.
The theoretical work on nickelates has proved challenging, as only recently a consensus has
been reached on the nature of the insulating phase of these compounds. It is not our purpose here to
give a detailed account of the evolution of the theory of nickelates. We will limit ourselves to
summarizing the current status of the field, and to comment briefly on the relative position of the
model-Hamiltonian and first-principles simulation approaches to these materials.
The most common approach to studying nickelates, or any strongly correlated oxide for that
matter, starts with the formulation of model Hamiltonians, which are then solved analytically or
numerically. The numerical solutions range from exact diagonalization for small clusters, to
approaches combining dynamical mean field theory (DMFT) and density functional theory (DFT)
techniques, which permit a better embedding of the part of the system that is treated more
accurately. Such numerical techniques are characterized by providing a good description of
exchange and correlation effects in the many-body electronic problem. Thus, the critical step in the
model-Hamiltonian approach is the definition of the model itself and the determination of its
parameters, a task that is far from trivial in the case of nickelates.
Today we know that a successful model for the insulating nickelates must account for the
possible formation of holes in the ligand anions; it must also effectively incorporate the effects of a
breathing distortion of the O6 octahedra as an important, or even necessary, condition for the
insulating state to emerge. The relevance of these factors was summarized by Park et al. 25, whose
models (including the t2g and eg Ni orbitals in the DMFT basis set and solving by DFT+DMFT
methods) capture the insulating state of these compounds, the differentiated magnetic and electronic
behavior of the two Ni sites (termed “site-selective Mott transition” by these authors), and the
26
enigmatic absence of charge order. More recently, Johnston et al. worked with a model-
Hamiltonian for Ni eg and O p orbitals, which they solved for small clusters of atoms by exact
diagonalization. In agreement with Park et al., these authors showed that the insulating state of the
nickelates is associated with two distinct Ni sites with (d8L2)S=0 and (d8)S=1 configurations, which
18
they described as “charge ordering without actual movement of the charge”. Finally, other relevant
27
modeling studies were performed by Subedi et al. and, more recently, Seth et al. 87, who showed
that a minimal (low energy) Hamiltonian including only the Ni eg bands and solved via
DFT+DMFT methods was sufficient to capture the insulating ground state of these materials for a
27
wide range of input parameters. The work of Subedi et al. emphasizes the importance of the
breathing (monoclinic) distortion of the structure, as this causes a modulation of the effective on-
site energy of the eg orbitals, which, in turn, favors the differentiated behavior of the two Ni
27
sublattices. Further, in their original work and in a later analysis of experimental optical spectra
88
, Ruppen et al. find evidence that a better description of the nickelates may be achieved for
relatively small values of the Coulomb repulsion U (~ 1-2 eV), which comes close to the Hund’s
coupling constant J (~ 0.8 eV).
Hence, while there are open issues on the model Hamiltonian front, for example, regarding the
description of the experimentally-observed spin order, it seems fair to say that the basic picture of
the insulating nickelates is finally well-established. It is interesting to note that, in hindsight, many
of the ingredients in the final (correct) solution were already present in the early theoretical study of
Mizokawa et al. 41, although that work seemed to imply an actual transfer of charge between the Ni
sites, the absence of which has caused considerable confusion over the years.
Among other materials, rare earth nickelates have a peculiarity that makes them very challenging
for traditional model-Hamiltonian studies, namely, the critical role that structural distortions play in
the stabilization of the insulating phase. The importance of some atomistic features, in particular the
concerted rotations of the NiO6 octahedra, is apparent from the nickelate phase diagram. However,
the importance of the octahedral breathing was not so easy to grasp a priori in the context of
strongly correlated oxides, in which structural distortions, such as Jahn-Teller, can typically be
assumed to follow an electronic driving force. Realizing that such structural features are so
important, and incorporating their effects into the electronic Hamiltonians, has been crucial to
achieve a satisfactory description of these compounds.
The challenges posed by rare earth nickelates highlight the usefulness of first-principles
approaches that, like DFT, provide a complete, structural and electronic, self-consistent description
of the material. In fact, the major theoretical works mentioned above25,27 rely on a systematic
application of DFT methods to define and compute the parameters of the corresponding model
Hamiltonians. Extending the theoretical work on nickelates toward the design and optimization of
new materials will lean heavily on the performance of DFT-based methods, such as the Hubbard-
corrected “DFT+U”, that will allow the structural and electronic ground state of hypothetical
19
systems to be solved in an unbiased way. Fortunately, our understanding of how to treat nickelates
at the DFT+U level has improved very much over the past few years, to the point where these
methods now offer a standalone alternative to the investigation of these compounds.
Indeed, systematic DFT+U investigations of rare earth nickelates have shown that these methods
capture the splitting of the Ni sites into (d8L2)S=0 and (d8)S=1 sublattices in a natural and self-
consistent way 28, producing quantitatively correct results for all measured structural and magnetic
28,29,89 89
properties and, most recently, offering new insights into the triggered nature of the MIT .
For years, one of the few lingering issues with DFT+U methods has been the prediction that the
ground state of these compounds is ferromagnetically ordered 29; however, recently it was realized
31
that this result is strongly dependent on the Ueff correction used in the simulations . More
precisely, if large Ueff values (~ 7eV) are used, a collinear ferromagnetic solution is indeed favored;
however, if relatively small Ueff values (< 2eV) are used instead, a collinear antiferromagnetic
28,89 29
ground state is stabilized . Further, the results of Prosandeev et al. suggest that a similar U-
dependence is observed when allowing for non-collinear spin structures in the simulations. Hence,
while we still lack a detailed study finally settling this issue, the existing results indicate that
DFT+U simulations with a small U correction are able to correctly predict the basic features of the
non-collinear magnetic structure of nickelates.
It is interesting to note that the comparison of DFT+U and experimental results suggests that
nickelates are better treated by using small values of the effective Hubbard Ueff parameter, which, in
the most common approach 90, can be interpreted as the difference Ueff = U - J. This is in agreement
88
with the conclusions of the aforementioned Ref. , which used model Hamiltonians to interpret
measured optical spectra. Further, noting that the choice of Ueff in DFT+U simulations determines
all the computed properties – structural, electronic, magnetic –, and that the agreement with
28,89
experiment is good for all of them , we can be optimistic about the robustness and reliability of
current DFT+U simulations of these materials.
Let us also point out that, while DFT+U simulations do not incorporate statistical averages by
91
default, it is perfectly possible to use these schemes within e.g. molecular dynamics or time-
92
dependent DFT frameworks and access the temperature-dependent behavior of the material. An
application of such methods to nickelates, including the possibility that spins disorder, would be
computationally intensive, but perfectly viable as a matter of principle. In fact, recent DFT+U
results suggest that this would be a promising strategy to elucidate the conditions for the
28
simultaneous, or separated, MIT and spin-ordering transitions. Note that Varignon et al. have
20
shown, using DFT+U methods, that the bulk RNiO3 compounds with R = Lu-Sm support an
insulating state irrespective of the spin order, a result that is compatible with the existence of a spin-
disordered insulating phase for those materials. In contrast, the same authors predict that bulk
NdNiO3 and bulk PrNiO3 would be metallic if their spins were to adopt a ferromagnetic
arrangement. It thus follows that, when in a magnetically disordered state, with spins dynamically
exploring all sorts of locally antiferromagnetic (insulating) and ferromagnetic (metallic)
configurations, such compounds will appear to be metals. This is consistent with the experimental
observations for bulk nickelates, and suggests that DFT+U simulations correctly capture the effects
behind the temperature-driven transitions in the family.
Interestingly, as we will see below, it is experimentally known that epitaxial effects in NdNiO3
93
films can induce an insulating paramagnetic state not accessible in the bulk material . This
suggests that a DFT+U calculation should predict such NdNiO3 films to be insulating irrespective of
the spin order, a conjecture that remains to be tested. Predictably, in the near future, we will see a
growing number of DFT+U simulations of nickelate heterostructures; this will hopefully confirm
that these methods retain the accuracy obtained for the bulk compounds, and are thus useful to
understand and design evermore complex materials.
Let us note that our current picture for the insulating phase of the rare earth nickelates was already
hinted at in early DFT+U works. For example, in 2002 Yamamoto and Fujiwara94 investigated
YNiO3 and reported the ground state that is commonly accepted today. Interestingly, in agreement
with most (if not all) DFT+U publications in the literature, these authors observed that the
computed charge difference between Ni sites is negligible, a result that may have seemed in conflict
with experiments and model theories in the past, but which we now know is perfectly correct. The
same ground state was obtained and further analyzed by Mazin et al.44 who emphasized that the
coexistence of two distinct Ni sublattices provides a mechanism for the MIT of nickelates and
explains the absence of on-site Jahn-Teller distortions. It is also relevant to mention here the work
that Raebiger et al.95 published in 2008 on how charge orders upon changing the nominal oxidation
state of transition metal atoms. These authors explained a widespread DFT finding – that
“calculations show only negligible changes in the local transition metal charge as the oxidation state
is altered” – in terms of a “negative charge-feedback” mechanism, which they illustrated with
various examples. In absence of a detailed investigation, it does seem that the alleged charge order
of rare earth nickelates follows the “negative charge-feedback” rules, as proposed in Ref. 95.
21
In summary, the MIT in RNiO3 nickelates has been studied in great detail by a variety of
theoretical and quantum simulation approaches, all of which have by now converged to a common
picture of the physical mechanism behind it with a key role played by the bond disproportionation.
Hence, theory-wise, we are now well-equipped to tackle the RNiO3-based heterostructures of great
interest today.
While the first successful synthesis of bulk RNiO3s was demonstrated in 1971 by Demazeau et
al. 96, their physical properties were extensively characterized only in the early 90s in the pioneering
work of García-Muñoz et al. 34,36,66 The discovery of a distinctive MIT in all members of the family,
excluding the case of R = La, and the observation of a peculiar antiferromagnetic structure
characterizing the insulating phase, sparked interest in the compounds. Nonetheless, the
experimental study of nickelates was limited to polycrystalline ceramic samples, the single crystal
size being restricted to the micron scale. Furthermore, the synthesis of the materials involved
challenging conditions, requiring both extremely high oxygen pressure and high temperature for the
chemical stabilization of the perovskite phase. While the synthesis of high quality LaNiO3 single
37,38
crystals has recently been reported , achieving high quality nickelate samples in thin film form
has so far been essential for the experimental study of these compounds.
Nowadays, thin film deposition techniques, such as molecular beam epitaxy (MBE), pulsed laser
deposition (PLD) and radio-frequency off-axis magnetron sputtering, yield highly crystalline RNiO3
thin films and enable advanced atomic-scale control of their lattice and electronic properties.
With respect to the bulk counterparts, the behavior of thin films can be highly sensitive to the
choice of substrate, dimensionality and the possibility of combining together different compounds
into a superlattice structure.1 In the following, we attempt to cover the plethora of experimental and
theoretical work that applies such a strategy to the RNiO3 family.
22
Figure 7: Strategies to tune the metal-to-insulator transition in RNiO3 thin films.
The structure of nickelates can be modified by growing epitaxial films on top of different
substrates, providing varying levels of epitaxial strain (εxx). Since the in-plane lattice parameter of
the film has to match the lattice structure of the substrate, the film will experience, with respect to
its bulk counterpart, an in-plane biaxial strain constraint εxx, defined as:
𝑎!"# − 𝑎!"#$
𝜀!! =
𝑎!"#$
where abulk is the in-plane lattice parameter of the film material in its bulk form and asub is the in-
plane lattice parameter of the substrate. Due to this in-plane constraint, the out-of-plane lattice
23
parameter of the film will expand (compressive strain) or contract (tensile strain) in order to reduce
the volume change, resulting in a tetragonal distortion of films grown along the (001)pc
crystallographic axis. Consequently, the BO6 units composing the perovskite structure have two
main structural pathways in order to accommodate such a distortion in epitaxially coherent
heterostructures: The oxygen octahedra can either adopt a Jahn-Teller distortion (i.e. elongate in-
plane (out-of-plane) in response to tensile (compressive) strain and, consequently, stabilize the
occupation of the in-plane (out-of-plane) –oriented orbitals) or, alternatively, arrange themselves
into a lower symmetry rotation pattern, often promoting a more distorted B-O-B bond angle 11.
For the case of the nickelates, epitaxial strain is a widely investigated strategy to control their
functionalities and has been proved to be very efficient in this respect. Much research effort has
been spent on strain engineering for achieving orbital polarization of eg orbitals. For this purpose,
most of the experimental work has focused on LaNiO3, as described in section 4.2. Epitaxial strain
has also been heavily researched in the other members of the family as a way to tune the electronic
transitions. In this case, the applied epitaxial strain provides a fine control of the MIT over
69,97-106
temperature . As an example, Fig. 8a displays the resistance as a function of temperature of
NdNiO3 films grown on top of (001)pc-oriented perovskite substrates providing different levels of
biaxial strain. As εxx ranges from the tensile regime to the compressive regime, the film TMI is
gradually shifted to lower temperatures and, eventually, a single metallic phase is stabilized for
sufficiently high compressive strain. Furthermore, in all cases, even for effectively no strain, the
observed MIT occurs at temperatures below the bulk transition temperature (see Fig. 8c). This
transport behavior has been reported by several groups and observed for different RNiO3
compounds, indicating that a common mechanism intervenes in response to epitaxial strain 69,105.
Similar to the case of bulk nickelates subject to hydrostatic pressure described previously, the
lowering of TMI with compressive strain points to a bandwidth broadening driven by the applied
structural constraints. The relation between epitaxial strain and bandwidth change is well captured
by DFT calculations. As presented in Fig. 8b, a monotonic broadening of the system bandwidth W
with strain is driven by a corresponding straightening of the Ni-O-Ni angle out-of-plane,
accompanied by the stretching (squeezing) of the out-of-plane (in-plane) Ni-O bond length.
Thus, the RNiO3 lattice responds to epitaxial strain by modifying the degree of distortion
characterizing the system. These observations are summarized in Fig. 8c, which shows the
evolution of the electronic transitions of SmNiO3 and NdNiO3 as a function of strain, compared to
the phase diagram of bulk RNiO3 parameterized by the tolerance factor. Clearly, the effect of
compressive strain on the MI and Néel transitions is equivalent to a change of the R cation size, or,
equivalently, to a shift of the tolerance factor of the system. In contrast, increasing epitaxial tensile
24
strain applied in the (001)pc crystallographic plane does not result in an increase of TMI, as the
temperature of the MIT saturates lower than in the corresponding bulk nickelate. This observation
leads to two basic conclusions: First, thin films of nickelates grown epitaxially along the (001)pc
direction always stabilize a less-distorted Ni-O-Ni angle, with respect to their bulk counterparts,
indicating that the presence of the (001)pc-oriented perovskite substrate matrix alone promotes a
systematic change of the effective chemical pressure on the RNiO3 lattice; Second, as the strain
becomes more tensile, the lattice does not accommodate strain with a stronger orthorhombic
distortion. In fact, the transport measurements of RNiO3 films under tensile strain show that the
resistivity of the system increases for high levels of tensile strain (εxx ≥ 3%), suggesting that the
elastic energy cost for deforming the film lattice exceeds the formation energy of structural defects,
such as vacancies or dislocations, as discussed by Conchon and coworkers 107.
An alternative perspective has recently been proposed by Zhang et al., who suggested that the
occurring of the MIT may be intimately linked to symmetry. They showed that the MIT is
suppressed in NdNiO3/YAlO3 thin films, in which the epitaxial strain imposes the P21/m structure
that does not allow the structural transition to the monoclinic P21/n space group associated with the
bond-disproportionation 108.
Finally, Mikheev et al. have proposed an original phase diagram for NdNiO3 thin films, adopting
epitaxial strain and thickness as control parameters that drive the system from the paramagnetic
109
metallic to the antiferromagnetic insulating phase . In particular, the authors have shown that the
metallic regime can be described either as a non-Fermi liquid (NFL) or a Landau Fermi liquid
(LFL) depending on the thickness of the films and the applied strain, which tune, respectively, the
degree of disorder in the system and the amount of orbital polarization of the eg levels.
25
Figure 8: Tuning of the electronic properties of the perovskite nickelates by epitaxial strain. (a)
Temperature- and lattice mismatch-dependence of the resistivity of (001)pc-oriented NdNiO3 films.
Adapted from Ref. 101. (b) DFT calculations showing the strain-dependence, in the in-plane and out-
of-plane directions, of the bandwidth W (top panel), the Ni-O bond length (middle panel) and the
<Ni-O-Ni> bond angle (bottom panel). (c) Phase diagram of nickelates summarizing the evolution
of TMI and TNéel and how the two transitions are correlated to strain for NdNiO3 and SmNiO3 films
compared to the bulk compounds. (b) and (c) after reference 69.
While compressive strain, applied to the most-studied (001)pc-oriented RNiO3 films, has been
proven to act as an efficient control knob for straightening the Ni-O-Ni angle and thus broadening
the system bandwidth, it has been shown that such a control ceases to be effective as the strain
verges on the tensile regime, favoring defect formation rather than a continuous evolution of the Ni-
O-Ni angle to more distorted values. However, the latter effect can be observed by selecting
different crystallographic directions for the growth of RNiO3 films. In particular, the growth of
26
(111)pc oriented perovskite oxides has received much attention over the past few years, as (111)pc-
oriented perovskite planes form a buckled honeycomb lattice, which has been predicted to favor the
stabilization of exotic electronic phases in TMO-based heterostructures 110,111. Yet, for several years
the study of (111)pc-oriented perovskite oxides has been hindered by the difficulties in synthesizing
high quality thin films along this crystallographic direction. The strong polarity of (111)pc
perovskite planes favors the formation of surface reconstruction or structural and chemical defects
112,113
that neutralize the polarity in the materials during the growth . Nevertheless, intense research
efforts have been dedicated to growing (111)pc-oriented nickelate films, resulting in high quality
93,114-118
defect-free thin samples which display novel electronic behavior, as discussed in the
following.
Interestingly, Lian et al.119 and Catalano et al.93 observe a surprising change of behavior in
NdNiO3 films grown under tensile strain on NdGaO3 substrates (εxx= +1.5%) along crystallographic
directions different from the usual (001)pc-orientation. Remarkably, the change of orientation
dramatically affects the MIT temperature. For example, NdNiO3/NdGaO3 films oriented along the
(100)o axis ((110)pc in pseudocubic notation) exhibit TMI〜200 K119, well above the TMI〜150 K
displayed by the (001)pc films. Even more strikingly, (101)o-oriented ((111)pc in pseudocubic
notation) NdNiO3/NdGaO3 films display TMI〜335 K and TNéel〜200 K, resembling the TMI > TNéel
behavior of more distorted members of the family (see Fig. 9d) 93. These observations demonstrate
that such, less conventional, orientations can stabilize a more distorted Ni-O-Ni angle than in the
case of films grown in the (001)pc direction.
Such a profound effect results from a change in epitaxial constraints applied to the system by a
given substrate symmetry. In fact, the NdGaO3 substrate is distorted orthorhombically with a B-O-B
angle of Θ〜153°. NdNiO3/NdGaO3 films oriented in the (001)o and (100)o directions have a greater
in-plane anisotropy of lattice parameter mismatch at the interface with respect to the (110)o
(corresponding to (001)pc) case. This difference in lattice mismatch along the two main in-plane
axes imposes a stronger orthorhombic distortion in the film. Lian and coworkers119 attribute the
observed behavior to this effect.
Meanwhile, Catalano and coworkers93 discuss an alternative epitaxial mechanism occurring at
(111)pc-oriented film/substrate interfaces. Along the (001)pc direction, each oxygen octahedron is
connected to the next through a single apical oxygen, forming a rectangular lattice in the plane, as
shown in Fig. 9a. In this configuration, the NiO6 octahedral units can rather easily adopt the most
energetically favorable tilt pattern, balancing the applied strain and the internal chemical pressure
120
. In the case of interfaces normal to the (111)pc-direction, on the other hand, each of the corner-
27
sharing octahedra of the substrate are connected to those of the film through three oxygen atoms
instead of only one, imposing a substantially stronger epitaxial requirement. As a consequence, the
structure favors a rotation of the NiO6 units of the film in order to adopt the tilt angle of the
substrate, maintaining the connectivity across the interface, as can be visualized in Fig. 9c. In this
way the substrate tilt pattern is propagated into the film. Bearing in mind the RNiO3 phase diagram
5,97
, one can observe on Fig. 9e that the MIT of (111)pc NdNiO3/NdGaO3 at 335 K would correspond
to Θ = 153.8°, an angle which is remarkably close to the tilts characterizing NdGaO3 (Θ = 153.2°).
Interestingly, the impact of the epitaxial constraints at the (111)pc interface is stunningly robust,
affecting up to 17 nm thick (111)pc NdNiO3/ NdGaO3 films.
It is clear, then, that growth along crystallographic directions, different from the conventional
(001)pc, can be exploited to further implement the control over the MIT in RNiO3 films by imposing
particular epitaxial constraints to the film lattice. Moreover, these results suggest that selecting such
orientations for the deposition of perovskite films could help in engineering novel materials
requiring a precise distortion structure. As an example, room-temperature polar metals have been
experimentally demonstrated in NdNiO3 and LaNiO3 thin films grown on (111)pc-oriented LaAlO3
116
substrates . The strong constraint imposed by the substrate along the (111)pc-growth direction is
crucial in achieving the non-centrosymmetric structure in the metallic state. There also exist
intriguing theoretical calculations suggesting that exotic topological phases can be stabilized in
121-123
(111)pc-oriented RNiO3-based heterostructures . Concurrently, Middey et al. have recently
demonstrated that a novel electronic phase is stabilized in (111)pc-oriented NdNiO3 films, as
discussed later in this review 117.
28
Figure 9: (001)pc and (111)pc interfaces ((110)o and (101)o in orthorhombic notation, respectively).
(a) Along the (001)pc plane, the connectivity between the BO6 octahedra is maintained through one
apical oxygen only. (b) Top view of the (111)pc plane. (c) Along the (111)pc termination plane, each
BO6 unit shares three oxygen sites with the next one. (d) Temperature-dependence of the resistivity
of NdNiO3 films grown on (001) and (111) NdGaO3 substrates. (e) Extrapolation of the Ni-O-Ni
angle for NdNiO3/NdGaO3 based on the experimental TMI and TNéel values of the films and the
RNiO3 phase diagram. From Catalano et al 93.
Different works have explored the effect of electron (n) and hole (p) doping on the MIT of
nickelates by substitution of the R cation. Originally, the first experiments were performed on
polycrystalline NdNiO3 samples doped by R cation substitution with both electron donor (Th4+,
Ce4+) and hole donor (Sr2+, Ca2+) cations 67,124
. Interestingly, it was observed that by varying the
dopant concentration between 1% and 10% the MIT is progressively suppressed independent of
whether the dopant is n- or p-type. In particular, García Muñoz et al. showed that there are two
contributions to the surpession of TMI: (i) the lattice distortion induced by steric effects and (ii) the
124
change in the electron concentration . They observed that while the steric contribution mainly
affects the R-O distance, the Ni-O bond length is concomitantly modified by the injection of carriers
on the Ni site. In fact, the purely electronic contribution plays a major role in controlling TMI, so
29
that doping with Ce and Th, the size-effect of which is expected to increase the lattice distortion and
consequently stabilize the insulating phase, results in a suppression of the MIT due to electron
doping. Nikulin et al. also studied the consequence of electron doping induced by oxygen non-
stoichiometry in polycrystalline SmNiO3 and NdNiO3 125. Again, the authors observed a suppression
of TMI proportional to the content of oxygen vacancies in the samples, which was varied between
5% and 10%. All together, such studies suggest that light carrier doping induces extra electronic
states within the RNiO3 band gap, causing a suppression of TMI.
More recently, the effects of doping have been studied in RNiO3 thin films, providing new
important insights on the electronic behavior of these materials. In particular, Ramanathan and
coworkers have explored a new approach to electron doping, based on electro-chemical reaction
126-128
rather than cation substitution . In their work, these authors reduced the valence of the Ni
cation in SmNiO3 thin films from nominal Ni3+ to Ni2+ by either oxygen vacancy formation induced
by ionic liquid gating or hydrogenation. In both cases, the change of the Ni oxidation state results in
a dramatic increase of the room temperature resistivity of the system, which hints at a purely
electronically-driven metal-to-insulator transition: meaning that as an extra electron is doped onto
the Ni site, the strong Coulomb interaction triggers the electronic localization. While such results
may seem to contradict the observations of García-Muñoz et al. on bulk compounds, it is important
to notice that the doping mechanism investigated by Ramanathan et al. is based on a chemical
reaction that transforms the Ni oxidation state without modifying the original lattice properties, as
in the case of R-cation substitution. Interestingly, in a recent work, Heo et al. also focused on the
effect of oxygen deficiency in NdNiO3 thin films, finding that the insulating phase is stabilized as
129
the oxygen vacancy concentration increases , in agreement with the results of Ramanathan et al.
Concerning the discrepancy with the observations of Nikulin et al. on the bulk parent compounds,
the authors point out that in previous works the oxygen non-stoichiometry was controlled by
varying the oxygen pressure during the growth, a method which affects the oxygen content and the
cation stoichiometry concurrently. Conversely, Heo et al. controlled the oxygen content of the
material by varying the growth temperature of NdNiO3 thin films under tensile strain, succeeding in
changing the oxygen stoichiometry independently from the R-cation stoichiometry.
30
In addition to strain engineering and optical excitation, the transport properties of RNiO3 films
have also been investigated by field effect (FE) experiments, allowing the charge carrier density of
the RNiO3 conducting channel to be modulated as a function of gate voltage applied through a
130 53,131
dielectric or ionic liquid layer. FE-based devices have been realized with LaNiO3 , NdNiO3
and SmNiO3 films 126,132. In LaNiO3 /SrTiO3 heterostructures, a 10% change of the carrier density of
a 6 u.c. thick LaNiO3 conducting channel was obtained by applying a gate voltage varying from 300
to -300 V through the SrTiO3 dielectric layer, indicating that p-type conduction dominates the
130
transport characteristics of the device . Analogous measurements performed on NdNiO3/LaAlO3
films demonstrated FE as a practical tool for tuning the temperature of the MIT over a 50 K range53.
These measurements took advantage of the electric double layer transistor (EDLT) FE technique, in
which an ionic liquid allows an in-plane 2D carrier density modulation of the order of 1014 - 1015
cm-2 to be achieved. We will see in a later section that such a tuning of electronic properties opens
the way for nickelate-based transistors. Motivated by these promising results, Son et al. propose the
design of RNiO3-based modulation-doped Mott FE transistor (MMFET), in which an MIT can be
131
switch on and off through the applied gate voltage . Marshall et al. have realized a ferroelectric
field effect device which exploits the resistivity modulation of an LaNiO3 conducting channel
133
induced through a ferroelectric Pb(Zr,Ti)O3 gate . Similar ferroelectric field effect devices have
134,135
also been designed adopting Sm0.5Nd0.5NiO3 films as conducting channel . Finally, Shi et al.
have developed an original SmNiO3-based FE transistor, which mimics the behavior of neural
synapses 127, whose concept and operation are described extensively in section 6.2.1.
31
for SmNiO3/LaAlO3 films 137, whose response could be probed over a remarkably broad THz range
(100-800 cm-1). In addition, by pumping the system in both TNéel < T < TMI and T < TNéel< TMI
temperature ranges, the authors also demonstrate that the charge order melts regardless of the
presence of the antiferromagnetic ordering.
In order to clarify the underlying physics of such dynamically excited metastable metallic
phases, Caviglia et al. and Först et al. carried out a series of measurements combining optical
138-140
excitation with time-resolved non-resonant and resonant x-ray diffraction . By identifying and
following the diffraction peaks characterizing the charge-ordered ((½ ½ ½)pc) and
antiferromagnetically-ordered ((¼ ¼ ¼)pc) superstructures of the RNiO3 insulating phase, the
authors could directly investigate the time-dependent evolution of the melting and disentangle the
dynamics of the charge, spin and lattice degrees of freedom across the insulator to metal transition
(IMT). Most notably, they found that the melting of the insulating phase begins at the buried
film/substrate interface and propagates through the film faster than the bulk speed of sound of the
material, indicating an electronic mechanism driving the IMT. In particular, further analysis reveals
that ultrafast charge redistribution is at first induced after dynamic light excitation and is followed
by the melting of the spin ordering, while both the charge and spin redistributions evolve faster than
the phonon speed in the films.
Within the family of rare earth nickelates, LaNiO3 is the odd-one-out as it remains paramagnetic
and metallic at all temperatures in bulkA. When grown in the form of ultrathin layers, however,
LaNiO3, in common with the other RNiO3s, enters an electronically localized state with no observed
71,141
bond disproportionation . As dimensionality is reduced, the films first go from a metallic
regime to one where a resistivity upturn is seen, which has been attributed to weak localization. The
critical thickness whereupon the films become insulating can vary from 2 – 6 u.c., depending on the
substrate selected for epitaxial growth 52,130,141. This thickness-dependent transport behavior is likely
driven by a profound alteration to the electronic structure of LaNiO3, marked by drastic changes to
the Fermi surface upon reducing the thickness towards the atomic limit, as photoemission studies
have shown (Fig. 10a) 48 142.
A
The recent results reported in Ref. 38 Z. Li, H. Guo, Z. Hu, L. Zhao, C.-Y. Kuo, W. Schmidt, A. Piovano, T.
Pi, D. Khomskii, and L. Tjeng, arXiv preprint arXiv:1705.02589 (2017).) indicate an antiferromagnetically
ordered state for very low resistivity single crystals.
32
This behavior may have a largely structural origin. Hardening of the phonon modes associated
with oxygen octahedral distortions, concomitant with the dimensionality-induced electronic
143
localization, has recently been reported . In the same study it was found that the film structure
differentiates into three different local structures – one found on the surface, one towards the
interface with the substrate and the third in the interior layers where the boundary effects are
minimal. See Fig. 10c. Indeed, it has been shown that the polar LaNiO3 surface is prone to
developing buckling distortions, as illustrated in Fig. 10b, which very likely leads to an increase in
the resistivity through a reduction of the orbital overlap 144. This effect can be mostly alleviated with
the addition of an encapsulating layer of, for instance, LaAlO3 but when it comes to ultrathin films,
1-2 u.c., not even this is enough to preserve the metallic state.
Efforts to understand the nature of the ultrathin insulating state of LaNiO3 are ongoing. Oxygen
vacancies, interfacial cation intermixing, dimensionality, electron-phonon coupling or defects could
all have a part to play.
Another possibility is that LaNiO3 undergoes multiple structural phase transitions under strain,
with different structures under compression and tension 145,146. Beyond the experimentally observed
phase transitions, it is important to note that calculations suggest additional structural instabilities,
hinting at a rich energy landscape, which are expected to impact electronic properties.
33
Figure 10: (a) A photoemission study showing the disintegration of the Fermi surface as the
thickness of LaNiO3 is decreased to only 1 u.c. 48 (b) Surface buckling of Ni-O-Ni bonds 144 and (c)
local structure differentiation as revealed by transmission electron microscopy143.
There remains a strong focus on LaNiO3 as a potential building block in oxide-based functional
materials, driven, in large part, by the prediction of possible high-temperature superconductivity in
LaNiO3-based heterostructures. A key ingredient in achieving this is a large Ni orbital polarization,
147
further discussed later . There are clearly limitations to the orbital polarizability of thin films of
LaNiO3. One issue, which is general to perovskites, is the aforementioned tendency for the NiO6
octahedra to rotate, rather than to adopt a Jahn-Teller distortion, under the influence of biaxial strain
148
. The octahedra are able to remain rigid under a rotational distortion and thus the immediate
oxygen coordination environment would retain an almost cubic symmetry. Also, there is predicted
34
149
to be a strong on-site Hund’s rule coupling , which would directly oppose the high degree of
preferential occupation that is sought after.
One possible avenue to explore with thin layers of LaNiO3 is to go one step further and include
them in a multilayer heterostructure. The study of multilayers containing LaNiO3 will be discussed
in the following section.
In addition to the interest in nickelate thin films as the closest system to single crystals and the
capabilities they offer to tune their electronic transitions, it is fair to say that the present wave of
interest in nickelates has been motivated by the previously mentioned prediction of Chaloupka and
Khaliullin. They propose that a novel high temperature superconductor may be designed by orbital
engineering in 1-unit-cell-thick LaNiO3 layers sandwiched between layers of an isostructural
147
insulating oxide LaXO3 (X = Al, Ga, Sc, etc.) . One of the core requisites for the successful
fabrication of a synthetic superconductor is a high degree of Ni orbital polarization. The most
intuitive way to achieve this is to break the symmetry of the Ni environment, which happens
naturally when going from a bulk material to a thin film. Ideally, a stretching (or squashing) of the
NiO6 octahedra would be induced by the application of biaxial compressive (tensile) strain and
result in a preferential occupation of the 3z2 – r2 (x2 - y2) orbitals. The ab initio calculations
predicting the formation of a cuprate-like Fermi surface in these heterostructures were further
supported by DFT+DMFT studies including an accurate treatment of electronic correlations 150,151.
35
measurements has allowed the orbital occupation profile, and thus the orbital polarization of the
interface and inner layers, to be quantified 153. These studies have revealed that the strain induced by
lattice mismatch provides the largest effect compared to the chemical composition of the blocking
layers. In LaNiO3/LaAlO3 superlattices, an asymmetry in strain-dependent orbital occupation has
been reported, with tensile strain acting on orbital polarization and compressive strain on octahedral
154
distortions . A more recent alternative route proposes enhancing orbital polarization in LaNiO3-
based heterostructures through engineering of atomic-scale structural distortions in tri-component
155
superlattices . Doping the nickelate layer via interfacial charge transfer and breaking inversion
symmetry through the introduction of a third chemical compound in the heterostructures has led to
orbital polarizations of up to 50% in atomically layered superlattices of the type (1 u.c. LaTiO3/1
156
u.c. LaNiO3/1 u.c. LaAlO3), as illustrated in Fig. 11b . The attained orbital polarization values
and the difficulty in achieving a single band crossing the Fermi level in LaNiO3-based superlattices
are explained by DMFT calculations including the oxygen bands and the important role of
correlations through the Hund’s rule coupling 149,157.
Figure 11: Engineering orbital polarization in LaNiO3-based superlattices. (a) Schematics of the
evolution of the Ni 3d eg orbital occupation in LaNiO3-LaAlO3 superlattices, and orbital
polarization values inferred from x-ray absorption spectroscopy (XAS) and reflectivity
measurements for several LaNiO3/LaXO3 (X = Al, Ga, Sc) superlattices grown on a variety of
substrates. Adapted from references152,153. (b) Theoretically and experimentally determined atomic
36
structures for (1u.c. LaTiO3/1u.c. LaNiO3/3u.c. LaTiO3) showing the polar displacements, and XAS
spectra acquired with linearly-polarized light for two- and tri- component nickelate superlattices
where r is the hole ratio quantity related to orbital polarization. Adapted from references 155,156.
LaNiO3-based superlattices have also been widely investigated in combination with manganites,
where interfacial charge transfer between Ni and Mn is at the origin of many interesting phenomena
115,163-166
. In LaNiO3/LaMnO3, charge transfer occurs from Mn to Ni due to the difference in
electronegativity with the resulting (nominal valence) Ni2+-Mn4+ cations interacting strongly
165,167-169
ferromagnetically . Exchange bias has been observed at low temperature in
37
115,168
LaNiO3/LaMnO3 superlattices , with various models having been proposed to explain this
118,170,171
unexpected behavior . Interestingly, when grown along the less conventional (111)pc
direction and for digital superlattices of 7-monolayer-thick-LaNiO3, exchange bias has been
observed at low temperature followed by an antiferromagnetically-coupled state between the
118
LaMnO3 layers . This complex magnetic evolution is the result of a subtle interplay between
interface-driven phenomena and reduced dimensionality leading to the stabilization of a (¼ ¼ ¼)pc
magnetic spiral structure in LaNiO3 (see Fig. 12c).
For LaNiO3/La2/3Sr1/3MnO3 superlattices, an oscillatory interlayer coupling between the
164 172
manganite layers depending on the thickness of the LaNiO3 spacers occurs . As depicted in
Fig. 12c, such a non-collinear coupling has recently been explained by the formation of a helical
spin structure within the LaNiO3 layers, which persists almost up to room temperature 173.
38
Figure 12: Magnetism in LaNiO3-based superlattices. (a) Polarization-dependent azimuthal scans
used to infer the polarization plane (illustrated by the arrows) of the nickelate antiferromagnetic
spiral structure under tensile and compressive strain for (2u.c.LaNiO3/2u.c.LaAlO3) and
(2u.c.LaNiO3/2u.c.DyScO3) superlattices and nickelate (PrNiO3, NdNiO3) thin films. The same
color code is used in the measurements and sketches of the spin directions. Adapted from 71. (b)
LaNiO3-thickness-dependent non-collinear oscillatory coupling in (LaNiO3)N/(La2/3Sr1/3MnO3)9
superlattices and a schematic of the magnetic structure for the superlattice with 3 u.c. LaNiO3
layers, where the manganite and nickelate layers are shown in blue and pink, respectively. Adapted
from173. (c) The observation of half-order peaks in polarized reflectivity measurements at the Mn
L3-edge of (111)pc-oriented 7-monolayer-thick LaNiO3/LaMnO3 superlattices reveals an
antiferromagnetically-coupled state between the manganite layers triggered by the digital nature of
the nickelate layer. Adapted from118.
Rich behavior is also observed in superlattices combining members of the nickelate and titanate
families. In particular, charge transfer is expected to occur from the titanate to the nickelate layers
174
due to the difference in the electron affinities of the two materials . Interestingly, the resulting
electronic reconstruction has been proposed to enhance electronic correlations in superlattices of
175
LaNiO3 and the Mott insulator LaTiO3 . Experimentally, a correlated gap of 1.5 eV has been
estimated for the insulating (2 u.c. LaNiO3/2 u.c. LaTiO3) superlattices, which also exhibit
enhanced orbital polarization (ndx2-y2 > n3z2-r2) and a splitting of the Ni eg band attributed to
interfacial structural distortions 176.
Besides LaNiO3, other members of the RNiO3 family have been integrated into various
heterostructure multilayer designs, exploring an ensemble of interfacial effects induced at the
interface between the nickelates and other perovskite compounds.
Grisolia et al. have studied the charge transfer at the titanate/nickelate interface in a series of
bilayers consisting of 7 u.c. of RNiO3 (R = La, Nd, Sm) deposited on 7 u.c. of GdTiO3 on top of a
174
LaAlO3 single crystal substrate . Spectroscopy measurements confirm charge transfer from the
titanate to the nickelate layers, the number of transferred electrons being controlled by the choice of
the rare earth cation in the RNiO3 layer.
39
Inspired by the LaNiO3-based superlattice design proposed by Chaloupka and Khallioulin, a
number of works have focused on heterostructures wherein RNiO3 layers alternate with some
insulating perovskite acting as a charge spacer (typically, a member of the RAlO3 family). Hepting
et al. have shown that the charge ordered phase of bulk PrNiO3 can be suppressed in
(mPrNiO3/nPrAlO3) superlattices, as they observed a magnetically ordered ground state in the
absence of the bond length disproportionation 105. Promising results have also been attained with the
study of (mNdNiO3/nLaAlO3)N/LaAlO3 superlattices grown along the (111)pc crystallographic
117
direction . The NdNiO3 film embedded in such a system is found to host an orbitally-ordered
ground state and to display sizeable antiferromagnetic correlations absent in the bulk counterpart or
in the (001)pc-oriented analogous superlattices. In particular, the stabilization of this unique
electronic phase is ascribed to the confinement of the Ni 3d electrons onto the hexagonal lattice
enforced by the (111)pc-oriented perovskite surfaces.
Finally, Girardot et al. investigated the properties of the first all-nickelate multilayers, alternating
SmNiO3 and NdNiO3 thin films 177. In their work, the authors compare the complex evolution of the
Raman spectra of epitaxial films of the mixed compound Sm0.6Nd0.4NiO3 and of
(nSmNiO3/nNdNiO3)N superlattices, with the periodic thickness, n, varying from 100 u.c. down to
25 u.c. The resistivity vs. temperature curve of an ((SmNiO3)100/(NdNiO3)100)1 bilayer shows two
transitions: (i) a first hysteretic transition upon heating at TMI ≈ 160 K, which is very close to the
value of TMI = 158 K for bulk NdNiO3. (ii) a second transition, less pronounced at TMI(2) ≈ 370 K,
which shows no hysteresis, similar to SmNiO3; this temperature is lower than TMI = 393 K of a
single bulk-like SmNiO3 film, an effect likely due to strain.
6. SCOPE OF APPLICATIONS
6.0 Context
The early years of study on nickelates were naturally focused on the understanding of their
fundamental behavior, namely the MI phase transition, the magnetic structure and the connection
between the two, as discussed earlier. In addition to this ongoing quest, many of the more recent
studies are marked by the exploration of potential application of nickelates in a variety of devices.
This interest in applications has been triggered by several developments.
Correlated TMOs, in general, have attracted increasing attention due to their remarkable variety
of functionalities, linked to their fascinating electronic, magnetic and optical properties. This raises
the potential for new types of electronic devices. A distinctive element of complex TMOs is the
40
178
competition of electronic phases close to a phase boundary , which is often highly sensitive to
8
subtle structural changes . One of the properties observed in some correlated TMOs, absent in
conventional semiconductors, is the sharp metal-insulator phase transition, as discussed previously,
which can be tuned by external stimuli. Already this functionality can be exploited in a variety of
126,131,179-182
applications (see e.g. Refs. ), thus offering great potential to make use of the MIT in
nickelates.
In addition to MBE, PLD and sputtering, as discussed previously, more large-scale growth of
perovskite compounds is also feasible. This possibility, often coupled with nanofabrication, means
that realization of actual devices can go alongside the increasing range of device concepts inspired
by the more fundamental study of nickelates reviewed so far.
Table 1 summarizes the main device concepts reported so far, without claim for full
completeness. While most of the propositions today concern single layers, this application
landscape is extended by the additional opportunities arising from perovskite multilayers
integrating, or entirely consisting of, nickelates.
Table 1
Overview of applications proposed in literature for rare earth nickelates, RNiO3
183
Electrode material Anode LaNiO3- δ
184-188
Supercapacitor LaNiO3/NiO
189-198
In heterostructures LaNiO3
199-202
Buffer layer in HTS LaNiO3
127,203,204
Opto-electronics Tunable Photonics SmNiO3
205
Photodetection GdNiO3
206-208
Thermochromism SmNiO3, NdNiO3
209,210
Photovoltaics NdNiO3
211,212
Ion-conduction/ Fuel Cell, SOFC SmNiO3
catalysis
213-216
Oxygen reduction LaNiO3
126,217-219
Resistance Switching Neuromorphic device SmNiO3, NdNiO3
220
Neuromimetic circuits SmNiO3
126
Synaptic transistor SmNiO3
221
Transistor LaNiO3/Si
131,222
Mott Transistor NdNiO3
41
127
Proton-gated transistor SmNiO3
186,223-233
Sensing Gas sensing LaNiO3
234
Magnetism Magnetotransport SmNiO3
235-237
Multiferrocity SmNiO3, NdNiO3
SmNiO3'
Resistance'switching'
SmNiO3' LaNiO3'
Synap:c'transistor' Gas'sensing'
SmNiO3' SmNiO3'
Opto6electronics' Fuel'Cell'
RNiO3%
NdNiO3'
Photovoltaics'
42
Figure 13: Schematic view of potential applications of RNiO3 in the field of electronics, photonics,
sensing or fuel cells. The images inside the hexagons are adapted from literature and address
specific application examples such as photovoltaics209, opto-electronics203, sensing186, fuel cells128,
resistance switching127 and synaptic transistors238. For a more complete list of applications, please
refer to Table 1.
In number, the most discussed nickelate in view of applications is LaNiO3. LaNiO3 is the only
perovskite nickelate that is metallic at all temperatures in bulk. Its low electrical resistivity of about
100 µΩ·cm, makes it a good candidate for the electrode material in oxide-based microelectronic
devices. In most of the reported application schemes, LaNiO3 is implemented in oxide
heterostructures as a metallic layer and electrode, rather than as a functional element. The literature
on such structures is vast and we cite only some representative references 189-198. As a matter of fact,
together with the perovskite SrRuO3, LaNiO3 stands out as the preferred candidate perovskite for a
top or bottom electrode of all-perovskite microelectronic devices. When comparing LaNiO3 and
SrRuO3, we note that SrRuO3 has been reported to have the tendency to develop a rougher surface
at either small (< 30 nm) or large thickness (> 100 nm)239. In such cases, a highly defective and
possibly low dielectric interface may form between the electrode and the active layer, predictably
resulting in degraded functional performance. LaNiO3 seems less affected by these limitations with
the exception of the formation of cracks at very large thickness (> 300 nm), and therefore appears to
be a versatile electrode material. However, the conductive properties of LaNiO3 can be altered in
240,241
ultrathin epitaxial films, as previously discussed . Finally, we note that LaNiO3 has the added
advantage that it is inexpensive compared to other oxides such as SrRuO3. Having said all this, the
integration of LaNiO3 in all-perovskite devices relies on the scalability of high quality LaNiO3 thin
films.
43
the HTS layer and electrically coupling the HTS layer to the metallic tape substrate, for instance in
the so-called RABiTS technology.
While applications of LaNiO3 rely on the fact that its conductivity remains high and rather stable
over a large temperature-range, the other members of the rare earth nickelates attract interest for
their abrupt and sharp MIT and its related properties, such as resistive switching or changes in
optical and opto-electronic properties. Most literature reports concern either SmNiO3 or NdNiO3,
with considerable focus on the former, which has an MIT above room temperature (130 °C)
217
allowing its potential integration into oxide electronics based on CMOS technology . The fine
control of the MI transition temperatures in nickelate thin films by multiple parameters is critical for
developing their potential in various applications.
MITs in TMOs are a longstanding topic in materials physics and chemistry, and the variety of
possible mechanisms triggering them in different materials has been discussed in a number of
59
reviews, in addition to the present one, see e.g. Ref. . The exploration of MI transitions in
applications has also been much discussed in literature and many of the currently suggested device
concepts for nickelates have been inspired by ideas proposed for other TMOs such as VO2 and
related oxides 182. Within this review, we will mainly focus on the advances that have been explored
more recently and that reveal a high potential for new innovative devices, based, for instance, on
resistance switching or opto-electronic properties.
One of the most popular applications discussed in recent literature on nickelates relates to their
memristive behavior, which is closely related to the MIT. Beyond nickelates, nano-scale resistive
switching devices, sometimes termed memristors, have recently generated significant interest for
their use in memory, logic and neuromorphic applications 242.
The term resistance switching is used here to describe the memristive behavior observed in some
nickelate devices. It should not be confused with the resistive switching behavior observed in many
243
oxides (see for instance Ref. ) and related to the formation of filamentary conducting channels
linked to oxygen vacancy formation. Generally speaking, memristors stand in the context of an
increasing exploration, by the electronic engineering community, of novel paradigms for
44
information processing inspired by the operation of the brain, i.e. neuromorphic computation
schemes. In a simplified vision, the brain is based on two types of building blocks: synapses and
neurons. Neurons integrate the information received from other neurons and emit a spike of voltage
when the integrated level reaches a threshold. Synapses connect two neurons and modulate the ease
with which information is transmitted; the tunability of their “transmissivity” or synaptic weight is
at the heart of the learning process. Building a neuromorphic computation architecture thus requires
220
building electronic equivalents of the synapses and of the neurons . While neurons can be
emulated by just one CMOS transistor, synaptic behavior is harder to mimic with classical circuits.
This has motivated the development of a new hardware component able to replicate the behavior of
synapses: the memristor.
Memristors are simple two-terminal analogue nanodevices whose resistance varies with the
voltage (or current) they previously experienced. A memristor is also known as the fourth
244
fundamental circuit element, in addition to the better known resistor, capacitor and inductor .
245
Memristors act as “resistors” with “memory”, hence the name . The analogue response of
memristors, the adjustability of their resistance level and their non-volatile character qualifies them
as electronic synapses. The overall interest of memristors was renewed in 2008 by the discovery of
Pt/TiO2/Pt memristors, in which coupled migration of electrons and ions within the TiO2 layer
246
results in a continuous change of the total device resistance . This voltage-induced displacement
of ions at the nanoscale may hinder reliability and endurance. Another proposed alternative is a
ferroelectric memristor247,248, namely a tunnel junction based on a ferroelectric barrier, whose tunnel
resistance varies greatly depending upon the polarization direction of the barrier (pointing towards
one or the other electrode) and on the associated ferroelectric domain structure. Such ferroelectric
memristors are purely electronic and based on the control of ferroelectric domain structures. Among
the competing memristor concepts we may also cite anionic electronic conducting oxides 249,250.
Within this attractive field where different concepts and material classes compete, oxides with an
MIT have been proposed as alternative candidates 245,251. The first experimental demonstration of an
MI-based memristor was described by Driscoll et al. 245 for VO2 thin films. Importantly, the authors
view their observations from the perspective of “phase-transition-driven memristive systems”,
where the hysteresis with its local characteristics gives the memory aspect of the memristor.
Inspired by this, Ramanathan and co-workers217 have proposed exploiting the MIT in SmNiO3, for
which they report memristive characteristics that are the basis for their subsequent work on synaptic
126 220
transistors or neuromimetic circuits . In a first step, these authors synthesized SmNiO3 thin
films on LaAlO3 and Si substrates to realize systems with charge transport characteristics for both
45
in-plane and in out-of-plane devices217. For the SmNiO3/LaAlO3 structures, they report results that
suggest a history-dependent memristive behavior, opening the door to device applications based on
this resistive switching. In a later work, they propose a SmNiO3-based transistor that mimics a
biological synapse, by doping the SmNiO3 through an ionic liquid 126. Finally, the observation of a
colossal resistance switching by electron doping may be one of the avenues toward larger
conductance changes, opening up a greater scope for future applications 127.
The work on SmNiO3 has motivated similar studies on NdNiO3 heterostructures with the
218
originality of including an asymmetric proton concentration . Altogether, the demonstration of
memristive behavior in nickelates broadens the landscape of materials conceivable for future
memristive applications with the potential to parallel biological neural systems with circuit
220
architecture . Nickelates are competitive in this respect due to the large panel of parameters that
can modulate their electric response. A remaining question is how the presence of a structural phase
transition affects the cyclability.
Finally, as discussed in section 4.1.3, we stress that extrinsic effects such as doping or oxygen
vacancies (natural or induced by an electric-field) are expected to strongly influence the Mott phase
transition by a drastic change of the Ni carriers. This, in turn, offers additional complexity and
tuning parameters for memristors (or other applications).
46
have designed a self-powered photodetector with high sensitivity to light with a wavelength
between 365 nm and 650 nm. It is reported that the performance remains remarkably stable even
205
after 6 months .A similar band gap tuning has been reported for NdNiO3, leading to the claim of
potential for novel photovoltaic devices 209,210.
Other authors127,203 have explored electron doping by lithium or hydrogen intercalation into
SmNiO3 thin films for tuning both the band gap up to 3 eV and the refractive index. Such variations
are spectacular and the fact that such films can be reliably synthesized, are stable at ambient
conditions and have a reversibly controlled level of doping, opens the door to the realization of a
203
variety of photonic devices. Several device concepts have been discussed in Ref. : (i) Electric-
field tunable solid-state devices exploring electrochromic properties. The stability of nickelates in
the presence of oxygen or moisture compares favorably to alternative organic or inorganic
materials. (ii) Devices where the transparency can be reversibly controlled through electron doping
to create reconfigurable photonic devices. Holograms and optical memories have been envisaged.
(iii) Exploring the large range of accessible refractive indices for the control of optical absorption /
transmissivity in smart windows or for radiation control allowing thermal management in space
technologies. (iv) Provided that the electron-doping-induced changes are not accompanied by a
203
structural phase transition (as suggested in Ref. ) this would make possible applications that take
advantage of ultrafast switching such as planar optical modulators. All these device concepts raise
high expectations and ask for an even better understanding to face the recurring challenge of
endurance, which is inherent whenever phase transitions are involved in the technology.
Another way to develop new optical devices is the mastering of defect engineering, as described
in Ref. 204 for developing active optical metasurfaces. Here, the authors use ion irradiation through a
nanometer-scale mask to selectively defect engineer VO2. This is another illustration of the
potential of nano-fabrication, and it is expected that such selective ion beam engineering could also
be of interest for transforming nickelates into photonic structures.
47
Last but not least, more “simple” devices such as bolometers, where the power of incident
electromagnetic radiation is measured via the heating of a material with a temperature-dependent
electrical resistance, such as the nickelates, can also take advantage of the approaches described
previously.
48
speaking, heterostructures are a powerful route toward manipulation of the electronic states
of transition metal oxides by exploiting strain effects, interface-driven phenomena or
reduced dimensionalities. Recent works on nickelate-based superlattices have opened a
variety of avenues to further control their functionalities and even realize novel collective
phases. The characterization and understanding relies on state-of-the-art techniques. To
date, most work has concentrated on LaNiO3-based superlattices and some recent studies
integrate other RNiO3 compounds; in contrast, all-nickelate multilayers (not counting
LaNiO3) have been investigated very little. Also, the original prediction by Chaloupka and
Khallilun of a superconducting phase stabilized in RNiO3 films through confinement and
heterostructuring147 has yet to be confirmed. Advances in material design and fabrication
techniques may allow such ideas to be further explored.
B. Advancement of understanding. In the past few years, a coherent picture of the metal-
insulator transition of perovskite nickelates has emerged, and a notable agreement between
experimental observations, model theories, and first-principles simulations has been
reached. Today there is a deeper understanding of the key role of the O6-breathing
distortions (bond disproportionation) in the insulating state, and the nature of the exotic
"charge order” that accompanies them, which does not involve any significant charge
transfer between nickels but, rather, a self-doping effect resulting in oxygen holes.
Admittedly, some important questions remain open, for instance, regarding the nature (and
onset of) the magnetic order, the possible occurrence of multiferroic effects and the
insulating behavior of ultrathin LaNiO3 films. Yet, in terms of theoretical models and
predictive simulation tools, it seems fair to say that today we are well equipped to tackle
those remaining challenges as well as to explore the wealth of possibilities that nickelate-
based nanostructures and heterostructures potentially offer.
C. Interface physics. Artificial structures, such as multilayers, either purely composed of
nickelates or in combination with other functional perovskites, enrich the interface
physics12 . In such structures, emergent effects, for instance, confinement, proximity
3,4
phenomena, charge transfer and particular local structures, have been observed . We
expect that an expanding theoretical and experimental understanding of exotic electronic,
orbital and spin states at interfaces will extend the discussion here.
D. Applications. A large panel of potential applications for nickelate thin films has been
identified. While some remain at the level of ideas or concepts, more and more reports
demonstrate actual feasibility and the integration of nickelate films into devices with
increasingly complex or innovative architectures. The versatility of nickelate films stems
49
from the fact that their properties - be they electronic, optical or opto-electronic in nature -
can be modulated to a large extent by strain, chemistry, stoichiometry, electron doping,
defect engineering, ionic liquid gating etc., which in turn opens a large window of
functionalities. Interface physics provides the additional parameter space for improving or
even extending the applications discussed for individual films and multilayers. This may
also concern magnetic properties, which have, so far, been little-explored in nickelate-based
devices. Several proposed devices would rely on the epitaxial integration of nickelates with
Si, Ge, and other semiconductors, which is almost unexplored and certainly will need
further attention in the future.
Acknowledgments
JK and JI acknowledge support from the National Research Fund, Luxembourg through a Pearl
Grant (FNR/P12/4853155). This work was supported by the Swiss National Science Foundation
through Division II. The research leading to these results has received funding from the European
Research Council under the European Union’s Seventh Framework Program (FP7/2007-2013)/ERC
Grant Agreement no. 319286 (Q-MAC).
50
References
51
25 H. Park, A. J. Millis, and C. A. Marianetti, Physical Review Letters 109 (15),
156402 (2012).
26 S. Johnston, A. Mukherjee, I. Elfimov, M. Berciu, and G. A. Sawatzky, Physical
Review Letters 112 (10), 106404 (2014).
27 A. Subedi, O. E. Peil, and A. Georges, Physical Review B 91 (7) (2015).
28 J. Varignon, M. N. Grisolia, J. Íñiguez, A. Barthélémy, and M. Bibes, npj Quantum
Materials 2 (1), 21 (2017).
29 S. Prosandeev, L. Bellaiche, and J. Iniguez, Physical Review B 85 (21) (2012).
30 S. Yamamoto and T. Fujiwara, Journal of Physics and Chemistry of Solids 63 (6-
8), 1347 (2002).
31 A. Hampel and C. Ederer, arXiv preprint arXiv:1707.02161 (2017).
32 R. J. Green, M. W. Haverkort, and G. A. Sawatzky, Physical Review B 94 (19),
195127 (2016).
33 J. Perez-Cacho, J. Blasco, J. Garcia, M. Castro, and J. Stankiewicz, Journal of
Physics-Condensed Matter 11 (2), 405 (1999).
34 J. L. Garcia-Muñoz, J. Rodriguez-Carvajal, and P. Lacorre, Physical Review B 50
(2), 978 (1994).
35 J. Rodriguez-Carvajal, S. Rosenkranz, M. Medarde, P. Lacorre, M. T. Fernandez-
Diaz, F. Fauth, and V. Trounov, Physical Review B 57 (1), 456 (1998).
36 J. L. García-Munoz, J. Rodríguez-Carvajal, P. Lacorre, and J. B. Torrance,
Physical Review B 46 (8), 4414 (1992).
37 J. Zhang, H. Zheng, Y. Ren, and J. F. Mitchell, Crystal Growth & Design 17 (5),
2730 (2017).
38 Z. Li, H. Guo, Z. Hu, L. Zhao, C.-Y. Kuo, W. Schmidt, A. Piovano, T. Pi, D. Khomskii,
and L. Tjeng, arXiv preprint arXiv:1705.02589 (2017).
39 J. A. Alonso, M. J. Martinez-Lope, M. T. Casais, M. A. G. Aranda, and M. T.
Fernandez-Diaz, Journal of the American Chemical Society 121 (20), 4754
(1999).
40 J. B. Torrance, P. Lacorre, C. Asavaroengchai, and R. M. Metzger, Physica C:
Superconductivity 182 (4-6), 351 (1991).
41 T. Mizokawa, D. I. Khomskii, and G. A. Sawatzky, Physical Review B 61 (17),
11263 (2000).
42 M. Medarde, A. Fontaine, J. L. García-Muñoz, J. Rodríguez-Carvajal, M. de Santis,
M. Sacchi, G. Rossi, and P. Lacorre, Physical Review B 46 (23), 14975 (1992).
43 V. Scagnoli, U. Staub, A. M. Mulders, M. Janousch, G. I. Meijer, G. Hammerl, J. M.
Tonnerre, and N. Stojic, Physical Review B 73 (10), 100409 (2006).
44 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,
Physical Review Letters 98 (17), 176406 (2007).
45 T. Mizokawa, A. Fujimori, T. Arima, Y. Tokura, N. Mōri, and J. Akimitsu, Physical
Review B 52 (19), 13865 (1995).
46 V. Bisogni, S. Catalano, R. J. Green, M. Gibert, R. Scherwitzl, Y. Huang, V. N.
Strocov, P. Zubko, S. Balandeh, J.-M. Triscone, G. Sawatzky, and T. Schmitt,
Nature Communications 7, 13017 (2016).
52
47 R. S. Dhaka, T. Das, N. C. Plumb, Z. Ristic, W. Kong, C. E. Matt, N. Xu, K. Dolui, E.
Razzoli, M. Medarde, L. Patthey, M. Shi, M. Radović, and J. Mesot, Physical
Review B 92 (3), 035127 (2015).
48 P. D. C. King, H. I. Wei, Y. F. Nie, UchidaM, AdamoC, ZhuS, HeX, BozovicI, D. G.
Schlom, and K. M. Shen, Nature Nanotechnology 9 (6), 443 (2014).
49 R. Eguchi, A. Chainani, M. Taguchi, M. Matsunami, Y. Ishida, K. Horiba, Y. Senba,
H. Ohashi, and S. Shin, Physical Review B 79 (11), 115122 (2009).
50 S. Lee, R. Chen, and L. Balents, Physical Review Letters 106 (1), 016405
(2011).
51 J. Son, B. Jalan, A. P. Kajdos, L. Balents, S. J. Allen, and S. Stemmer, Applied
Physics Letters 99 (19), 192107 (2011).
52 J. Son, P. Moetakef, J. M. LeBeau, D. Ouellette, L. Balents, S. J. Allen, and S.
Stemmer, Applied Physics Letters 96 (6), 062114 (2010).
53 R. Scherwitzl, P. Zubko, I. G. Lezama, S. Ono, A. F. Morpurgo, G. Catalan, and J.-M.
Triscone, Advanced Materials 22 (48), 5517 (2010).
54 S. D. Ha, R. Jaramillo, D. M. Silevitch, F. Schoofs, K. Kerman, J. D. Baniecki, and S.
Ramanathan, Physical Review B 87 (12) (2013).
55 S. W. Cheong, H. Y. Hwang, B. Batlogg, A. S. Cooper, and P. C. Canfield, Physica
B: Condensed Matter 194 (Part 1), 1087 (1994).
56 X. Granados, J. Fontcuberta, X. Obradors, and J. B. Torrance, Physical Review B
46 (24), 15683 (1992).
57 X. Granados, J. Fontcuberta, X. Obradors, L. Mañosa, and J. B. Torrance, Physical
Review B 48 (16), 11666 (1993).
58 N. Gayathri, A. Raychaudhuri, X. Xu, J. Peng, and R. Greene, Journal of Physics:
Condensed Matter 10 (6), 1323 (1998).
59 M. Imada, A. Fujimori, and Y. Tokura, Reviews of Modern Physics 70 (4), 1039
(1998).
60 S. R. Barman, A. Chainani, and D. D. Sarma, Physical Review B 49 (12), 8475
(1994).
61 M. Medarde, P. Lacorre, K. Conder, F. Fauth, and A. Furrer, Physical Review
Letters 80 (11), 2397 (1998).
62 U. Staub, G. I. Meijer, F. Fauth, R. Allenspach, J. G. Bednorz, J. Karpinski, S. M.
Kazakov, L. Paolasini, and F. d'Acapito, Physical Review Letters 88 (12),
126402 (2002).
63 M. Medarde, C. Dallera, M. Grioni, B. Delley, F. Vernay, J. Mesot, M. Sikora, J. A.
Alonso, and M. J. Martínez-Lope, Physical Review B 80 (24), 245105 (2009).
64 J. A. Alonso, M. J. Martínez-Lope, I. A. Presniakov, A. V. Sobolev, V. S. Rusakov, A.
M. Gapochka, G. Demazeau, and M. T. Fernández-Díaz, Physical Review B 87
(18), 184111 (2013).
65 S. Prosandeev, L. Bellaiche, and J. Íñiguez, Physical Review B 85 (21), 214431
(2012).
66 J. L. Garcia-Muñoz, J. Rodriguez-Carvajal, and P. Lacorre, Europhysics Letters
20 (3), 241 (1992).
67 I. Vobornik, L. Perfetti, M. Zacchigna, M. Grioni, G. Margaritondo, J. Mesot, M.
Medarde, and P. Lacorre, Physical Review B 60 (12), R8426 (1999).
53
68 V. Scagnoli, U. Staub, Y. Bodenthin, M. García-Fernández, A. M. Mulders, G. I.
Meijer, and G. Hammerl, Physical Review B 77 (11), 115138 (2008).
69 S. Catalano, M. Gibert, V. Bisogni, O. E. Peil, F. He, R. Sutarto, M. Viret, P. Zubko,
R. Scherwitzl, A. Georges, G. A. Sawatzky, T. Schmitt, and J.-M. Triscone, APL
Materials 2 (11), 116110 (2014).
70 Y. Bodenthin, U. Staub, C. Piamonteze, M. Garcia-Fernandez, M. J. Martinez-
Lope, and J. A. Alonso, Journal of Physics-Condensed Matter 23 (3) (2011).
71 A. Frano, E. Schierle, M. W. Haverkort, Y. Lu, M. Wu, S. Blanco-Canosa, U.
Nwankwo, A. V. Boris, P. Wochner, G. Cristiani, H. U. Habermeier, G. Logvenov,
V. Hinkov, E. Benckiser, E. Weschke, and B. Keimer, Physical Review Letters
111 (10), 106804 (2013).
72 J. S. Zhou and J. B. Goodenough, Physical Review B 69 (15), 153105 (2004).
73 Y. Lu, Z. Zhong, M. W. Haverkort, and P. Hansmann, Physical Review B 95 (19),
195117 (2017).
74 J. Ruppen, J. Teyssier, I. Ardizzone, O. E. Peil, S. Catalano, M. Gibert, J. M.
Triscone, A. Georges, and D. van der Marel, Physical Review B 96 (4), 045120
(2017).
75 X. Obradors, L. M. Paulius, M. B. Maple, J. B. Torrance, A. I. Nazzal, J.
Fontcuberta, and X. Granados, Physical Review B 47 (18), 12353 (1993).
76 P. C. Canfield, J. D. Thompson, S. W. Cheong, and L. W. Rupp, Physical Review B
47 (18), 12357 (1993).
77 J. S. Zhou, J. B. Goodenough, and B. Dabrowski, Physical Review Letters 95 (12)
(2005).
78 J. L. García-Muñoz, M. Amboage, M. Hanfland, J. A. Alonso, M. J. Martínez-Lope,
and R. Mortimer, Physical Review B 69 (9), 094106 (2004).
79 A. Y. Ramos, C. Piamonteze, H. C. N. Tolentino, N. M. Souza-Neto, O. Bunau, Y.
Joly, S. Grenier, J.-P. Itié, N. E. Massa, J. A. Alonso, and M. J. Martinez-Lope, Phys.
Rev. B 85 (4), 045102 (2012).
80 Y. Ohta, T. Tohyama, and S. Maekawa, Physical Review Letters 66 (9), 1228
(1991).
81 H. Park, A. J. Millis, and C. A. Marianetti, Physical Review B 89 (24), 245133
(2014).
82 M. Medarde, J. Mesot, P. Lacorre, S. Rosenkranz, P. Fischer, and K. Gobrecht,
Physical Review B 52 (13), 9248 (1995).
83 J. B. Torrance, P. Lacorre, C. Asavaroengchai, and R. M. Metzger, Physica C 182
(4-6), 351 (1991).
84 G. Giovannetti, S. Kumar, D. Khomskii, S. Picozzi, and J. van den Brink, Physical
Review Letters 103 (15), 156401 (2009).
85 D. Khomskii, Physics 2, 20 (2009).
86 T. Kimura, T. Goto, H. Shintani, K. Ishizaka, T. Arima, and Y. Tokura, Nature 426
(6962), 55 (2003).
87 P. Seth, O. E. Peil, L. Pourovskii, M. Betzinger, C. Friedrich, O. Parcollet, S.
Biermann, F. Aryasetiawan, and A. Georges, arXiv preprint arXiv:1707.09820
(2017).
88 J. Ruppen, J. Teyssier, O. E. Peil, S. Catalano, M. Gibert, J. Mravlje, J. M. Triscone,
A. Georges, and D. van der Marel, Physical Review B 92 (15), 155145 (2015).
54
89 Mercy, Bieder, J. Íñiguez, and P. Ghosez, (to be published).
90 S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J. Humphreys, and A. P. Sutton,
Physical Review B 57 (3), 1505 (1998).
91 R. Car and M. Parrinello, Physical Review Letters 55 (22), 2471 (1985).
92 E. Runge and E. K. U. Gross, Physical Review Letters 52 (12), 997 (1984).
93 S. Catalano, M. Gibert, V. Bisogni, F. He, R. Sutarto, M. Viret, P. Zubko, R.
Scherwitzl, G. A. Sawatzky, T. Schmitt, and J.-M. Triscone, APL Materials 3 (6),
062506 (2015).
94 S. Yamamoto and T. Fujiwara, Journal of the Physical Society of Japan 71 (5),
1226 (2002).
95 H. Raebiger, S. Lany, and A. Zunger, Nature 453 (7196), 763 (2008).
96 G. Demazeau, A. Marbeuf, M. Pouchard, and P. Hagenmuller, Journal of Solid
State Chemistry 3 (4), 582 (1971).
97 G. Catalan, R. M. Bowman, and J. M. Gregg, Physical Review B 62 (12), 7892
(2000).
98 F. Conchon, A. Boulle, R. Guinebretiere, C. Girardot, S. Pignard, J. Kreisel, F.
Weiss, E. Dooryhee, and J. L. Hodeau, Applied Physics Letters 91 (19) (2007).
99 F. Conchon, A. Boulle, R. Guinebretiere, E. Dooryhee, J. L. Hodeau, C. Girardot, S.
Pignard, J. Kreisel, and F. Weiss, Journal of Physics-Condensed Matter 20 (14)
(2008).
100 J. A. Liu, M. Kareev, B. Gray, J. W. Kim, P. Ryan, B. Dabrowski, J. W. Freeland, and
J. Chakhalian, Applied Physics Letters 96 (23) (2010).
101 J. Liu, M. Kargarian, M. Kareev, B. Gray, P. J. Ryan, A. Cruz, N. Tahir, Y.-D.
Chuang, J. Guo, J. M. Rondinelli, J. W. Freeland, G. A. Fiete, and J. Chakhalian,
Nature Communications 4, 2714 (2013).
102 D. Meyers, S. Middey, M. Kareev, M. van Veenendaal, E. J. Moon, B. A. Gray, J. Liu,
J. W. Freeland, and J. Chakhalian, Physical Review B 88 (7), 075116 (2013).
103 F. Y. Bruno, K. Z. Rushchanskii, S. Valencia, Y. Dumont, C. Carrétéro, E. Jacquet,
R. Abrudan, S. Blügel, M. Ležaić, M. Bibes, and A. Barthélémy, Physical Review
B 88 (19), 195108 (2013).
104 P.-H. Xiang, N. Zhong, C.-G. Duan, X. D. Tang, Z. G. Hu, P. X. Yang, Z. Q. Zhu, and J.
H. Chu, Journal of Applied Physics 114 (24), 243713 (2013).
105 M. Hepting, M. Minola, A. Frano, G. Cristiani, G. Logvenov, E. Schierle, M. Wu, M.
Bluschke, E. Weschke, H. U. Habermeier, E. Benckiser, M. Le Tacon, and B.
Keimer, Physical Review Letters 113 (22) (2014).
106 M. K. Stewart, J. Liu, M. Kareev, J. Chakhalian, and D. N. Basov, Physical Review
Letters 107 (17), 176401 (2011).
107 F. Conchon, A. Boulle, R. Guinebretiere, E. Dooryhee, J. L. Hodeau, C. Girardot, S.
Pignard, J. Kreisel, F. Weiss, L. Libralesso, and T. L. Lee, Journal of Applied
Physics 103 (12) (2008).
108 J. Y. Zhang, H. Kim, E. Mikheev, A. J. Hauser, and S. Stemmer, 6, 23652 (2016).
109 E. Mikheev, A. J. Hauser, B. Himmetoglu, N. E. Moreno, A. Janotti, C. G. Van de
Walle, and S. Stemmer, Science Advances 1 (10) (2015).
110 S. Okamoto, Physical Review Letters 110 (6), 066403 (2013).
111 S. Okamoto, W. Zhu, Y. Nomura, R. Arita, D. Xiao, and N. Nagaosa, Physical
Review B 89 (19), 195121 (2014).
55
112 J. L. Blok, X. Wan, G. Koster, D. H. A. Blank, and G. Rijnders, Applied Physics
Letters 99 (15), 151917 (2011).
113 S. Middey, P. Rivero, D. Meyers, M. Kareev, X. Liu, Y. Cao, J. W. Freeland, S.
Barraza-Lopez, and J. Chakhalian, Scientific Reports 4, 6819 (2014).
114 S. Middey, D. Meyers, M. Kareev, E. J. Moon, B. A. Gray, X. Liu, J. W. Freeland, and
J. Chakhalian, Applied Physics Letters 101 (26), 261602 (2012).
115 M. Gibert, P. Zubko, R. Scherwitzl, J. Íñiguez, and J.-M. Triscone, Nature
Materials 11 (3), 195 (2012).
116 T. H. Kim, D. Puggioni, Y. Yuan, L. Xie, H. Zhou, N. Campbell, P. J. Ryan, Y. Choi, J.
W. Kim, J. R. Patzner, S. Ryu, J. P. Podkaminer, J. Irwin, Y. Ma, C. J. Fennie, M. S.
Rzchowski, X. Q. Pan, V. Gopalan, J. M. Rondinelli, and C. B. Eom, Nature 533
(7601), 68 (2016).
117 S. Middey, D. Meyers, D. Doennig, M. Kareev, X. Liu, Y. Cao, Z. Yang, J. Shi, L. Gu,
P. J. Ryan, R. Pentcheva, J. W. Freeland, and J. Chakhalian, Physical Review
Letters 116 (5), 056801 (2016).
118 M. Gibert, M. Viret, P. Zubko, N. Jaouen, J. M. Tonnerre, A. Torres-Pardo, S.
Catalano, A. Gloter, O. Stephan, and J. M. Triscone, Nature Communications 7,
11227 (2016).
119 X. K. Lian, F. Chen, X. L. Tan, P. F. Chen, L. F. Wang, G. Y. Gao, S. W. Jin, and W. B.
Wu, Applied Physics Letters 103, 172110 (2013).
120 F. He, B. O. Wells, Z. G. Ban, S. P. Alpay, S. Grenier, S. M. Shapiro, W. Si, A. Clark,
and X. X. Xi, Physical Review B 70 (23), 235405 (2004).
121 A. Rüegg and G. A. Fiete, Physical Review B 84 (20), 201103 (2011).
122 K.-Y. Yang, W. Zhu, D. Xiao, S. Okamoto, Z. Wang, and Y. Ran, Physical Review B
84 (20), 201104 (2011).
123 D. Doennig, W. E. Pickett, and R. Pentcheva, Phys. Rev. B 89 (12), 121110
(2014).
124 J. L. García-Muñoz, M. Suaaidi, M. J. Martínez-Lope, and J. A. Alonso, Physical
Review B 52 (18), 13563 (1995).
125 I. V. Nikulin, M. A. Novojilov, A. R. Kaul, S. N. Mudretsova, and S. V. Kondrashov,
Oxygen nonstoichiometry of NdNiO3−δ and SmNiO3−δ. (2004), pp.775.
126 J. Shi, S. D. Ha, Y. Zhou, F. Schoofs, and S. Ramanathan, Nature Communications
4 (2013).
127 J. Shi, Y. Zhou, and S. Ramanathan, Nature Communications 5, 4860 (2014).
128 Y. Zhou, X. Guan, H. Zhou, K. Ramadoss, S. Adam, H. Liu, S. Lee, J. Shi, M.
Tsuchiya, D. D. Fong, and S. Ramanathan, Nature 534 (7606), 231 (2016).
129 S. Heo, C. Oh, J. Son, and H. M. Jang, Scientific Reports 7 (1), 4681 (2017).
130 R. Scherwitzl, P. Zubko, C. Lichtensteiger, and J. M. Triscone, Applied Physics
Letters 95 (22), 222114 (2009).
131 J. Son, S. Rajan, S. Stemmer, and S. J. Allen, Journal of Applied Physics 110 (8)
(2011).
132 S. H. Lee, M. Kim, S. D. Ha, J.-W. Lee, S. Ramanathan, and S. Tiwari, Applied
Physics Letters 102 (7), 072102 (2013).
133 M. S. J. Marshall, A. Malashevich, A. S. Disa, M.-G. Han, H. Chen, Y. Zhu, S. Ismail-
Beigi, F. J. Walker, and C. H. Ahn, Physical Review Applied 2 (5), 051001
(2014).
56
134 X. Chen, X. Zhang, M. A. Koten, H. Chen, Z. Xiao, L. Zhang, J. E. Shield, P. A.
Dowben, and X. Hong, Advanced Materials 29 (31), 1701385 (2017).
135 L. Zhang, X. G. Chen, H. J. Gardner, M. A. Koten, J. E. Shield, and X. Hong, Applied
Physics Letters 107 (15), 152906 (2015).
136 A. D. Caviglia, R. Scherwitzl, P. Popovich, W. Hu, H. Bromberger, R. Singla, M.
Mitrano, M. C. Hoffmann, S. Kaiser, P. Zubko, S. Gariglio, J. M. Triscone, M. Först,
and A. Cavalleri, Phys. Rev. Lett. 108 (13), 136801 (2012).
137 W. Hu, S. Catalano, M. Gibert, J. M. Triscone, and A. Cavalleri, Physical Review B
93 (16), 161107 (2016).
138 A. D. Caviglia, M. Först, R. Scherwitzl, V. Khanna, H. Bromberger, R. Mankowsky,
R. Singla, Y. D. Chuang, W. S. Lee, O. Krupin, W. F. Schlotter, J. J. Turner, G. L.
Dakovski, M. P. Minitti, J. Robinson, V. Scagnoli, S. B. Wilkins, S. A. Cavill, M.
Gibert, S. Gariglio, P. Zubko, J. M. Triscone, J. P. Hill, S. S. Dhesi, and A. Cavalleri,
Physical Review B 88 (22), 220401 (2013).
139 M. Forst, A. D. Caviglia, R. Scherwitzl, R. Mankowsky, P. Zubko, V. Khanna, H.
Bromberger, S. B. Wilkins, Y. D. Chuang, W. S. Lee, W. F. Schlotter, J. J. Turner, G.
L. Dakovski, M. P. Minitti, J. Robinson, S. R. Clark, D. Jaksch, J. M. Triscone, J. P.
Hill, S. S. Dhesi, and A. Cavalleri, Nature Materials 14, 883 (2015).
140 M. Först, K. R. Beyerlein, R. Mankowsky, W. Hu, G. Mattoni, S. Catalano, M.
Gibert, O. Yefanov, J. N. Clark, A. Frano, J. M. Glownia, M. Chollet, H. Lemke, B.
Moser, S. P. Collins, S. S. Dhesi, A. D. Caviglia, J. M. Triscone, and A. Cavalleri,
Physical Review Letters 118 (2), 027401 (2017).
141 R. Scherwitzl, S. Gariglio, M. Gabay, P. Zubko, M. Gibert, and J. M. Triscone,
Physical Review Letters 106 (24), 246403 (2011).
142 H. K. Yoo, S. I. Hyun, Y. J. Chang, L. Moreschini, C. H. Sohn, H.-D. Kim, A.
Bostwick, E. Rotenberg, J. H. Shim, and T. W. Noh, Physical Review B 93 (3),
035141 (2016).
143 J. Fowlie, M. Gibert, G. Tieri, A. Gloter, J. Íñiguez, A. Filippetti, S. Catalano, S.
Gariglio, A. Schober, M. Guennou, J. Kreisel, O. Stéphan, and J.-M. Triscone,
Advanced Materials 29 (18), 1605197 (2017).
144 D. P. Kumah, A. S. Disa, J. H. Ngai, H. Chen, A. Malashevich, J. W. Reiner, S. Ismail-
Beigi, F.-J. Walker, and C. H. Ahn, Advanced Materials 26 (12), 1935 (2014).
145 M. C. Weber, M. Guennou, N. Dix, D. Pesquera, F. Sánchez, G. Herranz, J.
Fontcuberta, L. López-Conesa, S. Estradé, F. Peiró, J. Iñiguez, and J. Kreisel,
Physical Review B 94 (1), 014118 (2016).
146 J. Chakhalian, J. M. Rondinelli, J. Liu, B. A. Gray, M. Kareev, E. J. Moon, N. Prasai, J.
L. Cohn, M. Varela, I. C. Tung, M. J. Bedzyk, S. G. Altendorf, F. Strigari, B.
Dabrowski, L. H. Tjeng, P. J. Ryan, and J. W. Freeland, Physical Review Letters
107 (11), 116805 (2011).
147 J. Chaloupka and G. Khaliullin, Physical Review Letters 100 (1), 016404
(2008).
148 S. J. May, J. W. Kim, J. M. Rondinelli, E. Karapetrova, N. A. Spaldin, A.
Bhattacharya, and P. J. Ryan, Physical Review B 82 (1), 014110 (2010).
149 O. E. Peil, M. Ferrero, and A. Georges, Physical Review B 90 (4), 045128
(2014).
57
150 P. Hansmann, A. Toschi, X. Yang, O. K. Andersen, and K. Held, Physical Review B
82 (23), 235123 (2010).
151 P. Hansmann, X. Yang, A. Toschi, G. Khaliullin, O. K. Andersen, and K. Held,
Physical Review Letters 103 (1), 016401 (2009).
152 M. Wu, E. Benckiser, M. W. Haverkort, A. Frano, Y. Lu, U. Nwankwo, S. Brück, P.
Audehm, E. Goering, S. Macke, V. Hinkov, P. Wochner, G. Christiani, S. Heinze, G.
Logvenov, H. U. Habermeier, and B. Keimer, Physical Review B 88 (12),
125124 (2013).
153 E. Benckiser, M. W. Haverkort, S. Brück, E. Goering, S. Macke, A. Frañó, X. Yang,
O. K. Andersen, G. Cristiani, H.-U. Habermeier, A. V. Boris, I. Zegkinoglou, P.
Wochner, H.-J. Kim, V. Hinkov, and B. Keimer, Nature Materials 10 (3), 189
(2011).
154 J. W. Freeland, J. Liu, M. Kareev, B. Gray, J. W. Kim, P. Ryan, R. Pentcheva, and J.
Chakhalian, Europhysics Letters 96, 57004 (2011).
155 H. Chen, D. P. Kumah, A. S. Disa, F. J. Walker, C. H. Ahn, and S. Ismail-Beigi,
Physical Review Letters 110 (18), 186402 (2013).
156 A. S. Disa, D. P. Kumah, A. Malashevich, H. Chen, D. A. Arena, E. D. Specht, S.
Ismail-Beigi, F. J. Walker, and C. H. Ahn, Physical Review Letters 114 (2),
026801 (2015).
157 M. J. Han, X. Wang, C. A. Marianetti, and A. J. Millis, Physical Review Letters 107
(20), 206804 (2011).
158 F. Y. Bruno, M. Gibert, S. M. Walker, O. E. Peil, A. d. l. Torre, S. Riccò, Z. Wang, S.
Catalano, A. Tamai, F. Bisti, V. N. Strocov, J.-M. Triscone, and F. Baumberger,
APL Materials 5 (1), 016101 (2017).
159 S. J. May, T. S. Santos, and A. Bhattacharya, Physical Review B 79 (11), 115127
(2009).
160 J. Hwang, J. Y. Zhang, J. Son, and S. Stemmer, Applied Physics Letters 100 (19)
(2012).
161 A. V. Boris, Y. Matiks, E. Benckiser, A. Frano, P. Popovich, V. Hinkov, P. Wochner,
M. Castro-Colin, E. Detemple, V. K. Malik, C. Bernhard, T. Prokscha, A. Suter, Z.
Salman, E. Morenzoni, G. Cristiani, H.-U. Habermeier, and B. Keimer, Science
332 (6032), 937 (2011).
162 Y. Lu, A. Frano, M. Bluschke, M. Hepting, S. Macke, J. Strempfer, P. Wochner, G.
Cristiani, G. Logvenov, H. U. Habermeier, M. W. Haverkort, B. Keimer, and E.
Benckiser, Physical Review B 93 (16), 165121 (2016).
163 H. Tanaka, N. Okawa, and T. Kawai, Solid State Communications 110 (4), 191
(1999).
164 K. R. Nikolaev, A. Y. Dobin, I. N. Krivorotov, W. K. Cooley, A. Bhattacharya, A. L.
Kobrinskii, L. I. Glazman, R. M. Wentzovitch, E. D. Dahlberg, and A. M. Goldman,
Physical Review Letters 85 (17), 3728 (2000).
165 J. Hoffman, I. C. Tung, B. B. Nelson-Cheeseman, M. Liu, J. W. Freeland, and A.
Bhattacharya, Physical Review B 88 (14), 144411 (2013).
166 A. J. Grutter, H. Yang, B. J. Kirby, M. R. Fitzsimmons, J. A. Aguiar, N. D. Browning,
C. A. Jenkins, E. Arenholz, V. V. Mehta, U. S. Alaan, and Y. Suzuki, Physical
Review Letters 111 (8), 087202 (2013).
58
167 C. Piamonteze, M. Gibert, J. Heidler, J. Dreiser, S. Rusponi, H. Brune, J. M.
Triscone, F. Nolting, and U. Staub, Physical Review B 92 (1), 014426 (2015).
168 J. C. Rojas Sánchez, B. Nelson-Cheeseman, M. Granada, E. Arenholz, and L. B.
Steren, Physical Review B 85 (9), 094427 (2012).
169 M. Gibert, M. Viret, A. Torres-Pardo, C. Piamonteze, P. Zubko, N. Jaouen, J. M.
Tonnerre, A. Mougin, J. Fowlie, S. Catalano, A. Gloter, O. Stéphan, and J. M.
Triscone, Nano Letters 15 (11), 7355 (2015).
170 S. Dong and E. Dagotto, Physical Review B 87 (19), 195116 (2013).
171 A. T. Lee and M. J. Han, Physical Review B 88 (3), 035126 (2013).
172 K. R. Nikolaev, A. Bhattacharya, P. A. Kraus, V. A. Vas’ko, W. K. Cooley, and A. M.
Goldman, Applied Physics Letters 75 (1), 118 (1999).
173 J. D. Hoffman, B. J. Kirby, J. Kwon, G. Fabbris, D. Meyers, J. W. Freeland, I. Martin,
O. G. Heinonen, P. Steadman, H. Zhou, C. M. Schlepütz, M. P. M. Dean, S. G. E. te
Velthuis, J.-M. Zuo, and A. Bhattacharya, Physical Review X 6 (4), 041038
(2016).
174 M. N. Grisolia, J. Varignon, G. Sanchez-Santolino, A. Arora, S. Valencia, M. Varela,
R. Abrudan, E. Weschke, E. Schierle, J. E. Rault, J. P. Rueff, A. Barthelemy, J.
Santamaria, and M. Bibes, Nat Phys 12 (5), 484 (2016).
175 H. Chen, A. J. Millis, and C. A. Marianetti, Physical Review Letters 111 (11),
116403 (2013).
176 Y. W. Cao, X. R. Liu, M. Kareev, D. Choudhury, S. Middey, D. Meyers, J. W. Kim, P.
J. Ryan, J. W. Freeland, and J. Chakhalian, Nature Communications 7 (2016).
177 C. Girardot, S. Pignard, F. Weiss, and J. Kreisel, Appl. Phys. Lett. 98, 241903
(2011).
178 E. Dagotto, Science 309 (5732), 257 (2005).
179 M. Nakano, K. Shibuya, D. Okuyama, T. Hatano, S. Ono, M. Kawasaki, Y. Iwasa,
and Y. Tokura, Nature 487 (7408), 459 (2012).
180 J. Jeong, N. Aetukuri, T. Graf, T. D. Schladt, M. G. Samant, and S. S. P. Parkin,
Science 339 (6126), 1402 (2013).
181 N. Shukla, A. V. Thathachary, A. Agrawal, H. Paik, A. Aziz, D. G. Schlom, S. K.
Gupta, R. Engel-Herbert, and S. Datta, Nature Communications 6 (2015).
182 Z. Yang, C. Y. Ko, and S. Ramanathan, in Annual Review of Materials Research,
Vol 41, edited by D. R. Clarke and P. Fratzl (2011), Vol. 41, pp. 337.
183 R. N. Singh, L. Bahadur, J. P. Pandey, S. P. Singh, P. Chartier, and G. Poillerat,
Journal of Applied Electrochemistry 24 (2), 149 (1994).
184 Y. Cao, B. P. Lin, Y. Sun, H. Yang, and X. Q. Zhang, Electrochimica Acta 174, 41
(2015).
185 L. Hu, Y. F. Deng, K. Liang, X. J. Liu, and W. C. Hu, Journal of Solid State
Electrochemistry 19 (3), 629 (2015).
186 D. K. Hwang, S. Kim, J. H. Lee, I. S. Hwang, and I. D. Kim, Journal of Materials
Chemistry 21 (6), 1959 (2011).
187 K. Liang, N. Wang, M. Zhou, Z. Y. Cao, T. L. Gu, Q. Zhang, X. Z. Tang, W. C. Hu, and
B. Q. Wei, Journal of Materials Chemistry A 1 (34), 9730 (2013).
188 X. Liu, G. Du, J. L. Zhu, Z. F. Zeng, and X. H. Zhu, Applied Surface Science 384, 92
(2016).
59
189 F. Y. Bruno, S. Boyn, S. Fusil, S. Girod, C. Carretero, M. Marinova, A. Gloter, S.
Xavier, C. Deranlot, M. Bibes, A. Barthelemy, and V. Garcia, Advanced Electronic
Materials 2 (3) (2016).
190 B. V. Mistry, R. Pinto, and U. S. Joshi, Journal of Materials Science-Materials in
Electronics 27 (2), 1812 (2016).
191 H. N. Xu, Y. Liu, B. Xu, Y. D. Xia, G. S. Wang, J. Yin, and Z. G. Liu, Journal of Physics
D-Applied Physics 49 (37) (2016).
192 Z. W. Dong, S. P. Pai, R. Ramesh, T. Venkatesan, M. Johnson, Z. Y. Chen, A.
Cavanaugh, Y. G. Zhao, X. L. Jiang, R. P. Sharma, S. Ogale, and R. L. Greene,
Journal of Applied Physics 83 (11), 6780 (1998).
193 D. H. Bao, N. Wakiya, K. Shinozaki, N. Mizutani, and X. Yao, Journal of Applied
Physics 90 (1), 506 (2001).
194 S. T. Zhang, X. J. Zhang, H. W. Cheng, Y. F. Chen, Z. G. Liu, N. B. Ming, X. B. Hu, and
J. Y. Wang, Applied Physics Letters 83 (21), 4378 (2003).
195 J. Ge, X. L. Dong, Y. Chen, F. Cao, and G. S. Wang, Applied Physics Letters 102
(14) (2013).
196 H. Han, J. Zhong, S. Kotru, P. Padmini, X. Y. Song, and R. K. Pandey, Applied
Physics Letters 88 (9) (2006).
197 A. D. Li, C. Z. Ge, P. Lu, D. Wu, S. B. Xiong, and N. B. Ming, Applied Physics
Letters 70 (12), 1616 (1997).
198 K. V. R. Prasad, K. B. R. Varma, A. R. Raju, K. M. Satyalakshmi, R. M. Mallya, and
M. S. Hegde, Applied Physics Letters 63 (14), 1898 (1993).
199 K. Kim, M. Paranthaman, D. P. Norton, T. Aytug, C. Cantoni, A. A. Gapud, A.
Goyal, and D. K. Christen, Superconductor Science & Technology 19 (4), R23
(2006).
200 G. Z. Sun, P. H. Wu, Y. Feng, Z. M. Ji, S. Z. Yang, M. Wang, W. W. Xu, and L. Kang,
Thin Solid Films 471 (1-2), 248 (2005).
201 T. Aytug, J. Z. Wu, C. Cantoni, D. T. Verebelyi, E. D. Specht, M. Paranthaman, D. P.
Norton, D. K. Christen, R. E. Ericson, and C. L. Thomas, Applied Physics Letters
76 (6), 760 (2000).
202 Q. He, D. K. Christen, R. Feenstra, D. P. Norton, M. Paranthaman, E. D. Specht, D.
F. Lee, A. Goyal, and D. M. Kroeger, Physica C 314 (1-2), 105 (1999).
203 Z. Y. Li, Y. Zhou, H. Qi, Q. W. Pan, Z. Zhang, N. N. Shi, M. Lu, A. Stein, C. Y. Li, S.
Ramanathan, and N. F. Yu, Advanced Materials 28 (41), 9117 (2016).
204 J. Rensberg, S. Zhang, Y. Zhou, A. S. McLeod, C. Schwarz, M. Goldflam, M. K. Liu, J.
Kerbusch, R. Nawrodt, S. Ramanathan, D. N. Basov, F. Capasso, C. Ronning, and
M. A. Kats, Nano Letters 16 (2), 1050 (2016).
205 L. Wang, L. Chang, X. Yin, L. You, J.-L. Zhao, H. Guo, K. Jin, K. Ibrahim, J. Wang, A.
Rusydi, and J. Wang, Applied Physics Letters 110 (4), 043504 (2017).
206 B. D. Ngom, J. B. K. Kana, O. Nemraoui, N. Manyala, M. Maaza, R. Mdjoe, and A. C.
Beye, in Laser and Plasma Applications in Materials Science, edited by E. H.
Amara, S. Boudjemai, and D. Doumaz (2008), Vol. 1047, pp. 280.
207 T. Shao, Z. M. Qi, Y. Y. Wang, Y. Y. Li, M. Yang, Y. Wang, G. B. Zhang, and M. Liu,
Applied Physics Letters 107 (2) (2015).
208 P. Kiria, G. Hyett, and R. Binions, Advanced Materials Letters 1 (2), 86 (2010).
60
209 L. Chang, L. Wang, L. You, Y. Zhou, L. Fang, S. W. Wang, and J. L. Wang, Journal
of Physics D-Applied Physics 49 (44) (2016).
210 L. Wang, S. Dash, L. Chang, L. You, Y. Q. Peng, X. He, K. J. Jin, Y. Zhou, H. G. Ong, P.
Ren, S. W. Wang, L. Chen, and J. L. Wang, Acs Applied Materials & Interfaces 8
(15), 9769 (2016).
211 Y. Zhou, X. F. Guan, H. Zhou, K. Ramadoss, S. Adam, H. J. Liu, S. Lee, J. Shi, M.
Tsuchiya, D. D. Fong, and S. Ramanathan, Nature 534 (7606), 231 (2016).
212 J. B. Goodenough, Reports on Progress in Physics 67 (11), 1915 (2004).
213 H. Y. Ma, B. G. Wang, Y. S. Fan, and W. C. Hong, Energies 7 (10), 6549 (2014).
214 J. Zhang, Y. B. Zhao, X. Zhao, Z. L. Liu, and W. Chen, Scientific Reports 4, 6
(2014).
215 J. Ma, Y. Liu, Y. Liu, Y. Yan, and P. Zhang, Fuel Cells 8 (6), 394 (2008).
216 R. N. Singh, A. N. Jain, S. K. Tiwari, G. Poillerat, and P. Chartier, Journal of
Applied Electrochemistry 25 (12), 1133 (1995).
217 S. D. Ha, G. H. Aydogdu, and S. Ramanathan, Applied Physics Letters 98 (1)
(2011).
218 C. Oh, S. Heo, H. M. Jang, and J. Son, Applied Physics Letters 108 (12) (2016).
219 P. Pandey, R. Rana, S. Tripathi, and D. S. Rana, Applied Physics Letters 103 (3)
(2013).
220 S. D. Ha, J. Shi, Y. Meroz, L. Mahadevan, and S. Ramanathan, Physical Review
Applied 2 (6) (2014).
221 S. H. Lee, M. Kim, S. D. Ha, J.-W. Lee, S. Ramanathan, and S. Tiwari, Applied
Physics Letters 102 (7), 072102 (2013).
222 S. Asanuma, P. H. Xiang, H. Yamada, H. Sato, I. H. Inoue, H. Akoh, A. Sawa, K.
Ueno, H. Shimotani, H. Yuan, M. Kawasaki, and Y. Iwasa, Applied Physics
Letters 97 (14) (2010).
223 T. R. Ling, Z. B. Chen, and M. D. Lee, Catalysis Today 26 (1), 79 (1995).
224 X. C. Lu, T. X. Xu, and X. H. Dong, Sensors and Actuators B-Chemical 67 (1-2),
24 (2000).
225 R. Bouregba, G. Le Rhun, and G. Poullain, Integrated Ferroelectrics 70, 67
(2005).
226 F. Hou, Y. N. Qin, and T. X. Xu, Rare Metal Materials and Engineering 31, 296
(2002).
227 F. Hou, Y. N. Qin, T. X. Xu, and M. X. Xu, Journal of Electroceramics 8 (3), 243
(2002).
228 T. Kobayashi, R. Kondou, K. Nakamura, M. Ichiki, and R. Maeda, Japanese
Journal of Applied Physics Part 1-Regular Papers Brief Communications &
Review Papers 46 (10B), 7073 (2007).
229 A. Lazauskas, J. Baltrusaitis, V. Grigaliunas, and I. Prosycevas, Applied Surface
Science 344, 159 (2015).
230 H. L. Lung, S. C. Lai, H. Y. Lee, T. B. Wu, R. Liu, and C. Y. Lu, Ieee Transactions on
Electron Devices 51 (6), 920 (2004).
231 C. Suo, C. J. Gao, X. Y. Wu, Y. Zuo, X. C. Wang, and J. F. Jia, Rsc Advances 5 (112),
92107 (2015).
232 M. Sygnatowicz, M. Snure, and A. Tiwari, Journal of Crystal Growth 310 (15),
3590 (2008).
61
233 B. J. Wang, S. Q. Gu, Y. P. Ding, Y. L. Chu, Z. Zhang, X. Ba, Q. L. Zhang, and X. R. Li,
Analyst 138 (1), 362 (2013).
234 K. Ramadoss, N. Mandal, X. Dai, Z. Wan, Y. Zhou, L. Rokhinson, Y. P. Chen, J. P.
Hu, and S. Ramanathan, Physical Review B 94 (23), 6 (2016).
235 S. W. Cheong and M. Mostovoy, Nature Materials 6, 13 (2007).
236 J. V. d. Brink and D. I. Khomskii, Journal of Physics: Condensed Matter 20 ( 43),
249 (2008).
237 G. Catalan, Phase Transitions 81, 729 (2008).
238 S. D. Ha, J. Shi, Y. Meroz, L. Mahadevan, and S. Ramanathan, Physical Review
Applied 2 (6), 064003 (2014).
239 C. Guerrero, F. Sánchez, C. Ferrater, J. Roldán, M. V. Garcı́a-Cuenca, and M.
Varela, Applied Surface Science 168 (1), 219 (2000).
240 R. Scherwitzl, S. Gariglio, M. Gabay, P. Zubko, M. Gibert, and J. M. Triscone,
Physical Review Letters 106 (24) (2011).
241 J. Fowlie, M. Gibert, G. Tieri, A. Gloter, J. Íñiguez, A. Filippetti, S. Catalano, S.
Gariglio, A. Schober, M. Guennou, J. Kreisel, O. Stéphan, and J.-M. Triscone,
Advanced Materials, DOI: 10.1002/adma.201605197 (2017).
242 S. H. Jo, T. Chang, I. Ebong, B. B. Bhadviya, P. Mazumder, and W. Lu, Nano
Letters 10 (4), 1297 (2010).
243 A. Sawa, Materials Today 11 (6), 28 (2008).
244 C. L. O., IEEE Trans. Circuit Theory 18, 507 (1971).
245 T. Driscoll, H.-T. Kim, B.-G. Chae, M. D. Ventra, and D. N. Basov, Applied Physics
Letters 95 (4), 043503 (2009).
246 D. B. Strukov, G. S. Snider, D. R. Stewart, and R. S. Williams, Nature 453 (7191),
80 (2008).
247 A. Chanthbouala, V. Garcia, R. O. Cherifi, K. Bouzehouane, S. Fusil, X. Moya, S.
Xavier, H. Yamada, C. Deranlot, N. D. Mathur, M. Bibes, A. Barthélémy, and J.
Grollier, Nat Mater 11 (10), 860 (2012).
248 D. J. Kim, H. Lu, S. Ryu, C. W. Bark, C. B. Eom, E. Y. Tsymbal, and A. Gruverman,
Nano Letters 12 (11), 5697 (2012).
249 F. Messerschmitt, M. Kubicek, S. Schweiger, and J. L. M. Rupp, Advanced
Functional Materials 24 (47), 7448 (2014).
250 S. Schweiger, M. Kubicek, F. Messerschmitt, C. Murer, and J. L. M. Rupp, Acs
Nano 8 (5), 5032 (2014).
251 M. D. Ventra, Y. V. Pershin, and L. O. Chua, Proceedings of the IEEE 97 (10),
1717 (2009).
252 A. Boileau, F. Capon, P. Laffez, S. Barrat, J. L. Endrino, R. E. Galindo, D. Horwat,
and J. F. Pierson, The Journal of Physical Chemistry C 118 (11), 5908 (2014).
253 C. Napierala - Dutfoy, M. Edely, P. Laffez, and L. Sauques, Thermo-optical effect
of Nd0.3Sm0.7NiO3 ceramic in the infrared range. (2009), pp.1498.
254 K. Ramadoss, N. Mandal, X. Dai, Z. Wan, Y. Zhou, L. Rokhinson, Y. P. Chen, J. P.
Hu, and S. Ramanathan, Physical Review B 94 (23) (2016).
62
63