HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.

  • failed: chemformula

Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.

License: CC BY-NC-SA 4.0
arXiv:2402.00747v1 [cond-mat.str-el] 01 Feb 2024

Mott resistive switching initiated by topological defects

Alessandra Milloch [email protected] Department of Mathematics and Physics, Università Cattolica del Sacro Cuore, Brescia I-25133, Italy ILAMP (Interdisciplinary Laboratories for Advanced Materials Physics), Università Cattolica del Sacro Cuore, Brescia I-25133, Italy Department of Physics and Astronomy, KU Leuven, B-3001 Leuven, Belgium    Ignacio Figueruelo-Campanero [email protected] IMDEA Nanociencia, Cantoblanco, 28049 Madrid, Spain Facultad Ciencias Físicas, Universidad Complutense, 28040 Madrid, Spain    Wei-Fan Hsu Department of Physics and Astronomy, KU Leuven, B-3001 Leuven, Belgium    Selene Mor Department of Mathematics and Physics, Università Cattolica del Sacro Cuore, Brescia I-25133, Italy ILAMP (Interdisciplinary Laboratories for Advanced Materials Physics), Università Cattolica del Sacro Cuore, Brescia I-25133, Italy    Simon Mellaerts Department of Physics and Astronomy, KU Leuven, B-3001 Leuven, Belgium    Francesco Maccherozzi Diamond Light Source, Didcot, Oxfordshire OX11 0DE, UK    Larissa Ishibe Veiga Diamond Light Source, Didcot, Oxfordshire OX11 0DE, UK    Sarnjeet S. Dhesi Diamond Light Source, Didcot, Oxfordshire OX11 0DE, UK    Mauro Spera Department of Mathematics and Physics, Università Cattolica del Sacro Cuore, Brescia I-25133, Italy    ** Won Seo Department of Materials Engineering, KU Leuven, 3001 Leuven, Belgium    Jean-Pierre Locquet Department of Physics and Astronomy, KU Leuven, B-3001 Leuven, Belgium    Michele Fabrizio Scuola Internazionale Superiore di Studi Avanzati (SISSA), Via Bonomea 265, 34136 Trieste, Italy    Mariela Menghini IMDEA Nanociencia, Cantoblanco, 28049 Madrid, Spain    Claudio Giannetti [email protected] Department of Mathematics and Physics, Università Cattolica del Sacro Cuore, Brescia I-25133, Italy ILAMP (Interdisciplinary Laboratories for Advanced Materials Physics), Università Cattolica del Sacro Cuore, Brescia I-25133, Italy CNR-INO (National Institute of Optics), via Branze 45, 25123 Brescia, Italy
Abstract

Resistive switching is the fundamental process that triggers the sudden change of the electrical properties in solid-state devices under the action of intense electric fields [1]. Despite its relevance for information processing, ultrafast electronics, neuromorphic devices, resistive memories and brain-inspired computation [2, 3, 4, 5, 6, 7, 1, 8, 9, 10, 11, 12, 13, 14], the nature of the local stochastic fluctuations that drive the formation of metallic nuclei out of the insulating state has remained hidden.

Here, using operando𝑜𝑝𝑒𝑟𝑎𝑛𝑑𝑜operandoitalic_o italic_p italic_e italic_r italic_a italic_n italic_d italic_o X-ray nano-imaging, we have captured the early-stages of resistive switching in a V22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT-based device under working conditions. V22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT is a paradigmatic Mott material [3], which undergoes a first-order metal-to-insulator transition coupled to a lattice transformation that breaks the threefold rotational symmetry of the rhombohedral metal phase [2, 5, 6, 8, 9, 10, 11, 15]. We reveal a new class of volatile electronic switching triggered by nanoscale topological defects of the lattice order parameter of the insulating phase. Our results pave the way to the use of strain engineering approaches to manipulate topological defects and achieve the full control of the electronic Mott switching. The concept of topology-driven reversible electronic transition is of interest for a broad class of quantum materials, comprising transition metal oxides, chalcogenides and kagome metals, that exhibit first-order electronic transitions coupled to a symmetry-breaking order.

Refer to caption
Figure 1: a) Non-primitive hexagonal unit cell of the high-temperature rhombohedral phase. b) Sketch of the rhombohedral-to-monoclinic distortion taking place along each of the three equivalent hexagonal axes. c) PEEM experimental setup. X-ray radiation with tunable energy resonant with the vanadium L2,323{}_{2,3}start_FLOATSUBSCRIPT 2 , 3 end_FLOATSUBSCRIPT edge im**es on the sample and the emitted electrons are collected and imaged through electrostatic and magnetic lenses. The \chV_2O_3 film of thickness t𝑡titalic_t = 20 nm is coated with gold metal electrodes separated by a 2 µm width (w𝑤witalic_w) ×\times× 50 µm length (l𝑙litalic_l) gap, allowing to drive a current through the device while simultaneously acquiring XLD-PEEM images.

The insulator-to-metal transition (IMT) in Mott materials is a key mechanism for the development of the next generation Mottronic devices [3, 13]. The intrinsic correlated nature of the Mott insulating state makes these systems fragile to external stimuli [16, 17], such as the application of an electric field, which can drive the collapse of the electronic band structure and the sudden release of a large number of free carriers [18, 19]. At the macroscopic level, this phenomenon manifests in the resistive switching process, i.e., a sharp increase of the current flow when the applied voltage exceeds a threshold value [20, 21, 22, 6, 23, 24, 25, 26, 27]. This strong non-linearity triggered many efforts to develop neuromorphic building blocks for the hardware implementation of neural networks [14] or for ultrafast volatile and non-volatile memories or processors [28, 29, 12]. The state-of-the-art macroscopic models [30] are based on resistor networks that consider interconnected nodes transforming from the insulating to metallic state in the presence of an electric field. Above a certain threshold, a percolative transition takes place, thus leading to the formation of a conductive filament and a sudden resistivity drop [22, 31].

The full control and, therefore, the exploitation of this process is prevented by the limited knowledge of the early-stage firing dynamics. Microscopically, little is known about the nature of the nanoscale regions that trigger the avalanche process. Also the relation between the electronic and structural properties of the switched regions and those of the pristine insulating template is a matter of debate. Pioneering optical microscopy experiments captured the real-time formation of macroscopic metallic channels [4, 32, 33, 34, 35], while lacking the resolution and sensitivity to address the origin of the switching process.

Here, we introduce a resonant X-ray microscopy to take nanoscale snapshots of the switching dynamics at the nanoscale during the application of an electric field to a V22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT-based nanodevice. Our results unveil the fundamental role played by the topology of the underlying lattice nanotexture. The breaking of the C3subscript𝐶3C_{3}italic_C start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT symmetry upon transition to the insulating monoclinic phase leads to the formation of three twin domains whose boundaries are oriented along the three hexagonal directions [36, 37]. The geometrical constraints produce strain topological defects at the corners of monoclinic domains crossing with an angle of 60{}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPT. These nanoscale topological defects act as seeds of the metallic phase, thus triggering the macroscopic volatile switching. The nature of the transient metallic state is discussed in view of achieving ultrafast and reversible all-electronic switching.

\ch

V_2O_3 is a prototypical Mott insulator that undergoes a thermally-driven transition from a high-temperature paramagnetic metal with rhombohedral lattice symmetry to a low-temperature antiferromagnetic insulator with monoclinic structure [38, 39, 40]. The lattice transformation across the critical temperature TIMTsubscript𝑇𝐼𝑀𝑇T_{IMT}italic_T start_POSTSUBSCRIPT italic_I italic_M italic_T end_POSTSUBSCRIPT implies the breaking of the C3subscript𝐶3C_{3}italic_C start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT symmetry of the non-primitive hexagonal unit cell of the high-temperature phase (see Fig. 1a). The structural transition can thus be described [37] by a vector order parameter:

ϵ=(ϵ31,ϵ23)=ϵ(cosϕn,sinϕn)italic-ϵsubscriptitalic-ϵ31subscriptitalic-ϵ23italic-ϵcossubscriptitalic-ϕ𝑛sinsubscriptitalic-ϕ𝑛\overrightarrow{\epsilon}=(\epsilon_{31},\epsilon_{23})=\epsilon\,\big{(}% \mathrm{cos}\phi_{n},\mathrm{sin}\phi_{n}\big{)}over→ start_ARG italic_ϵ end_ARG = ( italic_ϵ start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT , italic_ϵ start_POSTSUBSCRIPT 23 end_POSTSUBSCRIPT ) = italic_ϵ ( roman_cos italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , roman_sin italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) (1)

associated to the shear strain components ϵ31subscriptitalic-ϵ31\epsilon_{31}italic_ϵ start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT and ϵ23subscriptitalic-ϵ23\epsilon_{23}italic_ϵ start_POSTSUBSCRIPT 23 end_POSTSUBSCRIPT that characterize the monoclinic distortion. Below TIMTsubscript𝑇𝐼𝑀𝑇T_{IMT}italic_T start_POSTSUBSCRIPT italic_I italic_M italic_T end_POSTSUBSCRIPT, the amplitude of the order parameter, ϵitalic-ϵ\epsilonitalic_ϵ, becomes non-zero, while the phase can assume three different values:

ϕn=(2n1)π3subscriptitalic-ϕ𝑛2𝑛1𝜋3\phi_{n}=(2n-1)\frac{\pi}{3}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ( 2 italic_n - 1 ) divide start_ARG italic_π end_ARG start_ARG 3 end_ARG (2)

corresponding to the distortion along the three equivalent hexagonal axes of the rhombohedral phase, indicated in the following by the versors ϵ^nsubscript^italic-ϵ𝑛\hat{\epsilon}_{n}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, n𝑛nitalic_n=1,2,3 (see Fig. 1b).

Resistive switching from the insulating to the metallic state can be induced by applying an electric field across patterned micro-gaps at temperatures close to TIMTsubscript𝑇𝐼𝑀𝑇T_{IMT}italic_T start_POSTSUBSCRIPT italic_I italic_M italic_T end_POSTSUBSCRIPT [41, 4, 33]. The resistive switching device investigated here is formed by a 20 nm \chV_2O_3 film coated with gold electrodes. \chV_2O_3 is grown by oxygen-assisted Molecular Beam Epitaxy on a (0001)-\chAl_2O_3 substrate with a 40 nm \chCr_2O_3 buffer layer to reduce any interfacial residual strain [42]. The resulting \chV_2O_3 film has the c𝑐citalic_c axis oriented parallel to the surface normal and undergoes the IMT at TIMTsubscript𝑇𝐼𝑀𝑇T_{IMT}italic_T start_POSTSUBSCRIPT italic_I italic_M italic_T end_POSTSUBSCRIPT = 145 K (see Supplementary Information Fig. S1). Two gold electrodes allow the application of an electric bias across the gap of width w𝑤witalic_w = 2 µm and length l𝑙litalic_l = 30 µm (figure 1c). The gap region between the electrodes is imaged using Photo Emission Electron Microscopy (PEEM), combined with X-ray Linear Dichroism (XLD) at the \chL_2,3 vanadium edge (513-530 eV, see Figure S2) [43, 36, 37]. The XLD-PEEM images are obtained from the normalized difference between images recorded with the light electric field vector, E𝐸\overrightarrow{E}over→ start_ARG italic_E end_ARG, parallel and perpendicular to the surface normal at a photon energy 520.6 eV. Since the XLD signal depends on the angle between the in-plane component of E𝐸\overrightarrow{E}over→ start_ARG italic_E end_ARG and ϵ(𝐫)italic-ϵ𝐫\overrightarrow{\epsilon}(\mathbf{r})over→ start_ARG italic_ϵ end_ARG ( bold_r ) [37] (see Fig. 1b and c), this technique provides a map - with similar-to\sim30 nm spatial resolution - of the three different monoclinic domains and their melting during the resistive switching process.

Figure 2a) shows the PEEM image obtained in the monoclinic insulating phase of \chV_2O_3 at T𝑇Titalic_T = 120 K. The in-gap \chV_2O_3 features the nanotexture typical of the monoclinic insulating phase [36, 37]. Monoclinic domains with different ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT give rise to different XLD contrast, which can be appreciated as different color intensities within the PEEM image. The minimization of the total strain leads to the formation of stripe-like domains, whose directions are constrained by the symmetry of the system [37], as it will be discussed later. Each monoclinic insulating domain with a specific ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, extends over a few micrometers in length, thus connecting the two electrodes, and it is characterized by a lateral size of wdomsimilar-tosubscript𝑤𝑑𝑜𝑚absentw_{dom}\simitalic_w start_POSTSUBSCRIPT italic_d italic_o italic_m end_POSTSUBSCRIPT ∼200 nm [37].

Refer to caption
Figure 2: PEEM images before (a) and during (b-j) electric current application at T𝑇Titalic_T = 120 K. The homogeneous regions on top and bottom of each image are the gold electrodes; the area in between the electrodes represents the exposed \chV_2O_3 surface exhibiting the striped domains nanotexture typical of the antiferromagnetic monoclinic phase. For currents I𝐼Iitalic_I larger than 1.5 mA, the striped domains disappear in the region delimited by white dashed lines. This shows that a rhombohedral metallic filament is formed and widens as the current running through the device is increased.

The PEEM imaging is then repeated while driving a current I𝐼Iitalic_I through the device and measuring the voltage drop V𝑉Vitalic_V across the contacts. Figures 2b)-j) show the PEEM images acquired at increasing values of I𝐼Iitalic_I, following the upward branch of the hysteresis cycle. The presence of an in-plane electric field across the electrodes introduces a weak image blurring that becomes significant for electric potential differences V𝑉Vitalic_V larger than 6-8 V. Despite this effect, the nanodomains are clearly resolved during the resistive switching process, which manifests itself in the voltage drop observed between 0.08 mA and 1.1 mA (Fig. 2 c) and d) respectively). As the current is further increased beyond the threshold value necessary for the switching, we progressively observe the melting of the monoclinic nanotexture in the region delimited by white dashed lines (Fig. 2 e)-j)) and the appearance of a channel with homogenous intensity. The XLD contrast measured in this region corresponds to the signal of the high-temperature rhombohedral phase. This is also confirmed from the angle dependence of the XLD signal [37]. As shown in the Supplementary Information Fig. S3, images collected with two different angles of the X-rays polarization with respect to the in-plane \chV_2O_3 axes show no intensity variation upon sample rotation in the metallic channel, as opposed to the lateral monoclinic domains, whose signal depends on the angle between the light polarization and ϵ(𝐫)italic-ϵ𝐫\overrightarrow{\epsilon}(\mathbf{r})over→ start_ARG italic_ϵ end_ARG ( bold_r ). We can thus attribute the flat signal region in the middle of the gap to the formation of a metallic channel with rhombohedral lattice structure (ϵitalic-ϵ\epsilonitalic_ϵ=0). In the Supplementary Information Fig. S4, we report PEEM images obtained when repeating the experiment in the same conditions but with a larger field of view that allows to capture the whole gap of the device. The metallic channel always forms in the same location within the gap and no other metallic paths are observed. Furthermore, when the applied current is removed, the metallic channel disappears and the monoclinic domains form again with the same pre-switching configuration, indicating a volatile process.

Refer to caption
Figure 3: a) Detail of the sample nanotexture in the region where the metallic filament is formed upon application of an above-threshold current. b) Sketch of the arrangement of monoclinic domains crossing at 60superscript6060^{\circ}60 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT angle and forming a topological defect. Blue, red and yellow areas identify the three possible monoclinic domains corresponding to the three equivalent order parameter directions ϵ^nsubscript^italic-ϵ𝑛\hat{\epsilon}_{n}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. The order parameter at the boundaries between different domains is oriented along ϵ^1+ϵ^2subscript^italic-ϵ1subscript^italic-ϵ2\hat{\epsilon}_{1}+\hat{\epsilon}_{2}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (2π/32𝜋32\pi/32 italic_π / 3) for the red-blue interface and along ϵ^2+ϵ^3subscript^italic-ϵ2subscript^italic-ϵ3\hat{\epsilon}_{2}+\hat{\epsilon}_{3}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (4π/34𝜋34\pi/34 italic_π / 3) for the blue-yellow interface. The mixed red-yellow triangular region indicates the local suppression of the strain at the topological defect. The energy functional F𝐹Fitalic_F as function of the order parameter η𝜂\etaitalic_η, sketched in the left and right panels, shows how a smaller value of ϵ2superscriptitalic-ϵ2\epsilon^{2}italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT due to the topological defect (green plot, solid line) results in a decrease of the insulator-metal energy difference ΔΔ\Deltaroman_Δ.

The formation of the metallic channel is pinned by a specific topology of the lattice nanotexture, characterized by V-shaped domains, i.e., at the crossing point of domains with the same ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and directions that differ by a π/3𝜋3\pi/3italic_π / 3 angle. In Fig. 3a) we report a detail of the switching region using a colorscale that highlights the nature of the three different domains: 1) light blue, which corresponds to a monoclinic distortion along the ϵ^2subscript^italic-ϵ2\hat{\epsilon}_{2}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT direction (ϕ2subscriptitalic-ϕ2\phi_{2}italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT=π𝜋\piitalic_π); 2) red, which corresponds to a monoclinic distortion along the ϵ^1subscript^italic-ϵ1\hat{\epsilon}_{1}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT direction (ϕ1subscriptitalic-ϕ1\phi_{1}italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT=π𝜋\piitalic_π/3); 3) yellow, which corresponds to a monoclinic distortion along the ϵ^3subscript^italic-ϵ3\hat{\epsilon}_{3}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT direction (ϕ3subscriptitalic-ϕ3\phi_{3}italic_ϕ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT=5π𝜋\piitalic_π/3). As demonstrated in Ref. 37, the stabilization of the monoclinic nanotexture is driven by the Saint-Venant compatibility condition, which guarantees the continuity of the medium during a deformation and corresponds to the curl-free condition:

×ϵ(𝐫)=0.italic-ϵ𝐫0\overrightarrow{\nabla}\times\overrightarrow{\epsilon}(\mathbf{r})=0.over→ start_ARG ∇ end_ARG × over→ start_ARG italic_ϵ end_ARG ( bold_r ) = 0 . (3)

The conservation of the parallel component of the position dependent order parameter, ϵ(𝐫)italic-ϵ𝐫\overrightarrow{\epsilon}(\mathbf{r})over→ start_ARG italic_ϵ end_ARG ( bold_r ), across an interface between two different domains, as implied by the curl-free condition, has two major consequences:

  • i)

    the interface between two different monoclinic domains must be oriented along the direction corresponding to the order parameter of the third domain;

  • ii)

    the interface between a monoclinic and a rhombohedral metallic domain must be oriented perpendicularly to the order parameter of the monoclinic domain.

If we consider, for example, a domain with order parameter along ϵ^2subscript^italic-ϵ2\hat{\epsilon}_{2}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (light blue in Figure 3b)), its interface is oriented along ϵ^1subscript^italic-ϵ1\hat{\epsilon}_{1}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, i.e. at π𝜋\piitalic_π/3 angle, when it neighbours with a ϵ^3subscript^italic-ϵ3\hat{\epsilon}_{3}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT domain (yellow), whereas it is oriented along ϵ^3subscript^italic-ϵ3\hat{\epsilon}_{3}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, i.e. at 2π𝜋\piitalic_π/3 angle when it neighbours with a ϵ^1subscript^italic-ϵ1\hat{\epsilon}_{1}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT domain (red), in agreement with the nanotexture reported in Fig. 3. The Saint-Venant condition corresponds to a fixed phase jump δϕ𝛿italic-ϕ\delta\phiitalic_δ italic_ϕ=2π𝜋\piitalic_π/3 of ϵ(𝐫)italic-ϵ𝐫\overrightarrow{\epsilon}(\mathbf{r})over→ start_ARG italic_ϵ end_ARG ( bold_r ) across any interface between two monoclinic domains. We note that this condition is easily satisfied throughout the domains of the V22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT sample, except for the vertex of the V-shaped structures formed by the merging of two domains with ϵ^2subscript^italic-ϵ2\hat{\epsilon}_{2}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT order parameter direction (blue) and boundaries oriented along the ϵ^1subscript^italic-ϵ1\hat{\epsilon}_{1}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ϵ^3subscript^italic-ϵ3\hat{\epsilon}_{3}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT directions. If we consider a circuit Γ1subscriptΓ1\Gamma_{1}roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT across the boundary between two striped domains, the total phase shift is given by δϕ𝛿italic-ϕ\delta\phiitalic_δ italic_ϕ=+2π𝜋\piitalic_π/3-2π𝜋\piitalic_π/3=0 thus respecting the curl-free condition. In contrast, the topology of the V-shaped structure is such that, if we move around the internal apex (Γ2subscriptΓ2\Gamma_{2}roman_Γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT), the total phase-shift is constrained to δϕ𝛿italic-ϕ\delta\phiitalic_δ italic_ϕ=+2π𝜋\piitalic_π/3+2π𝜋\piitalic_π/3=4π𝜋\piitalic_π/3 thus breaking the curl-free condition. The direct consequence is that the vertex of the V shaped domains acts as a topological defect with fractional Hopf index (see Supplementary Information Section S6). These topological defects are inherently characterized by the strong frustration of the local value of the order parameter ϵ(𝐫)italic-ϵ𝐫\overrightarrow{\epsilon}(\mathbf{r})over→ start_ARG italic_ϵ end_ARG ( bold_r ) and local fluctuations on spatial and temporal scales that cannot be captured by the present experiment. We further note that the formation of this kind of topological defect is a direct and unavoidable consequence of the quasi-1D confined geometry of the system. Whereas the component of the order parameter parallel to the electrodes (ϵ||\epsilon_{||}italic_ϵ start_POSTSUBSCRIPT | | end_POSTSUBSCRIPT, see Fig. 3b) can be compensated outside the gap, the perpendicular component (ϵsubscriptitalic-ϵperpendicular-to\epsilon_{\perp}italic_ϵ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT) should be minimized to avoid excessive strain energy accumulation within the gap region. Thus, considering the directions of ϵ(𝐫)italic-ϵ𝐫\overrightarrow{\epsilon}(\mathbf{r})over→ start_ARG italic_ϵ end_ARG ( bold_r ) at the boundaries between different monoclinic domains (see Fig. 3b), the formation of V-shaped domains is the only configuration that fulfils the requirement ϵsubscriptitalic-ϵperpendicular-to\epsilon_{\perp}italic_ϵ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT=0.

Refer to caption
Figure 4: a) I-V curve measured along with the collection of the PEEM images. The drop in voltage measured at I = 1.05 mA signals the first resistive switching. b) Line profiles of the PEEM images in Fig. 2. The grey shaded area indicates the region where the rhombohedral metallic channel is formed. c) Width d𝑑ditalic_d of the metallic filament as a function of current I𝐼Iitalic_I in proximity of Ithsubscript𝐼𝑡I_{th}italic_I start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT. The blue and red markers represent the values of d𝑑ditalic_d obtained experimentally from PEEM images. The green solid line is the estimate of the width d𝑑ditalic_d according to the two parallel resistors model, which predicts a jump to a 200 µµ\textmuroman_µm wide filament at Ithsubscript𝐼𝑡I_{th}italic_I start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT (see Supplementary Information Section S5).

The suppression of the symmetry-breaking order parameter, ϵ(𝐫)italic-ϵ𝐫\overrightarrow{\epsilon}(\mathbf{r})over→ start_ARG italic_ϵ end_ARG ( bold_r ), at topological defects has far-reaching implications related to the nature of the resistive switching process. The electronic IMT can be described by a scalar order parameter η(𝐫)𝜂𝐫\eta(\mathbf{r})italic_η ( bold_r ) [37], which depends on the position 𝐫𝐫\mathbf{r}bold_r and is such that η=1𝜂1\eta=-1italic_η = - 1 in the metallic state and η=+1𝜂1\eta=+1italic_η = + 1 in the insulating state. The coupling between the electronic and structural transitions can be described by the energy functional [37]:

F[ϵ,η]𝑑𝐫{(η2(𝐫)1)2g(ϵ2(𝐫)ϵt2(V))η(𝐫)},proportional-to𝐹italic-ϵ𝜂differential-d𝐫superscriptsuperscript𝜂2𝐫12𝑔superscriptitalic-ϵ2𝐫subscriptsuperscriptitalic-ϵ2𝑡𝑉𝜂𝐫F\left[\epsilon,\eta\right]\propto\!\int\!d\mathbf{r}\,\Big{\{}\big{(}\eta^{2}% (\mathbf{r})-1\big{)}^{2}-g\big{(}\epsilon^{2}(\mathbf{r})-\epsilon^{2}_{t}(V)% \big{)}\,\eta(\mathbf{r})\Big{\}}\,,italic_F [ italic_ϵ , italic_η ] ∝ ∫ italic_d bold_r { ( italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) - 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_g ( italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) - italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_V ) ) italic_η ( bold_r ) } , (4)

where g𝑔gitalic_g is the coupling between the electronic order parameter and the strain and ϵt(V)subscriptitalic-ϵ𝑡𝑉\epsilon_{t}(V)italic_ϵ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_V ) is a threshold parameter that controls the first-order IMT and can depend on the applied voltage V𝑉Vitalic_V. When ϵ2(𝐫)>ϵt2(V)superscriptitalic-ϵ2𝐫subscriptsuperscriptitalic-ϵ2𝑡𝑉\epsilon^{2}(\mathbf{r})>\epsilon^{2}_{t}(V)italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) > italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_V ), the insulating phase with η=+1𝜂1\eta=+1italic_η = + 1 is locally favoured, whereas for strain smaller than the threshold value, i.e. ϵ2(𝐫)<ϵt2(V)superscriptitalic-ϵ2𝐫subscriptsuperscriptitalic-ϵ2𝑡𝑉\epsilon^{2}(\mathbf{r})<\epsilon^{2}_{t}(V)italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) < italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_V ), the metallic solution is stabilized. ϵt2(V)subscriptsuperscriptitalic-ϵ2𝑡𝑉\epsilon^{2}_{t}(V)italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_V ) thus represents the threshold above which the insulating monoclinic state (η=+1𝜂1\eta=+1italic_η = + 1, ϵ0italic-ϵ0\epsilon\neq 0italic_ϵ ≠ 0) becomes stable. The description of the electric-field induced transition is based on the observation [18] that the electric field directly couples to the electronic bandstructure of a Mott insulator and makes the metallic phase more stable. The transition can thus be described assuming that ϵt2(V)subscriptsuperscriptitalic-ϵ2𝑡𝑉\epsilon^{2}_{t}(V)italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_V ) increases as the voltage V𝑉Vitalic_V is increased. The energy difference between the insulating and metallic phase can be expressed as Δ(𝐫,V)=F[1]F[+1]g[ϵ2(𝐫)ϵt2(V)]Δ𝐫𝑉𝐹delimited-[]1𝐹delimited-[]1similar-to-or-equals𝑔delimited-[]superscriptitalic-ϵ2𝐫subscriptsuperscriptitalic-ϵ2𝑡𝑉\Delta(\mathbf{r},V)=F[-1]-F[+1]\simeq g\left[\epsilon^{2}(\mathbf{r})-% \epsilon^{2}_{t}(V)\right]roman_Δ ( bold_r , italic_V ) = italic_F [ - 1 ] - italic_F [ + 1 ] ≃ italic_g [ italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) - italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_V ) ]. If we start from the insulating phase with ϵ2(𝐫)>ϵt2(V=0)superscriptitalic-ϵ2𝐫subscriptsuperscriptitalic-ϵ2𝑡𝑉0\epsilon^{2}(\mathbf{r})>\epsilon^{2}_{t}(V=0)italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) > italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_V = 0 ), the IMT takes place when V𝑉Vitalic_V is increased up to the point that Δ(𝐫,V)Δ𝐫𝑉\Delta(\mathbf{r},V)roman_Δ ( bold_r , italic_V )=0. A topological defect, which locally suppresses ϵ2(𝐫)superscriptitalic-ϵ2𝐫\epsilon^{2}(\mathbf{r})italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ), thus naturally acts as a seed with lower threshold as compared to the rest of the system.

Intriguingly, we also note from Eq. 4 that the IMT can take place at a non-zero value of ϵ2(𝐫)superscriptitalic-ϵ2𝐫\epsilon^{2}(\mathbf{r})italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ), which makes possible the formation of a non-thermal metallic state (η=1𝜂1\eta=-1italic_η = - 1) with finite monoclinic distortion (ϵ2(𝐫)ϵt2(V)less-than-or-similar-tosuperscriptitalic-ϵ2𝐫subscriptsuperscriptitalic-ϵ2𝑡𝑉\epsilon^{2}(\mathbf{r})\lesssim\epsilon^{2}_{t}(V)italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) ≲ italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_V )), as already observed in non-equilibrium experiments [37, 44]. The nature of the early-stage switching process can be inferred by a direct comparison between the electrical state of the device and the melting of the monoclinic domains. The IV𝐼𝑉I-Vitalic_I - italic_V curve of the device, as measured in-situ during the PEEM experiment, is plotted in Fig. 4a). PEEM images are recorded at specific values of I𝐼Iitalic_I. The IV𝐼𝑉I-Vitalic_I - italic_V plot shows that the first resistive switching event occurs at the threshold current Ithsubscript𝐼𝑡I_{th}italic_I start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT = 1.05 mA. In Figure 4b) we report a linecut of the PEEM image acquired at specific values of I𝐼Iitalic_I; the image profile is taken along a line crossing the monoclinic domains in the middle of the device gap (see white solid line in Fig. 4, top panel). For large currents running through the device, the line profile in Fig. 4b) displays a flat region, which indicates the melting of the monoclinic nanodomains due to the formation of the rhombohedral metallic channel. As highlighted by the grey area in Fig. 4b), the width d𝑑ditalic_d of the metallic filament increases with the current, from d=0.23±0.05𝑑plus-or-minus0.230.05d=0.23\pm 0.05italic_d = 0.23 ± 0.05 µm at I𝐼Iitalic_I = 1.5 mA to d=3.7±0.2𝑑plus-or-minus3.70.2d=3.7\pm 0.2italic_d = 3.7 ± 0.2 µm at I𝐼Iitalic_I = 10 mA.

Modelling the device as a circuit with two parallel resistors (see Supplementary Information Section S5) allows us to estimate the expected width d𝑑ditalic_d of the rhombohedral filament corresponding to the observed voltage drop. For large currents running through the device, the experimental values of d𝑑ditalic_d match well with what is expected to be the case when the metallic channel forming in the gap has the resistivity of the high-temperature rhombohedral phase, as shown in Fig. 4. However, in correspondence of the first resistive switching event at Ithsubscript𝐼𝑡I_{th}italic_I start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT=1.05 mA, this model already predicts the sudden formation of a 200similar-toabsent200\sim 200∼ 200 nm wide metallic rhombohedral filament, which is not visible from PEEM measurements (see Fig. 4c and Supplementary Information Fig. S5), despite being well above the experimental resolution of the microscope. To explain this discrepancy, one might suspect that a rhombohedral metallic filament forms below the surface of the \chV_2O_3 film, where it is not detected by PEEM which is mainly a surface technique, sensitive to the first few nanometers. In fact, two arguments act against this possibility: i) the presence of the \chCr_2O_3 buffer layer allows to reduce the substrate-film lattice mismatch from 4.2%percent4.24.2\%4.2 % to 0.1%percent0.10.1\%0.1 %, thus removing almost entirely the residual epitaxial strain in the V22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT film [42], which is known to suppress the monoclinic phase and favour interfacial metallicity [42, 45]. In contrast to highly-strained films, in which the metal to insulator resistivity jump is strongly suppressed [45], our films display the 5-order of magnitude resistivity change typical of the unstrained metal-to-insulator transition (see Fig. S1); ii) the curl-free conditions force the interface between monoclinic and rhombohedral metallic regions to be oriented perpendicularly to the order parameter of the monoclinic domain. The formation of a below-surface metallic layer would lead to a sharp (much-less-than\ll 20 nm) monoclinic-rhombohedral interface parallel to ϵitalic-ϵ\overrightarrow{\epsilon}over→ start_ARG italic_ϵ end_ARG, thus leading to a dramatic increase of the strain energy of the system. Our results are compatible with a complex scenario in which the topology-driven resistive switching likely occurs first via the sudden transformation of a single 200 nm wide insulating monoclinic domain into a metallic channel with a non-thermal monoclinic lattice structure. At a second stage, the Joule heating leads to the thermally driven monoclinic-to-rhombohedral structural transition and the formation of rhombohedral metallic channels perpendicular to both the metallic electrodes and the ϵ^2subscript^italic-ϵ2\hat{\epsilon}_{2}over^ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT order parameter direction, as observed in Fig. 2.

The X-ray-based nanoimaging of a Mott device under operating conditions allowed us to simultaneously capture the formation of nanoscale conductive paths and the topology of the underlying symmetry-broken nanotexture. The present results expand our knowledge of the resistive switching process in Mott materials by demonstrating the leading role of inherent topological defects in the firing process. The combination of the methodologies here introduced with nanoscale strain engineering approaches unlocks the gate to new opportunities to manipulate topological defects and control the electronic switching dynamics in real devices, such as Mott-transition-based RRAM [46, 47], Mott memristor [48, 49, 50] and artificial neurons [51, 52]. The concept of topology-driven resistive switching will be key to assessing the possible non-thermal nature of the early stage electronic phase [37] as well as the microscopic origin of memory and non-volatile effects recently observed in Mott devices [6]. We finally note that the intimate relation between topological defects and electronic phase transitions is a general concept, potentially extendable to any system that undergoes a first-order phase transition accompanied by some form of symmetry breaking, as described by the energy functional (4). Relevant examples embrace transition-metal oxides [53, 3], such as vanadates, nickelates and manganites, and layered materials, such as 1T𝑇Titalic_T-TaS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT [54, 55, 56, 57], in which the insulator-to-metal transition is accompanied by charge-, lattice- and orbital-ordered states with reduced symmetry. Other interesting platforms are cuprate superconductors [58] and kagome metals [59] in which light- or magnetic-induced discontinuous electronic transitions coexist with charge-order. We thus provide a new framework for non-equilibrium electronic phase transitions and optical control of hidden states [60, 61, 57] in a broad class of quantum materials, in which the early-stage dynamics can be reversibly controlled by light or voltage, exploiting topological defects of the underlying symmetry-breaking order parameter [62, 57, 63].

We thank Diamond Lights Source for the provision of beamtime under proposal numbers MM-27218, MM-31711 and MM-34455. We thank Manuel R. Osorio and Fernando J. Urbanos for the fabrication of sample electrodes at the Centre for Micro and Nanofabrication of IMDEA Nanociencia. A.M., S.M. and C.G. acknowledge financial support from MIUR through the PRIN 2015 (Prot. 2015C5SEJJ001) and PRIN 2017 (Prot. 20172H2SC4_005) programs and from the European Uninion - Next Generation EU through the MUR-PRIN2022 (Prot. 20228YCYY7) program. Funded by the European Union - Next Generation EU C.G. acknowledges support from Università Cattolica del Sacro Cuore through D.1, D.2.2 and D.3.1 grants. S.M. acknowledges partial financial support through the grant ”Finanziamenti ponte per bandi esterni” from Università Cattolica del Sacro Cuore. I.F.C. and M.M. acknowledge support from the “Severo Ochoa” Programme for Centres of Excellence in R&D (CEX2020-001039-S) and the Spanish AEI-MCIN PID2021-122980OB-C52 (ECoSOx-ECLIPSE). I.F.C holds a FPI fellowship from the Spanish AEI-MCIN (PRE2020-092625). W.-F.H., S.M., J.W.S. and J.-P.L. acknowledge financial support by the KU Leuven Research Funds Project No. C14/21/083, iBOF/21/084, KAC24/18/056 and C14/17/080, as well as the FWO AKUL/13/19 and AKUL/19/023, and the Research Funds of the INTERREG-E-TEST Project (EMR113) and INTERREG-VL-VL-PATHFINDER Project (0559).

References

  • Wang et al. [2020] Z. Wang, H. Wu, G. W. Burr, C. S. Hwang, K. L. Wang, Q. Xia, and J. J. Yang, Resistive switching materials for information processing, Nature Reviews Materials 5, 173 (2020).
  • Zhou and Ramanathan [2015] Y. Zhou and S. Ramanathan, Mott Memory and Neuromorphic Devices, Proceedings of the IEEE 103, 1289 (2015).
  • Tokura et al. [2017] Y. Tokura, M. Kawasaki, and N. Nagaosa, Emergent functions of quantum materials, Nature Physics 13, 1056 (2017).
  • del Valle et al. [2018] J. del Valle, J. G. Ramirez, M. J. Rozenberg, and I. K. Schuller, Challenges in materials and devices for resistive-switching-based neuromorphic computing, Journal of Applied Physics 124, 211101 (2018).
  • Adda et al. [2018] C. Adda, B. Corraze, P. Stoliar, P. Diener, J. Tranchant, A. Filatre-Furcate, M. Fourmigué, D. Lorcy, M.-P. Besland, E. Janod, and L. Cario, Mott insulators: A large class of materials for Leaky Integrate and Fire (LIF) artificial neuron, Journal of Applied Physics 124, 152124 (2018).
  • del Valle et al. [2019] J. del Valle, J. G. Ramirez, M. J. Rozenberg, and I. K. Schuller, Subthreshold firing in Mott nanodevices, Nature 569, 388 (2019).
  • Pérez-Tomás [2019] A. Pérez-Tomás, Functional Oxides for Photoneuromorphic Engineering: Toward a Solar Brain, Advanced Materials Interfaces 6, 1900471 (2019).
  • del Valle et al. [2020] J. del Valle, P. Salev, Y. Kalcheim, and I. K. Schuller, A caloritronics-based Mott neuristor, Scientific Reports 10, 4292 (2020).
  • Zhang et al. [2021] Z. Zhang, S. Mondal, S. Mandal, J. M. Allred, N. A. Aghamiri, A. Fali, Z. Zhang, H. Zhou, H. Cao, F. Rodolakis, J. L. McChesney, Q. Wang, Y. Sun, Y. Abate, K. Roy, K. M. Rabe, and S. Ramanathan, Neuromorphic learning with Mott insulator NiO, Proceedings of the National Academy of Sciences 118, e2017239118 (2021).
  • Deng et al. [2021] X. Deng, S.-Q. Wang, Y.-X. Liu, N. Zhong, Y.-H. He, H. Peng, P.-H. Xiang, and C.-G. Duan, A Flexible Mott Synaptic Transistor for Nociceptor Simulation and Neuromorphic Computing, Advanced Functional Materials 31, 2101099 (2021).
  • Deng et al. [2023] S. Deng, H. Yu, T. J. Park, A. N. Islam, S. Manna, A. Pofelski, Q. Wang, Y. Zhu, S. K. Sankaranarayanan, A. Sengupta, and S. Ramanathan, Selective area do** for Mott neuromorphic electronics, Science Advances 9, eade4838 (2023).
  • Ran et al. [2023] Y. Ran, Y. Pei, Z. Zhou, H. Wang, Y. Sun, Z. Wang, M. Hao, J. Zhao, J. Chen, and X. Yan, A review of Mott insulator in memristors: The materials, characteristics, applications for future computing systems and neuromorphic computing, Nano Research 16, 1165 (2023).
  • Yang et al. [2011] Z. Yang, C. Ko, and S. Ramanathan, Oxide electronics utilizing ultrafast metal-insulator transitions, Annual Review of Materials Research 41, 337 (2011).
  • Mehonic and Kenyon [2022] A. Mehonic and A. J. Kenyon, Brain-inspired computing needs a master plan, Nature 604, 255 (2022).
  • Hsu et al. [2023] W.-F. Hsu, S. Mellaerts, C. Bellani, P. Homm, N. Uchida, M. Menghini, M. Houssa, J. W. Seo, and J.-P. Locquet, Raman spectroscopy and phonon dynamics in strained \chV2O3, Physical Review Materials 7, 074606 (2023).
  • Zhang and Averitt [2014] J. Zhang and R. D. Averitt, Dynamics and control in complex transition metal oxides, Annual Review of Materials Research 44, 19 (2014).
  • Basov et al. [2017] D. Basov, R. Averitt, and D. Hsieh, Towards properties on demand in quantum materials, Nature materials 16, 1077 (2017).
  • Mazza et al. [2016] G. Mazza, A. Amaricci, M. Capone, and M. Fabrizio, Field-driven mott gap collapse and resistive switch in correlated insulators, Physical review letters 117, 176401 (2016).
  • Suen et al. [2023] C. T. Suen, I. Marković, M. Zonno, S. Zhdanovich, N.-H. Jo, M. Schmid, P. Hansmann, P. Pupuhal, K. Fürsich, V. Zimmerman, et al., Nature of the current-induced insulator-to-metal transition in \chCa_2RuO_4 as revealed by transport-ARPES, arXiv:2308.05803  (2023).
  • Wouters et al. [2019] D. J. Wouters, S. Menzel, J. A. J. Rupp, T. Hennen, and R. Waser, On the universality of the I-V switching characteristics in non-volatile and volatile resistive switching oxides, Faraday Discuss. 213, 183 (2019).
  • Kalcheim et al. [2020] Y. Kalcheim, A. Camjayi, J. del Valle, P. Salev, M. Rozenberg, and I. K. Schuller, Non-thermal resistive switching in Mott insulator nanowires, Nature Communications 11, 2985 (2020).
  • Stoliar et al. [2013] P. Stoliar, L. Cario, E. Janod, B. Corraze, C. Guillot-Deudon, S. Salmon-Bourmand, V. Guiot, J. Tranchant, and M. Rozenberg, Universal Electric-Field-Driven Resistive Transition in Narrow-Gap Mott Insulators, Advanced Materials 25, 3222 (2013).
  • Guiot et al. [2013] V. Guiot, L. Cario, E. Janod, B. Corraze, V. Ta Phuoc, M. Rozenberg, P. Stoliar, T. Cren, and D. Roditchev, Avalanche breakdown in \chGaTa_4Se_8-xTe_x narrow-gap Mott insulators, Nature communications 4, 1722 (2013).
  • Nakamura et al. [2013] F. Nakamura, M. Sakaki, Y. Yamanaka, S. Tamaru, T. Suzuki, and Y. Maeno, Electric-field-induced metal maintained by current of the Mott insulator \chCa_2RuO_4, Scientific reports 3, 2536 (2013).
  • Fursina et al. [2009] A. Fursina, R. Sofin, I. Shvets, and D. Natelson, Origin of hysteresis in resistive switching in magnetite is Joule heating, Physical Review B 79, 245131 (2009).
  • Fursina et al. [2012] A. Fursina, R. Sofin, I. Shvets, and D. Natelson, Statistical distribution of the electric field-driven switching of the Verwey state in \chFe_3O_4, New Journal of Physics 14, 013019 (2012).
  • Janod et al. [2015] E. Janod, J. Tranchant, B. Corraze, M. Querré, P. Stoliar, M. Rozenberg, T. Cren, D. Roditchev, V. T. Phuoc, M.-P. Besland, et al., Resistive switching in Mott insulators and correlated systems, Advanced Functional Materials 25, 6287 (2015).
  • Waser and Aono [2007] R. Waser and M. Aono, Nanoionics-based resistive switching memories, Nature materials 6, 833 (2007).
  • Ielmini [2016] D. Ielmini, Resistive switching memories based on metal oxides: mechanisms, reliability and scaling, Semiconductor Science and Technology 31, 063002 (2016).
  • Lee et al. [2015] J. S. Lee, S. Lee, and T. W. Noh, Resistive switching phenomena: A review of statistical physics approaches, Applied Physics Reviews 2, 031303 (2015).
  • Stoliar et al. [2017] P. Stoliar, J. Tranchant, B. Corraze, E. Janod, M.-P. Besland, F. Tesler, M. Rozenberg, and L. Cario, A Leaky-Integrate-and-Fire Neuron Analog Realized with a Mott Insulator, Advanced Functional Materials 27, 1604740 (2017).
  • Lange et al. [2021] M. Lange, S. Guénon, Y. Kalcheim, T. Luibrand, N. M. Vargas, D. Schwebius, R. Kleiner, I. K. Schuller, and D. Koelle, Imaging of electrothermal filament formation in a mott insulator, Physical Review Applied 16, 054027 (2021).
  • Del Valle et al. [2021] J. Del Valle, N. M. Vargas, R. Rocco, P. Salev, Y. Kalcheim, P. N. Lapa, C. Adda, M.-H. Lee, P. Y. Wang, L. Fratino, et al., Spatiotemporal characterization of the field-induced insulator-to-metal transition, Science 373, 907 (2021).
  • Babich et al. [2021] D. Babich, J. Tranchant, C. Adda, B. Corraze, M.-P. Besland, P. Warnicke, D. Bedau, B. Bertoncini, J.-Y. Mevellec, B. Humbert, J. Rupp, T. Hennen, D. Wouters, R. Llopis, L. Cario, and E. Janod, Lattice contraction induced by resistive switching in chromium-doped V22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT: a hallmark of Mott physics, arXiv preprint arXiv:2105.05093  (2021).
  • Luibrand et al. [2023] T. Luibrand, A. Bercher, R. Rocco, Tahouni-Bonab, L. Varbaro, C. Rischau, and e. a. Domínguez, Characteristic lengthscales of the electrically-induced insulator-to-metal transition, arXiv preprint arXiv:2301.00456  (2023).
  • Ronchi et al. [2019] A. Ronchi, P. Homm, M. Menghini, P. Franceschini, F. Maccherozzi, F. Banfi, G. Ferrini, F. Cilento, F. Parmigiani, S. S. Dhesi, et al., Early-stage dynamics of metallic droplets embedded in the nanotextured Mott insulating phase of V22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT, Physical Review B 100, 075111 (2019).
  • Ronchi et al. [2022] A. Ronchi, P. Franceschini, A. De Poli, P. Homm, A. Fitzpatrick, F. Maccherozzi, G. Ferrini, F. Banfi, S. S. Dhesi, M. Menghini, et al., Nanoscale self-organization and metastable non-thermal metallicity in Mott insulators, Nature Communications 13, 3730 (2022).
  • McWhan et al. [1969] D. McWhan, T. Rice, and J. Remeika, Mott transition in \chCr-doped \chV_2O_3, Physical Review Letters 23, 1384 (1969).
  • McWhan and Remeika [1970] D. McWhan and J. Remeika, Metal-insulator transition in \ch(V_1- xCr_x)_2O_3, Physical Review B 2, 3734 (1970).
  • McWhan et al. [1973] D. McWhan, A. Menth, J. Remeika, W. Brinkman, and T. Rice, Metal-insulator transitions in pure and doped \chV_2O_3, Physical Review B 7, 1920 (1973).
  • Guénon et al. [2013] S. Guénon, S. Scharinger, S. Wang, J. Ramírez, D. Koelle, R. Kleiner, and I. K. Schuller, Electrical breakdown in a \chV_2O_3 device at the insulator-to-metal transition, Europhysics Letters 101, 57003 (2013).
  • Dillemans et al. [2014] L. Dillemans, T. Smets, R. R. Lieten, M. Menghini, C.-Y. Su, and J.-P. Locquet, Evidence of the metal-insulator transition in ultrathin unstrained \chV2O3 thin films, Applied Physics Letters 104, 071902 (2014).
  • Park et al. [2000] J.-H. Park, L. Tjeng, A. Tanaka, J. Allen, C. Chen, P. Metcalf, J. Honig, F. De Groot, and G. Sawatzky, Spin and orbital occupation and phase transitions in \chV_2O_3, Physical Review B 61, 11506 (2000).
  • Ronchi et al. [2021] A. Ronchi, P. Franceschini, P. Homm, M. Gandolfi, G. Ferrini, S. Pagliara, F. Banfi, M. Menghini, C. Giannetti, et al., Light-assisted resistance collapse in a \chV_2O_3-based mott-insulator device, Physical Review Applied 15, 044023 (2021).
  • Pofelski et al. [2023] A. Pofelski, S. Valencia, Y. Kalcheim, P. Salev, A. Rivera, C. Huang, M. A. Mawass, F. Kronast, I. K. Schuller, Y. Zhu, et al., Domain nucleation across the metal-insulator transition of self-strained V22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO33{}_{3}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT films, arXiv preprint arXiv:2312.09051  (2023).
  • Sawa [2008] A. Sawa, Resistive switching in transition metal oxides, Materials Today 11, 28 (2008).
  • Wang et al. [2019] Y. Wang, K.-M. Kang, M. Kim, H.-S. Lee, R. Waser, D. Wouters, R. Dittmann, J. J. Yang, and H.-H. Park, Mott-transition-based RRAM, Materials Today 28, 63 (2019).
  • Pickett et al. [2013] M. D. Pickett, G. Medeiros-Ribeiro, and R. S. Williams, A scalable neuristor built with Mott memristors, Nature Materials 12, 114 (2013).
  • Yoshida et al. [2015] M. Yoshida, R. Suzuki, Y. Zhang, M. Nakano, and Y. Iwasa, Memristive phase switching in two-dimensional 1\chT-TaS2 crystals, Science Advances 1, e1500606 (2015).
  • Kumar et al. [2017] S. Kumar, J. P. Strachan, and R. S. Williams, Chaotic dynamics in nanoscale \chNbO2 Mott memristors for analogue computing, Nature 548, 318 (2017).
  • Tesler et al. [2018] F. Tesler, C. Adda, J. Tranchant, B. Corraze, E. Janod, L. Cario, P. Stoliar, and M. Rozenberg, Relaxation of a spiking Mott artificial neuron, Phys. Rev. Appl. 10, 054001 (2018).
  • Zhang et al. [2020] X. Zhang, Y. Zhuo, Q. Luo, Z. Wu, R. Midya, Z. Wang, W. Song, R. Wang, N. K. Upadhyay, Y. Fang, F. Kiani, M. Rao, Y. Yang, Q. Xia, Q. Liu, M. Liu, and J. J. Yang, An artificial spiking afferent nerve based on Mott memristors for neurorobotics, Nature Communications 11, 51 (2020).
  • Imada et al. [1998] M. Imada, A. Fujimori, and Y. Tokura, Metal-insulator transitions, Rev. Mod. Phys. 70, 1039 (1998).
  • Vaskivskyi et al. [2015] I. Vaskivskyi, J. Gospodaric, S. Brazovskii, D. Svetin, P. Sutar, E. Goreshnik, I. A. Mihailovic, T. Mertelj, and D. Mihailovic, Controlling the metal-to-insulator relaxation of the metastable hidden quantum state in 1T-TaS22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTScience Advances 1, e1500168 (2015).
  • Hollander et al. [2015] M. J. Hollander, Y. Liu, W.-J. Lu, L.-J. Li, Y.-P. Sun, J. A. Robinson, and S. Datta, Electrically driven reversible insulator-metal phase transition in 1\chT-TaS2, Nano Letters 15, 1861 (2015).
  • Lee et al. [2019] S.-H. Lee, J. S. Goh, and D. Cho, Origin of the insulating phase and first-order metal-insulator transition in 1\chT-TaS_2, Phys. Rev. Lett. 122, 106404 (2019).
  • Gao et al. [2022] F. Y. Gao, Z. Zhang, Z. Sun, L. Ye, Y.-H. Cheng, Z.-J. Liu, J. G. Checkelsky, E. Baldini, and K. A. Nelson, Snapshots of a light-induced metastable hidden phase driven by the collapse of charge order, Science Advances 8, eabp9076 (2022).
  • Keimer et al. [2015] B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida, and J. Zaanen, From quantum matter to high-temperature superconductivity in copper oxides, Nature 518, 179 (2015).
  • Asaba et al. [2024] T. Asaba, A. Onishi, Y. Kageyama, T. Kiyosue, K. Ohtsuka, S. Suetsugu, Y. Kohsaka, T. Gaggl, Y. Kasahara, H. Murayama, K. Hashimoto, R. Tazai, H. Kontani, B. R. Ortiz, S. D. Wilson, Q. Li, H. H. Wen, T. Shibauchi, and Y. Matsuda, Evidence for an odd-parity nematic phase above the charge-density-wave transition in a kagome metal, Nature Physics 10.1038/s41567-023-02272-4 (2024).
  • Stojchevska et al. [2014] L. Stojchevska, I. Vaskivskyi, T. Mertelj, P. Kusar, D. Svetin, S. Brazovskii, and D. Mihailovic, Ultrafast switching to a stable hidden quantum state in an electronic crystal, Science 344, 177 (2014).
  • Zhang et al. [2016] J. Zhang, X. Tan, M. Liu, S. W. Teitelbaum, K. W. Post, F. **, K. Nelson, D. N. Basov, W. Wu, and R. D. Averitt, Cooperative photoinduced metastable phase control in strained manganite films, Nature Materials 15, 956 (2016).
  • Wandel et al. [2022] S. Wandel, F. Boschini, E. H. da Silva Neto, L. Shen, M. X. Na, S. Zohar, Y. Wang, S. B. Welch, M. H. Seaberg, J. D. Koralek, G. L. Dakovski, W. Hettel, M.-F. Lin, S. P. Moeller, W. F. Schlotter, A. H. Reid, M. P. Minitti, T. Boyle, F. He, R. Sutarto, R. Liang, D. Bonn, W. Hardy, R. A. Kaindl, D. G. Hawthorn, J.-S. Lee, A. F. Kemper, A. Damascelli, C. Giannetti, J. J. Turner, and G. Coslovich, Enhanced charge density wave coherence in a light-quenched, high-temperature superconductor, Science 376, 860 (2022).
  • Cheng et al. [2024] Y. Cheng, A. Zong, L. Wu, Q. Meng, W. Xia, F. Qi, P. Zhu, X. Zou, T. Jiang, Y. Guo, J. van Wezel, A. Kogar, M. W. Zuerch, J. Zhang, Y. Zhu, and D. Xiang, Ultrafast formation of topological defects in a two-dimensional charge density wave, Nature Physics 20, 54 (2024).