Next Article in Journal
Water Adsorption Dynamics on Metal–Organic Framework MOF-801: Comparative Study of Loose and Glued Grains, and Coatings
Previous Article in Journal
Hydrogel Contact Lenses Embedded with Amine-Functionalized Large-Pore Mesoporous Silica Nanoparticles with Extended Hyaluronic Acid Release
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Cerium Oxide-Based Memristors for Neuromorphic Computing

1
Department of Physics, Riphah International University, Lahore Campus, 13-KM Raiwand Road, Lahore 54000, Pakistan
2
Department of Physics, University of Okara, Okara 56300, Pakistan
3
Department of Physics “Ettore Pancini”, University of Naples ‘Federico II’, Piazzale Tecchio, 80, 80125 Naples, Italy
4
Department of Electrical Engineering, Sejong University, Seoul 05006, Republic of Korea
5
Department of Advanced Battery Convergence Engineering, Dongguk University, Seoul 04620, Republic of Korea
6
Division of Electronics and Electrical Engineering, Dongguk University, Seoul 04620, Republic of Korea
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(17), 2443; https://doi.org/10.3390/nano13172443
Submission received: 21 July 2023 / Revised: 23 August 2023 / Accepted: 27 August 2023 / Published: 28 August 2023
(This article belongs to the Topic Energy Storage Materials and Devices)

Abstract

:
This review article attempts to provide a comprehensive review of the recent progress in cerium oxide (CeO2)-based resistive random-access memories (RRAMs). CeO2 is considered the most promising candidate because of its multiple oxidation states (Ce3+ and Ce4+), remarkable resistive-switching (RS) uniformity in DC mode, gradual resistance transition, cycling endurance, long data-retention period, and utilization of the RS mechanism as a dielectric layer, thereby exhibiting potential for neuromorphic computing. In this context, a detailed study of the filamentary mechanisms and their types is required. Accordingly, extensive studies on unipolar, bipolar, and threshold memristive behaviors are reviewed in this work. Furthermore, electrode-based (both symmetric and asymmetric) engineering is focused for the memristor’s structures such as single-layer, bilayer (as an oxygen barrier layer), and doped switching-layer-based memristors have been proved to be unique CeO2-based synaptic devices. Hence, neuromorphic applications comprising spike-based learning processes, potentiation and depression characteristics, potentiation motion and synaptic weight decay process, short-term plasticity, and long-term plasticity are intensively studied. More recently, because learning based on Pavlov’s dog experiment has been adopted as an advanced synoptic study, it is one of the primary topics of this review. Finally, CeO2-based memristors are considered promising compared to previously reported memristors for advanced synaptic study in the future, particularly by utilizing high-dielectric-constant oxide memristors.

1. Introduction

Because of the variety of oxide materials, the resistive-switching (RS) mechanism is widely investigated in exploring resistance variation processes in memristors. However, improving RS mechanisms attributed to resistive random-access memories (RRAMs) is still elusive. Nevertheless, no crucial switching mechanism has been guaranteed to be ideal, i.e., 100% RS behavior, to date. Therefore, an extensive review is indispensable for elucidating the switching kinetics of the desired switching materials to enable the incorporation of synaptic devices into these studies, which is presented in this work [1,2,3]. More recently, memristive devices have displayed new opportunities in neuromorphic computing research [4,5,6]. RS devices’ mechanism has been significantly explored in various nanomaterials such as TiO2, Cr2O3, and FeOx [7,8,9], and in high-k dielectrics with large bandgaps such as Al2O3 and HfO2 [10]. As an illustration of pursuing a comprehensive understanding of neuromorphic dynamics, an intricately developed computational model was introduced, specifically focusing on Ag-PVP nanowire networks (NWNs). In this context, Ag/PVP/Ag junctions were established at the interfacial contacts between the nanowires. Notably, all electrical junctions were meticulously modeled as voltage-controlled memristors. Within this framework, investigation into the switching mechanisms encompassed various aspects, outlined under headings such as “Collective Junction Switching Drives Non-local Transport”, “Avalanche Switching Dynamics”, and “Order-Chaos Transition from Polarity-Driven Switching” [11]. However, they lacked an approach to achieve the desired RS behavior in memory technology. For example, electrode engineering, unipolar, bipolar (symmetric and asymmetric), and threshold behaviors are still questionable. Moreover, they do not offer remarkable outcomes in their filamentary mechanism in insulating layers containing metal/insulator/metal and metal/insulator/insulator/metal configurations, as reported in the literature. Meanwhile, among oxide-based materials, cerium oxide (CeO2-x)-based nanomaterials possess potential applications such as catalysts, solar cells, and gate nanomaterials belonging to metal oxide-based semiconductor devices [12]. Nevertheless, they still show limitations in achieving the targeted outcomes. There was also a focus on single-crystalline materials owing to the allowed disentangling defect behavior occurring at grain boundaries, which are considered the intrinsic property of bulk materials [13].
To date, cerium oxide (CeO2) has been suggested as a promising candidate showing analog-switching behavior of memristors. Furthermore, several researchers have reported significant digital-type resistance variation for RRAM implementation using a ceria switching layer [14,15,16,17]. Ismail et al. [16] demonstrated the bipolar switching mechanism in Al/CeO2/Au and Zr/CeOx/Pt architectures. Lin et al. also presented the nonpolar switching behavior due to the polycrystalline CeO2 layer of a Pt/CeO2/Pt structure [17]. Remarkably, CeO2-based analog switching is still being researched. In this respect, the demonstration of analog-based switching of a Pt/CeO2/Pt memristor has been displayed in memristive technology. Moreover, Pt/CeO2/Pt analog synaptic devices also possess a symmetric electrode structure, thereby prominently exhibiting emulated synaptic trends in the form of potentiation and depression behaviors in computation. Therefore, its comprehensive switching peculiarities for artificial synapses have been extensively studied [18]. In previous studies, an identical synaptic weight modulation has been prominently considered owing to the functioning of the neural setup for generating voltage pulses containing symmetric amplitude as well as bipolar functioning toward potentiation and depression trends. Again, analog resistance variations exhibit long-term stability (LTS) to ensure constancy in signal formation, simulated learning, and retention processes. The abovementioned properties have been achieved under low-power/energy consumption conditions, as they always remain the primary goal of research areas [19].
In addition, a Pt/ITO/CeO2/Pt memristive device and ITO behaving as a capping layer often play a vital role toward synaptic trends. The stability of analog resistance variation to emulate the potentiation and depression functions in an ITO free-layered Pt/CeO2/Pt memristor has been elaborately reported [18,20]. Further, the exhibited potentiation and depression dynamics of the Pt/CeO2/Pt memristor have shown nonlinearity and asymmetry instead of a symmetric architecture. Because it has a nonfilamentary route such as that of other memristors, this device also has the disadvantage of memory loss, which limits its stability in analog/multilevel synaptic weight modulation. However, a CeO2-based memristor having oxygen reactivity in an ITO electrode form exhibits enhanced conductivity caused by increased density of oxygen vacancies (OVs) [21]. Hence, oxygen ions (O2−) are exchanged where CeO2-ITO interfacial contact occurs in the Pt/ITO/CeO2/Pt device, leading to stability of resistance modulation. Therefore, improved linear behavior, symmetric potentiation and depression, and long-life stability are the fundamental steps in promoting the Pt/ITO/CeO2/Pt memristor for artificial synapse devices in neuromorphic systems [19].
Further, memristors are suggested as candidates for artificial synapses owing to comparable structures with presynaptic-neuron/synapse/postsynaptic-neuron architectures with higher density of neuromorphic synaptic systems. Furthermore, their resistance states always mimic analog synaptic weight functioning with a wide variety of synaptic dynamics, such as postsynaptic current, paired-pulse facilitation (PPF), short-term plasticity (STP), long-term plasticity (LTP), and spike-time-dependent plasticity (STDP) [22,23,24]. Herein, a promising analog synapse formation in the Pt/CeO2/Pt memristor under annealing of ceria layer has also been investigated to explore optimal effect of Pt/CeO2 interfacial states. In that study, the primary resistance change was attributed to the tuning of interfacial states to mimic analog synaptic weight modulation in biological synapses without involving filament formation [24]. Additionally, the CeO2 layer complemented the Si-based electronics in the synapse layer, showing a reasonable permittivity of ~26 with a bandgap of ~3.2 eV [25] and with fruitful outcomes for the integrated artificial synapses decorated over a Si-based neural hardware foundation.
For the advancement in analog resistance variation in Pt/CeO2/Pt memristors, an approach involving altering the post-deposition annealing environment for changing the switching parameters of the CeO2 synapse-based layer has been widely studied. However, from the conduction mode investigation data and spectroscopic study of chemical-bonding states, the reference device was found to be capable of voltage-controlled operation due to Pt/CeO2 interfacial contacts [26]. Furthermore, researchers have worked on many artificial intelligence-based devices showing functions identical to those of neurons to function in several tasks such as visual recognition, audio recognition, and artificial nociceptor functioning [27,28,29].
Memristors have been extensively discussed as capable artificial synaptic devices owing to inheriting synaptic weight change into neural networks [13,18]. Moreover, synaptic plasticity can also be modulated by various input signals as synaptic functions. In this regard, various oxide-based nanomaterials, such as ZnO [30], Y2O3 [31], CeO2 [32,33,34], HfOx [35], and ZrO2 [36], have been proposed for memristive memories as well as synaptic type devices. Among them, CeO2 has been suggested as the most promising material for renewable nonvolatile memories, as it exhibits a high dielectric constant of ~26 and a high bandgap of 6 eV, which covers the maximum residual bandgap in other oxide materials [37]. The flexible valence states of Ce3+ and Ce4+ in ceria can create conducting filaments (CFs) containing OVs or may disrupt them upon application of a biasing voltage signal. Meanwhile, the versatile asymmetric electrode memristor Ag/CeO2/Pt has been widely examined. The resistance state-functioning of an active metal/oxide/inert metal structure was initially tuned by rupturing of the silver CF; moreover, the CF containing defective OVs was also comparatively studied [38]. The Ag/CeO2/Pt memristor has been demonstrated to be an inherent bipolar resistance switching device with the realization of slow variation in resistance states with low operating voltage, which has enabled artificial synapse simulation [39].
Herein, the main focus areas of research are schematically presented and elaborated. The various types of CeO2-based memristors are classified according to filamentary-based mechanisms. Then, the electrode-based (symmetric and asymmetric) structures and their performance are examined. Further, the resistance variation tendency of unipolar, bipolar, and threshold memristors and their role as single- and double-insulating layers (with capping layer) for achieving optimal RS behavior is elaborated. Neuromorphic applications involving spike-based learning, potentiation and depression behaviors, synaptic plasticity, and synaptic weight decay with time modulation are significantly investigated. The extensive review concludes that various CeO2-based memristors have demonstrated excellent performance while proving themselves as advanced synaptic devices. Mainly, the focus is on Pavlov’s dog experiment in this study because it represents an advanced synaptic study; accordingly, it has been critically examined here to cover recent advances in CeO2-based synaptic devices. Finally, future perspectives for advancement in oxide-based materials toward synaptic trends are suggested and are achievable through the utilization of alternative high-k oxide materials.

2. Filamentary Mechanisms in CeO2-Based Memristors

The RS mechanisms of various materials have been widely analyzed for elucidating the resistance variation process; however, the resistance switching mechanism in RRAMs is still contentious. Several literature reviews have been presented to understand switching kinetics in different switching materials. Interestingly, no specific switching mechanism assures 100% RS behavior [1,40]. However, the conductive filamentary mechanisms are broadly predictable, whereas major disagreements still prevail concerning microstates, configuration, and shape of the creation and annihilation of the CFs. Initially, one or multiple filaments, i.e., local conducting channels are formed or ruptured when an external stimulus is applied. Some materials may require a forming state with the application of external voltages. In this way, CFs may be created by the occurrence of soft break-up in the insulating layer [41]. Therein, change from a high-resistance state (HRS) to a low-resistance state (LRS) corresponds to the SET process, whereas the converse variation represents the RESET process. Further, under application of a positive voltage at the top electrode (TE), charge carriers trapped by vacancies form a filamentary path, and the device goes into LRS. Conversely, during the reverse mechanism the device will occupy HRS. Akinaga H. stated that the thermally assisted redox reaction with anodization at the metal electrode and oxide-layer interface is responsible for the possible mechanisms [42]. Usually, Joule heating is considered the leading cause of rupturing of conducting paths [43]. However, the typical filamentary mechanism is generally not dependent upon the contact area associated with device electrodes [44,45].

2.1. Electrochemical Metallization Filamentary Mechanism

The basic working principle of RRAM cells is the basis for filamentary classification. The filamentary mechanism can be classified as electrochemical metallization mechanism (ECM), valence change mechanism (VCM) [46], and thermochemical mechanism (TCM) [47,48].

2.1.1. Electrochemical Metallization Mechanism

ECM is often considered to elaborate the working principle of RRAM devices by using metals having high mobility for dynamic electrodes, such as silver (Ag), copper (Cu), and nickel (Ni) [3]. Yang et al. reported the conductive filament mechanism for the Pt/SiO2/Ag structure using transmission electron microscopy (TEM) as shown in Figure 1a,b [49]. For new planar architecture formation, conductive channels are not formed within the SiO2 layer lying between Ag and Pt electrodes. After application of a positive potential to the Ag electrode, an oxidation reaction occurs (Ag - e → Ag+). During the movement of Ag+ toward the Pt electrode surface, Ag+ loses electrons to undergo reduction and becomes Ag. The Ag provides evidence of Ag conductive filament formation in an RRAM switching layer, thereby exhibiting LRS as illustrated by Figure 1a. By contrast, upon application of a positive voltage at the Pt electrode, the HRS state of the device is observed as shown in Figure 1b, supporting the breakdown of the conducting path.
Yan et al. examined the RS behavior of Pt/SrTiO3/Ag RRAM devices using X-ray photoelectron spectroscopy (XPS), investigating the unintended conducting filamentary process related to Ag in memory devices [50]. The prepared device was operated in HRS and LRS states, and complex XPS spectra of Ag and Pt electrodes were obtained. The Ag component always provides evidence of their occurrence between the TE and bottom electrode (BE). Additionally, a slight variation in dispersion intensity may easily be identified at various depths for the SrTiO3 film when the device is at LRS. However, in HRS, a slightly insignificant amount of Ag is observed in the SrTiO3 film and the Pt thin film. Thus, the role of RS should be implicated in formation and rupture in the established silver filament mechanism. Meanwhile, Tsuruoka et al. have reported the quantized conductance of the Ag/Ta2O5/Pt structure as presented in Figure 1c–e [51].
Figure 1. (a) TEM result for the cell containing a larger electrode showing a space after the forming process (scale bar: 200 nm). The red color square and zoomed-in image of the filament (indicated with the upper arrows) close to the dielectric/inert electrode edge (scale bar: 20 nm) and (b) The red color square and zoomed-in image and TEM image after erasing. Adopted from [52]. Copyright 2012 Nature Publishing Group. Conductance quantization with the synaptic trend of Ta2O5-based atomic switching: (c) observed increment in quantized conductance owing to the cell upon application of positive voltage pulses of width 20 ms during a 2-s interval. The inset shows the controlled-conductance level realized by inserting a series combination with the cell along with a current-controlling resistor with a resistance of 3 kΩ, (d) observed decrement in quantized conductance owing to the cell upon application of negative voltage pulses of width 20 ms during a 2-s interval. The input pulses are directly connected to the device, shown in the inset view, and (e) The quantized conductance histogram demonstration from measured data with zoomed-in image of conductance against time. Adopted from [53]. Copyright 2012 IOP Publishing.
Figure 1. (a) TEM result for the cell containing a larger electrode showing a space after the forming process (scale bar: 200 nm). The red color square and zoomed-in image of the filament (indicated with the upper arrows) close to the dielectric/inert electrode edge (scale bar: 20 nm) and (b) The red color square and zoomed-in image and TEM image after erasing. Adopted from [52]. Copyright 2012 Nature Publishing Group. Conductance quantization with the synaptic trend of Ta2O5-based atomic switching: (c) observed increment in quantized conductance owing to the cell upon application of positive voltage pulses of width 20 ms during a 2-s interval. The inset shows the controlled-conductance level realized by inserting a series combination with the cell along with a current-controlling resistor with a resistance of 3 kΩ, (d) observed decrement in quantized conductance owing to the cell upon application of negative voltage pulses of width 20 ms during a 2-s interval. The input pulses are directly connected to the device, shown in the inset view, and (e) The quantized conductance histogram demonstration from measured data with zoomed-in image of conductance against time. Adopted from [53]. Copyright 2012 IOP Publishing.
Nanomaterials 13 02443 g001

2.1.2. Valence Change Mechanism

VCM and ECM exhibit the same trends with respect to electrochemical reaction and ion mobility mechanisms. ECM is often considered as an electrochemical reaction involving active metals, whereas VCM is also electrochemical reaction, though it is based on OVs in oxide materials. Generally, the device resistance change is significantly affected by oxygen-related defects in electrochemical reactions. The extensive literature has reported that the transition between HRS and LRS corresponding to the desired memory device (such as RRAM) is caused by the damage or formation of OV filaments [3,54,55,56,57]. The steps involved in the VCM process are presented in Figure 2a–f. When positive biasing is applied at an inert TE, the oxide insulating layer becomes active in LRS; O2− changes their original position, and the OVs alternatively produce a conductive path along with drifting toward the anode/oxide-layer interface. In this way, a conductive path containing OVs is eventually created, increasing the conductivity owing to the functional layer film.
Conversely, upon application of the reverse voltage, the formed conductive channels are gradually destroyed, and the RRAM device turns transitions to the HRS. Consequently, O2−/OVs may be responsible for developing and rupturing conducting channels in RRAMs. Szot et al. reported the locality in conductivity related to SrTiO3 films and discussed it in detail by employing atomic force microscopy with a conducting tip, observing that SrTiO3 films may demonstrate bistable conductance switching between nonmetallic and metallic behavior upon application of an electrical field [56]. In addition, a later study reported in favor of bipolar resistance switching behavior, resulting in an ideal conductive channel among the Pt tops upon being influenced by Ta-rich clusters belonging to asymmetric TaOx, which were directly perceived by an in-situ detection approach [58]. Notably, VCM generates the OV – O2− anti-Frenkel pairing. However, the formation of OV – O2− anti-Frenkel pairing has not been reported in the literature [59,60]. The formation of OV – O2− anti-Frenkel pairing would cause instability and relaxation [47].

2.1.3. Thermochemical Mechanism

RRAM cells commonly show resistance variation with the application of an electric field and Joule heating-based O2− reactions associated with drifting that dominate the operation mechanism [61]. In these RRAM cells, the forming process is attributable to thermal rupturing within the device medium and the subsequent formation of CFs. By contrast, the reset process corresponds to thermal melting of the dominant CFs [62,63]. Moreover, unipolar/bipolar modes are considered more effective in RRAM cells owing to the free Joule heating development without a polarity effect. The different steps involved in TCM are presented in Figure 3a–g. The nonpolar switching in resistance change has exhibited a particular trend in some previous studies. Jung et al. demonstrated a Pt/NiO/Pt memory structure and discussed temperature-dependent resistance change in switching dynamics in NiO films, revealing the corresponding OFF-state current evaluated with respect to temperature and defect patterns.
The Pt/TiO2/Pt memory filamentary arrangement is openly determined during resistance variation by employing high-resolution TEM [64]. In situ voltage–current (V–I) curves at a lower temperature (130 K) evidenced the creation and rupturing of filaments owing to TinO2n-1. Upon increasing OV density, triangular-type filaments containing TinO2n-1 of the Magneli phase were slowly formed in the set process. This result was obtained using the Fast Fourier transform, from which the diffraction pattern of Ti4O7 was simulated. When the thermal influence is considered with a reversed external field, TinO2n-1 is converted back into TiO2 within the prevailing triangular state of the filament, where it rearranges the O2− concentration. Hence, Joule heating always plays a pivotal role in rupturing the CFs, which is broadly recognized. However, there is still no prominent experimental confirmation verifying the factor accountable for the formation of CFs. Several factors cause the creation and disruption of CFs, including electrode range, reaction, and switching layer thickness [65]. Additionally, various materials show a variety of storage mechanisms in memory devices. The above summary is only for filament mechanisms; however, specific mechanisms are still under discussion [66]. Because the physical mechanism for the RS phenomenon has not yet been elucidated, it should be focused on in future investigations [47].
Figure 3. Schematic depiction of thermochemical process showing the RS of (a) pristine state, HRS under (b) Joule heating and (c) pyrolysis, (d) set process associated with conducting channel creation, LRS under (e) Joule heating, (f) reset process associated with CF breakdown, and (g) reversion to the set process. Adopted from [67]. Copyright 2018 John Wiley & Sons, Ltd.
Figure 3. Schematic depiction of thermochemical process showing the RS of (a) pristine state, HRS under (b) Joule heating and (c) pyrolysis, (d) set process associated with conducting channel creation, LRS under (e) Joule heating, (f) reset process associated with CF breakdown, and (g) reversion to the set process. Adopted from [67]. Copyright 2018 John Wiley & Sons, Ltd.
Nanomaterials 13 02443 g003

3. Unipolar and Bipolar CeO2-Based Memristors

RRAM is a nonvolatile memory, showing two logical resistance states of the switching channel. It characteristically involves an insulating or semiconducting layer, mainly composed of high-k oxide nanomaterials sandwiched between two electrodes as shown in Figure 4a. These memory devices commonly exhibit hysteresis performance with current and voltage characteristics (i.e., negative differential resistance at a specific applied potential). In the SET state, when the applied voltage and current approach a threshold level, RRAM cells switch from the HRS (OFF-state) to the LRS (ON-state). Similarly, upon application of a specific voltage, the RRAM cells return to their OFF-state when the prevailing tendency is removed. In nonpolar RRAM, as shown in Figure 4b, the SET and RESET modes work with unidirectional applied voltage. By contrast, in bipolar RRAM, the resistance may vary step-wise, showing two distinct values; however, some devices exhibit gradual resistance variation depending upon the previous path of charges [68].
Sun et al. reported the bipolar switching mechanism in the Al/CeOx/Pt cell associated with multilevel resistance variation through the creation and rupturing of CFs containing OVs by valence change from Ce3+ to Ce4+ cations [14,18]. Many researchers have demonstrated a digital-type resistance variation for RRAM applications by utilizing the CeO2 switching layer [14,15,16,70] through typical I–V curves for the RS trend involving the TE area (0.25 mm2), employing a sweeping voltage at 300 K. Figure 5a depicts the statistics of the configuration for SET and RESET voltages. Younis et al. presented the switching operation evaluated through OVs in the Au/CeO2/ITO structure, as shown in Figure 5b [70]. In addition, Hsieh et al. [15] and Ismail et al. [16] have discussed the bipolar switching behavior of Al/CeO2/Au and Zr/CeOx/Pt structures, respectively, as shown in Figure 6c–e. Further, Lin et al. reported the same trend in the CeO2 conductive layer for Pt/CeO2/Pt structures as shown in Figure 5g [18]. Despite this, the analog switching of a CeO2 layer is still being researched [18]. An analog memristive switching along with an excellent resistance variation by a few orders of magnitude in a Pt/CeO2/Pt memristive nature cell has been investigated. Pt/CeO2/Pt possessed symmetric Pt electrodes (both top and bottom); therefore, it showed significant polarity-dependent and asymmetric diode RS mode as depicted Figure 5f [18]. In Younis et al.’s study [71], bipolar reversible RS behavior was observed when a self-assembled CeO2 nanocube-derivative layer of ~162-nm thickness was decorated on the Au-coated Si-substrate. Compared to the thin CeO2 layer, the bulky and uneven grains attributed to CeO2-dependent nano cubes exhibit extra OVs, thereby leading to improved performance in the form of an excellent OFF/ON ratio, small forming potential with a reasonable resistance variation process. Further, the regular cubic-type crystals occupying increased OVs at the surfaces rather than their counterparts contribute to reducing the set voltage while rearranging themselves in forming regular percolation paths, as it is deficient in bulk-type thin films. In addition, a CF model containing OVs with perfect distribution has been suggested to describe the RS process.
Bipolar resistive switching (BRS) and unipolar resistive switching (URS) have been achieved in a double-layer Ag/Ti/CeO2/Pt structure under different sweep rates along with CC. In BRS, the RS trend depends upon ECM, whereas TCM is dominated by URS functioning (Figure 6a,b). No forming process and small RS voltage make the device more attractive [39]. Here, a sweeping voltage mode associated with a sweeping potential (0.005 V) at a rate of 0.09 V/s is followed. Further, sweeping voltage operation in the order of 0 V → 1 V → −1 V → 0 V favors bipolar behavior of the device (as in Figure 6a). The device shows the HRS of 2800 Ω with a sweeping potential of 0–1 V. Then, the resistance state changes to LRS, exhibiting 270 Ω at 0.3 V. In this case, the CC (1 mA) is observed. Negative biasing is dominated by CF [72]. In addition, I–V curves exploring the space-charge-limited current (SCLC) obey Ohm’s law (I ∝ V) and Child’s law (I ∝ V3/2). Furthermore, to confirm the bipolar switching operation, double-log I–V curves are also depicted. However, the primary defect lying in the device is its deviation from the SCLC mechanism during RS functioning [39]. To establish the repeatability and reproducibility of bipolar switching behavior, a total of 120 cycles were executed. During this experiment, the SET voltage was maintained at 0.3 V, whereas the RESET voltage spanned from −0.3 to −0.6 V. Notably, no compliance current (CC) was attained during the RESET phase, thereby enabling the realization of maximum achievable current levels.
By contrast, a current of 2 mA was set during SET state, whereby URS behavior was observed as shown in Figure 6b. Further, the current increased to 1.4 V and the resistance states changed to SET. Afterward, at a positive voltage (without limiting current), the device changed from LRS to HRS operating at a potential of 0.7 V (but without the forming step). TCM [73,74] was followed to explain the URS trend associated with the CFs during the LRS state. The local current density in the conductive channel becomes very large, resulting in Joule heating, which sharply raises the temperature of the conductive filament. Then, the conducting channel breaks while returning the device to a high resistance. Further, bipolar RS characteristics were explored in a bilayer TaN/CeO2/TiO2/Pt structure. The set voltage was found to be dominated by CeOx layer thickness, indicating the switching process to be an electric field-induced mechanism. A CeO2/TiO2 ultrathin film was deposited by radio-frequency magnetron sputtering with an Ar/O2 mixture, followed by annealing at different temperatures (350 °C, 450 °C, 500 °C, and 550 °C. The annealed samples had decreased roughness and set/reset voltages (~1.1 and ~1 V) at 500 ºC as shown in Figure 6c. Here, the TiN/CeOx interface performed the principal role of controlling OV defects, thus controlling the creation/deforming of conductive paths. This device also showed cycling endurance with an HRS-to-LRS ratio of 105 with excellent performance under retention/stress test (104 s). Cycling endurance with data-retention performance was acceptable [75]. The coexisting RS behaviors were comprehensively characterized for the Ti/CeO2/Al/CeO2/Pt device. Similar SET and RESET voltages and similar current levels with the exact current compliance in the SET process validated that the device operation was very simple. Both coexisting RS trends were identical during the LRS and HRS modes. However, the switching uniqueness among both modes was investigated in the CeO2 devices involving an Al capping layer. The BRS characteristics of lower voltage with optimum operation window in CeO2/TiO2-based resistive memories with multilevel switching were analyzed. The CeO2/TiO2 ultrathin film was decorated through a radio-frequency (RF) magnetron sputtering route in an Ar/O2 environment, followed by annealing at various temperatures. The annealing process considerably reduced the unevenness and set/reset window of ~1.1/~1 V at 500 °C. A more negative value of Gibbs free energy owing to TiO2 compared to CeO2 leads to a more straightforward reoxidation process for filamentary oxygen exchange with a tantalum nitride electrode. In this respect, cumulative probability distribution of fundamental memory parameters has exhibited reasonable configuration of the SET/RESET potential. The said device offered outstanding endurance properties, such as DC switching operation at >2500 cycles, data retention >105 s, and a resistance ratio of >102 at a temperature of 500 °C [76].
Figure 6. (a) Demonstration of electrochemical process for the creation and rupture of conducting silver filaments corresponding to log (I)–V characterization for the Ag/Ti/CeO2/Pt structure. (b) Schematic depiction of the URS for the Ag/Ti/CeO2/Pt device, with I–V graphs for this memory device indicating unipolar operation trend. Adopted from [32]. Copyright 2020 Elsevier B.V. (c) Schematic illustration of the conducting channels in a TaN/CeO2/TiO2/Pt cell with corresponding graphs showing ideal BRS behavior at annealing temperatures of 350 °C, 450 °C, 500 °C, and 550 °C. Adopted from [76]. Copyright 2017 Elsevier B.V.
Figure 6. (a) Demonstration of electrochemical process for the creation and rupture of conducting silver filaments corresponding to log (I)–V characterization for the Ag/Ti/CeO2/Pt structure. (b) Schematic depiction of the URS for the Ag/Ti/CeO2/Pt device, with I–V graphs for this memory device indicating unipolar operation trend. Adopted from [32]. Copyright 2020 Elsevier B.V. (c) Schematic illustration of the conducting channels in a TaN/CeO2/TiO2/Pt cell with corresponding graphs showing ideal BRS behavior at annealing temperatures of 350 °C, 450 °C, 500 °C, and 550 °C. Adopted from [76]. Copyright 2017 Elsevier B.V.
Nanomaterials 13 02443 g006
The bipolar RS behavior of CeO2−x ultrathin layers with an ultrathin TiO contacting layer under an Ar/O2 mixture in the chamber is shown in Figure 7a. The titanium oxide layer with an appropriate Ar/O2 mixture ratio is considered more suitable for the stability as well as uniformity of the RS mechanism. Moreover, a CeO2−x film deposited in an Ar/O2 (6:24 ratio) environment exhibited outstanding RS behavior owing to the decisive role of existing OVs, as shown in Figure 7b–e. In the study, the SCLC with Schottky emission was the dominant transportation in the stronger-field segment. Ideal swapping homogeneity (ON/OFF ratio of 102), device-to-device uniformity with the best DC endurance of >500 cycles, retention performance of 104 s validates it as a suitable nonvolatile RRAM device [77]. The coexistence of both cerium ions (i.e., Ce+3 and Ce+2) in the oxygen defective CeO2-x layers decorated over an ITO/glass substrate by RF magnetron sputtering was also studied. Furthermore, the Ni/CeO2−x/ITO/glass structure displayed a multilevel RS trend as shown in Figure 7f, which was achieved by controlling the RESET-stop potential in the adjustment of SET-CC. The reproducible switching effect and excellent endurance of ~103 DC cycles with a retention time of >104 s was observed. I–V characteristics revealed the Schottky emission transportation mechanism at RESET-stop potentials ranging from ~1.5 to ~2.0 V as shown in Figure 7g,h. These results favor the said device to be beneficial for multilevel information-storage applications [78]. The Pt/CeO2/Pt device showed the necessary forming process under application of a potential of 9 V and a current of 5 mA (at CC). Semilogarithmic I–V curves for the device were plotted under unipolar voltage sweeps and ON/OFF state, where the resistance ratio could approach five orders of magnitude. The ON/OFF states were observed with a stability of 104 s at a potential of 0.3 V, showing exceptional nondestructive readout characteristics. Furthermore, the symmetric electrode structure of the same device exhibited RS functioning free from biasing in the sense that both positive and negative potentials may switch ON/OFF states to OFF/ON states (not shown here). In this study, the OFF state was reported to be less stable than the ON state.

4. Threshold CeO2-Based Memristors

Threshold-switching-type memristors have become attractive for selectors of crossbar memory structures. The threshold switching memristor involving a single layer of CeO2 transforms into a new memory over repeated cycles caused by an overgrowth of Ag filament, thereby changing the device to LRS. Hence, optimal threshold switching was suggested in the aforementioned diffusive memristors by taking CeO2 as a switching layer composed of Ag filaments (with low voltage) combined with a SiO2 diffusion barrier layer in controlling the filament formation to achieve an advanced memory device. In this respect, an amorphous-type SiO2 layer was placed underneath the CeO2 layer, behaving as an obstacle for the overgrowth of silver metal filaments and thereby enhancing the device stability. This tailored device has exhibited a forming-free volatile threshold switching nature (with 1 pA OFF-state current) and excellent selectivity of 104 with tuned uniformity and endurance characteristics. Moreover, Ag/CeO2/SiO2/Pt has shown outstanding direct current for 103 I–V cycles with low density at switching voltages. Because of such results, the implementation of CeO2-based threshold RRAMs as the best selectors is proposed when crossbar memories are adopted [79]. The I–V characteristics of both Ag/CeO2/Pt cell and Ag/CeO2/SiO2/Pt cells under 50 consecutive cycles are presented in Figure 8a,b, and the endurance performance of the two devices under pulse operation is shown in Figure 8c,d. The Ag/CeO2/SiO2/Pt structure exhibited stability in endurance performance for 104 cycles, sustaining an on/off ratio of 103. A pulse rate of 1 with 0.6 V and duration of 0.64 ms was applied for switching operation in the abovementioned devices by incorporating a read pulse of 0.2 V for a duration of 0.64 ms as shown in Figure 8e. Threshold switching of the Ag/CeO2/SiO2/Pt structure occurs under variable CC. However, an increment in the ON-state with CC is also investigated, as shown in Figure 8f.

5. Electrode-Based CeO2 Memory Devices

5.1. Symmetric Electrode-Based CeO2 Memristors

Memristors containing both electrodes (the same in nature) are regarded as symmetric electrode-based memristors. The widely used CeO2 memristors of this type are focused on in this section. In this respect, being symmetric electrode-based, spin-coated Pt/CeO2-x/Pt with a polycrystalline phase and mixed with fluorite cubic structures having 40-nm thickness of Pt layer are deposited through DC magnetron sputtering. The energy bands attributed to semiconductor–metal contact involving surface states with different reverse biases and the XPS spectrum during 30-s sputtering revealing the binding energy are shown in Figure 9a,b. Two fundamental damages occur in the CeO2-x film of Pt/CeO2-x/Pt cell during electroforming phenomena. Surface defects are generated in Pt/CeO2-x at the TE contact, which provides the source to create positive charge carriers into CeO2-x film. During the set process, electron tunneling is accompanied by OV ionization in Pt/CeO2-x at the TE junction. The various trapping states occurring at the metal–semiconductor junction during metal electrode fabrication is evaluated in the form of surface damage. The repeated potential field owing to the crystal is disturbed at the surface, resulting in extra energy levels. Every surface atom is ascribed to a single surface energy level existing in disallowed states. Therefore, ionized OVs would be allowed to be dispersed at energy levels, owing to which electrons are confined in trapping sites. Moreover, the greater radius of Ce3+ than that of Ce4+ may produce an extended lattice and distortion. The distortion in lattice and grain boundaries often breaks up the lattice order. Consequently, the ionized OVs drift from the TE toward the BE, and many OVs are detected as electron-trapping sites in the bulk state. The I–V characteristics, Schottky emission fitting curves, and F-N tunneling fitting lines are displayed in Figure 9c. Further, a schematic illustration of the Schottky barrier altitude optimization occurring at the bottom CeO2/Pt contact is shown in Figure 9f. It results from OV redistribution with initial and final application of positive voltage existing at the top Pt electrode when OVs are transferred toward the bottom contact as shown in Figure 9f [32]. The Pt/CeO2/Pt memristors were prepared by depositing a 50-nm-thick CeO2 resistance-based layer on a Pt electrode on the Ti/SiO2/Si substrate through RF magnetron sputtering when the CeO2 target was selected with an Ar medium. Subsequently, an annealing process (400 °C–600 °C for 1 h) was also performed in ambient air. A platinum TE having a diameter of 100 μm was deposited by electron beam evaporation.
Further, the Pt/CeO2/Pt structure exhibited various current values with the alteration of positive voltage sweeps, which gave higher values than negative voltage sweeps. The I–V curves of the reference device under annealing at 500 °C with 20 repeated voltage sweeps are shown in Figure 9g; the I–V curves reveal that the device is electrically nonsymmetric because of the bottom Pt/CeO2 junction formation, as CeO2 is deposited through sputtering. By contrast, the top Pt/CeO2 junction is established through Pt electrode deposition on the CeO2 layer through evaporation. In addition, the annealing process is performed after CeO2 film deposition, leading to a bottom interface, whereas the top interface does not experience annealing. Consequently, asymmetric electrical characteristics are achieved [26]. RS characteristics have also been exhibited by completely transparent ITO/CeO2/ITO devices. The physical schematics with endurance performance of such a device are presented in Figure 9j,m. CeO2 ultrathin films containing 15–25 nm thin layers are decorated by RF magnetron sputtering. Additionally, a CeO2 ceramic target is decorated over ITO coated with 200-nm-thick glass substrate in an Ar–O2 (ratio of 20:10) environment as illustrated in Figure 9k–m. For RS results, a 75-nm-thick of ITO TE with a diameter of 150 μm is deposited by RF magnetron sputtering. When ITO is deposited, the pressure in the chamber is controlled to a value of 10 m Torr under Ar gas atmosphere with 100-W sputtering power [80].
Figure 9. (a) Energy band diagram for semiconducting metal interface and surface states at different reverse biases. (b) Ce 3d deconvolution XPS results for 30-s sputtering sample. I–V trends for samples with fitting curves: (c) Schottky emission fitting curves, (d) Schottky emission fitting lines, (e) F–N tunneling fitting lines. Adopted from [81]. Copyright 2022 Elsevier B.V. (f) Schematic display of Schottky barrier-height tunning occurring at the CeO2/Pt junction by OV redistribution when positive voltages are applied to the top Pt- electrode, extracting OVs near the bottom contact. (g) I–V curves of the reference device under annealing of Pt/SiO2/CeO2/SiO2/Pt cell at 500 °C. (h) XPS spectra of the Ce 3d levels for nonannealing and at 400 °C, 500 °C, and 600 °C annealing for Pt/CeO2/Pt memristors. (i) Retention performance of Pt/CeO2/Pt memristors at 500 °C annealing upon repeated application of +3–+7 V pulses with 30 repetitions and measurement of current during 1800 s. Adopted from [26] Copyright 2022. Elsevier B.V. (j) Arrangement schematic of the ITO/CeO2/ITO device with I–V curves showing the starting forming mechanism with prior sweeping of the BRS trend for ITO/CeO2/ITO devices containing CeO2-layers with different thicknesses: (k) 15 and (l) 20 nm. (m) Endurance performance of the ITO/CeO2/ITO device. Adopted from [80] Copyright 2015 Elsevier B.V.
Figure 9. (a) Energy band diagram for semiconducting metal interface and surface states at different reverse biases. (b) Ce 3d deconvolution XPS results for 30-s sputtering sample. I–V trends for samples with fitting curves: (c) Schottky emission fitting curves, (d) Schottky emission fitting lines, (e) F–N tunneling fitting lines. Adopted from [81]. Copyright 2022 Elsevier B.V. (f) Schematic display of Schottky barrier-height tunning occurring at the CeO2/Pt junction by OV redistribution when positive voltages are applied to the top Pt- electrode, extracting OVs near the bottom contact. (g) I–V curves of the reference device under annealing of Pt/SiO2/CeO2/SiO2/Pt cell at 500 °C. (h) XPS spectra of the Ce 3d levels for nonannealing and at 400 °C, 500 °C, and 600 °C annealing for Pt/CeO2/Pt memristors. (i) Retention performance of Pt/CeO2/Pt memristors at 500 °C annealing upon repeated application of +3–+7 V pulses with 30 repetitions and measurement of current during 1800 s. Adopted from [26] Copyright 2022. Elsevier B.V. (j) Arrangement schematic of the ITO/CeO2/ITO device with I–V curves showing the starting forming mechanism with prior sweeping of the BRS trend for ITO/CeO2/ITO devices containing CeO2-layers with different thicknesses: (k) 15 and (l) 20 nm. (m) Endurance performance of the ITO/CeO2/ITO device. Adopted from [80] Copyright 2015 Elsevier B.V.
Nanomaterials 13 02443 g009

5.2. Asymmetric Electrode-Based CeO2 Memristors

Asymmetric electrode-based memristors contain different TE and BE and show comparatively superior characteristics than symmetric electrode-based devices. Therefore, CeO2-based asymmetric memristors are discussed in detail in this section. The Ag/CeO2/Pt device is prepared by providing a high-vacuum environment during magnetron sputtering and a coating structure exhibiting asymmetric electrode behavior. Ceria layers having a thickness of 27 nm are decorated over a Pt/Ti/SiO2/Si substrate by RF magnetron sputtering. Then, a silver TE with a diameter of 200 μm is deposited on the as-prepared film with a shadow mask. The ceria thin film exhibits good compactness, which is confirmed by its average roughness of 0.38 nm. This ceria film shows sufficient crystallinity and pure phase [32]. Meanwhile, an asymmetric Ag/CeO2/Pt device is favored according to Wang et al. Ag filaments incorporated into the ceria conducting layer are vertically stacked in the order of BE to TE [82], which significantly favors Ag+ ions with high mobility compared to electrons because the ceria lattice provides a fast diffusion channel for Ag+ ions [32].
In the Ag/CeO2/Pt memristive structure, the variation in the corresponding rate and voltage range of the device gradually changes the resistance. It is controllable through the creation and rupture of silver filaments; the conduction phenomena corresponding to each state are also investigated. The semilogarithmic I–V curves, endurance performance under 100 repetitions, retention characteristics depending on ambient temperatures, fitting results for three levels before the forming process, and transition of states are shown in Figure 10a–f. The resistance transition of the memristor is stable during 100 voltage sweeps; thus, the resistance states are retained for 5 × 104 s [32]. In another study, epitaxial CeO2 films were synthesized via pulsed laser deposition. In this structure, 0.7-wt.% Nb-doped SrTiO3 single crystals with sizes of 3 and 0.5 mm were used for the deposition substrate. The Pt TE was decorated over a CeO2 layer by DC magnetron sputtering along with a shadow mask having many holes with an aperture of 0.2 mm. Furthermore, the indium BE was tarnished on NSTO substrates for forming an ohmic contact. This device exhibited outstanding RS behavior with an RS ratio of 3 × 104. A significant photoresponse was detected in this device when it was illuminated with a 405-nm laser beam, which facilitated fine-switching operation at HRS. In addition, both light-controlled RS associated with voltage-controlled photo-responsiveness were evaluated in the device. These RS and photoresponse properties are ascribed to the Schottky barrier transport in the Pt/CeO2 junction, whereas electron trapping/de-trapping occurs because of OV insertion into the interface. The main focus is to explain the binding energy levels using the XPS spectra of CeO2/NSTO at the Ce 3d level. Furthermore, I–V characteristics under darkness and 405-nm illumination and retention behavior at LRS, IRS1, IRS2, and HRS in the dark were obtained. The retention and endurance at LRS, IRS1, IRS2, and during illumination; resistance levels during dark and illumination under different voltage pulses; short-circuit current; and open-circuit voltage during illumination with the provision of voltage pulses are presented in Figure 10d,e. Because of such peculiarities, they may afford potential applications for different levels of RS memories along with various functions in photoelectric sensors [83].
To add, a layer 10-nm-thick CeO2 film is decorated over the Pt BE of the Pt/Ti/SiO2/Si substrate, which is synthesized by RF magnetron sputtering. Further, a 5-nm-thick ZnO thin film is used as a second dielectric layer that is decorated over CeO2/Pt/Ti/SiO2/Si with the same parameters following the CeO2 layer deposition. The final structure is the double-layered structure of the M/ZnO/CeO2/Pt device, for which alternative electrodes such as Ti, TaN, and TiN are deposited at normal temperature. Here, magnetron sputtering is adopted with a metal shadow mask to be circular as TEs (100-nm thick) and with a diameter of 150 μm. Further, the effect of Gibbs free energy on oxidation with the electronegativity of the TE material is also analyzed. It has been observed that decreased operational voltages may occur incrementally because of the work function gap between the TE and BE. Interfacial contact between the TE and the oxide layer (ZnO) influences the RS phenomenon. The forming and subsequent RS functioning of various double-layered structures, such as M/ZnO/CeO2/Pt, TaN/ZnO/CeO2/Pt, TiN/ZnO/CeO2/Pt, Ti/ZnO/CeO2/Pt, and Ni/ZnO/CeO2/Pt, are elaborated in this study. Statistical information related to SET and RESET voltages, resistance states, and switching ratio of the M/ZnO/CeO2/Pt devices are presented as well. The devices containing TaN (TE) have exhibited excellent cycling uniformity corresponding to switching parameters, SET and RESET voltages, HRS and LRS resistances, and remarkable endurance properties (>1000 cycles) with LTS > 105 s [84]. Moreover, the endurance performance of TaN-TE, TiN-TE, and Ni-TE with a schematic demonstration revealing he conduction phenomena in TaN/ZnO/CeO2/Pt devices is provided. In addition, pristine resistance trends showing the interfacial barriers and LRS at the SET voltage and HRS at the RESET voltage are considered in this study.

6. Single-Layer and Bilayer CeO2 with Capping Layer

Most previous studies have reported that the RS behavior of CeO2 ultrathin films often requires a forming voltage [17]. On the contrary, a study has reported that Ce-silicate can also be structured at the CeOx/Si junction [85]. In this way, creating Ce-silicate may increase the density of OVs at the contact area [86]. Hence, Si (buffer layer) plays a role at the contact point and in the appearance of OVs, thereby facilitating the modulation of the forming process, possibly by CeO2 for achieving reversible RS. Additionally, the probability of the suppressed forming process is expected to be assured through interface engineering. However, inserting a Si buffer layer at the bottom contact would form an RRAM cell having the W/CeO2/Si/TiN structure with a small forming voltage and CC, offering a reasonable window and endurance features compared to those devices (without Si). It is expected to improve the development of Ce-silicate with the appearance of OVs at the interfaces caused by the dynamic production of conducting channels toward the forming process. Figure 11a,b reveals the RS characteristics of the W/CeO2/TiN and W/CeO2/Si/TiN devices. These devices were operated on similar grounds, and the sweeping voltage order was from 0→10, 0→−3→0, and 0→3→0 V in distinguishing the forming and set/reset processes, respectively. With the insertion of Si, the forming voltage was reduced from 7.2 to 2.8 V, and CC decreased from 5 to 1 mA. Results showed that the device containing the Si layer did not exhibit stability in the window; however, a distinguishable window greater than 10 was observed. Further, the typical RS role with the forming process, SET/RESET states, HRS/LRS during 10 switching cycles, cyclic stability of devices owing to W/CeO2 (20 nm)/Si (1 nm)/TiN as well as W/CeO2 (20 nm)/Si (2 nm)/TiN, respectively, were presented.
Moreover, the behavior of W/CeO2/Si/TiN is easily improvable by optimizing the thickness of the Si buffer layer. Furthermore, following the symmetric conducting channel model, we may describe the consecutive mechanism of rupturing channels depending upon the reset voltage. Hence, the abovementioned results may provide a suitable path for voltage control to optimize device performance and deep insight into the slow and steady reset process [87]. Figure 11c,d shows the SEM morphology with a graphic depiction of Ag/CeO2/La0.5Sr0.5CoO3 decorated over a sapphire substrate. It is obtained using the focused ion beam technique with a tungsten probe constantly mounted over the sample. Then, fabrication and investigation of Ag/CeO2/La0.5Sr0.5CoO3 comprising different levels are elaborated. The La0.5Sr0.5CoO3 (500-nm thick) and ceria (80-nm thick) layers were decorated over the sapphire substrate. Pulsed laser deposition at 700 °C was employed in an oxygen environment with a pressure of 150 mTorr. The pulse frequency in the laser ablation was set at 20 Hz and maintained with an energy density ranging from 2 to 3 J/cm2. A thermally evaporated 300-nm-thick Ag thin film was realized at normal temperature. The Ag/CeO2 micro interface was engineered through an ion beam over the La0.5Sr0.5CoO3 layer that exhibited renewability and reversibility in switching to an HRS state (OFF state) with insulating characteristics. By contrast, a semiconducting/metallic LRS state (ON state) showed resistance ratios up to 104.
In this study, the formation of cells with micron and submicron values was more effective. The impact of micro-scaling associated with microdefects was evaluated at the cell boundary, and the electrical properties were evaluated through ion etching. The I–V curves of the Ag/CeO2/La0.5Sr0.5CoO3 cells with dimensions of 1–150 × 150, 2–15 × 30, 3–5 × 10, and 4–2 × 4 μm were measured. Ideal I–V results prior to the electrical forming process corresponding to different cell sizes are illustrated in Figure 11e,f. The Ag/CeO2/La0.5Sr0.5CoO3 structures were evaluated for breakdown voltage (which increased with a decrement in contact area). It is possibly caused by the existence of defects within the film. In this research, the high-resistance nature of the I–V results corresponding to Pool–Frenkel coordinates with Ln(I), V1/2 scaling was observed. In this process, the stimulating influence toward conductivity is explained by the Pool Frenkel effect, Schottky effect, and hopping conducting channels [88].
MnO/CeO2 heterostructures were prepared on Pt BE (150-nm thick) by RF magnetron sputtering at CeO2 and MnO targets in an Ar atmosphere at normal temperature. In addition, the MnO layer top area was decorated with an ~100-μm-diameter Ag TE by thermal evaporation with a shadow mask. Further, the Ag/MnO/CeO2/Pt cells were evaluated in terms of the stability in BRS characteristics, showing a high resistance ratio during the forming-free mechanism. In this way, the device always remains in HRS. Accordingly, a positive voltage applied on the TE will form the LRS, while a negative voltage on the Ag TE will lead to HRS, confirming and displaying the typical BRS behavior of these structures. In this case, the electrical forming process is not needed. Consequently, the forming-free process is associated with the creation of many interfacial states and defects into oxide layers [89].
Moreover, oxide materials such as NiO and CeO2 in the form of thin films are spin-coated over pure Si/ITO substrates. The reversible BRS showing optimal ON/OFF ratio has been reported to be ~103. The bilayer species have exhibited a threefold improved ON/OFF ratio compared to the nanocomposite films. The charge carrier transport mechanism depends upon various p-n depletion structures. For composite films, the space charge field enlarges toward the inner p-n junction, leading to insensitive bias sweeping. In a study, the p-n junction was found to affect the production or rupture of CF channels [90]. The RS characteristics for memory devices depending upon the CeO2-x/La0.8Sr0.2MnO3 bilayer were also investigated. The BRS was evaluated in La0.8Sr0.2MnO3 at the micrometer scale and reproduced the results achieved at the nanoscale through conductive scanning force microscopy. Figure 11g presents the schematic illustration of the oxygen (O2) ion migration mechanism in a metal/CeO2-x/La0.8Sr0.2MnO3 fabricated device. In the study, I–V characteristics corresponding to individual memristors were typically investigated under systematic voltage sweeps (0→ +Vmax →0 → −Vmax→ 0). The measured curves of the CeO2/La0.8Sr0.2MnO3 structures were evaluated via a probing process over Ag electrodes as shown in Figure 11h. The forming process was required to induce BRS behavior and was established at low-voltage operation. Further, endurance efficiency with a resistance ratio of 102–103 at a reading voltage of −1 V was observed with consecutive memory functioning during 100 cycles as shown in Figure 11i. In addition, single La0.8Sr0.2MnO3 layers displayed stability in RS realized by metallic connections of probing operation in the form of oxygen exchange within the medium. Further, it was observed that inserting a CeO2-x layer between two electrodes and the metallic-nature La0.8Sr0.2MnO3-layer prominently altered RS properties, thereby allowing the production of good microelectronics devices. In the as-prepared devices, the BRS trend was analyzed in a single La0.8Sr0.2MnO3 layer, where the SET/RESET processes were activated at minimal voltages with an enhanced resistance ratio. Therefore, the CeO2-x layer (as an oxygen conductor) behaved like an O2− reservoir, facilitating easy ion exchange with low-voltage operation [91].
The oxygen annealing effect attributed to the RS properties of bilayer TiN/ZnO/CeO2-x/Pt structures was analyzed. Schematic illustrations of the RS mechanism corresponding to un-annealing and annealing of TiN/ZnO/CeO2-x/Pt structures are presented in Figure 11j The switching trend for annealing and un-annealing effects on these devices revealed that with a small increment in the forming voltage, there may be a SET/RESET voltage drop. In this study, the endurance performance was critically improved in annealed oxygen ambient at an optimal temperature of 400 °C (with reduced OVs) in the ZnO/CeO2-x film. Additionally, Schottky emission may have higher voltage limitations for annealed and unannealed (at 400 °C) structures. It is inferred that annealed devices in an oxygen environment improve the crystalline arrangement as the maximum number of defects are removed to improve the uniformity of RS characteristics at an appropriate forming voltage. Finally, the aforementioned upgrading relates to (Ce+4 to Ce+3) transition states. Accordingly, the transition may lead to OVs in the lattice creating charge balancing with an increase in their density in a bulk-type device interface. Hence, fewer CFs are produced, which is attributed to an allowed arrangement of crystallinity at the optimum annealing. Moreover, newly created CFs may provide an easy way for rupturing in the RESET process, thereby improving the memory window, as it is the basic need for improved nonvolatile memory characteristics [92]. Another bilayer TiN/HfO2/CeOx/TiN memristor has been reported to adopt RF magnetron-sputtering. The as-prepared devices were thermally annealed at various temperatures for performance advancement. Moreover, compared with unannealed devices, the effect on the coefficient of variation corresponding to SET/RESET voltages dropped by 35.1% and 59.4%, respectively. By contrast, the coefficients of variations in resistances attributed to resistances in LRS and HRS may decrease by 70.2% and 52.7%, respectively, under annealing at 400 °C for 2 min in an air medium. We performed I–V measurements depending upon the annealing effect by employing X-ray diffraction (XRD) and XPS analysis. It is concluded that the overall influence owing to grain boundary reduction and decreased OV density in the switching layer is responsible for improved switching parameters of TiN/HfO2/CeOx/TiN devices. In this way, this study provides us with a simple way to enhance memristor performance [93].
Figure 11. (a) Typical RS performance of the W/CeO2 (20 nm)/TiN and W/CeO2 (20 nm)/Si (1 nm)/TiN devices, (b) comparison of HRS and LRS during 50 cycles for RRAMs with and without 1-nm Si buffer layer. Device stability window without and with Si (distinguishable window greater than 10). Adopted from [87]. Copyright 2012 Elsevier B.V. (c) SEM photographs, (d) schematic view of Ag/CeO2/La0.5Sr0.5CoO3 decorated over sapphire substrate and analyzed by the focused ion beam technique, and (e) typical I–V measurements under electrical forming for various cell sizes. Adopted from [88]. Copyright 2014 Elsevier B.V. (f) Schematic demonstration of oxygen (O2−) ion movement in the metal/CeO2-x/La0.8Sr0.2MnO3 cell. (g) Graphical representation of the probe scanning over Ag electrodes for CeO2/La0.8Sr0.2MnO3, the forming step is needed to induce the BRS, which is maintained at low voltage operation. (h) Endurance cycling measurement of LSMO structures. The resistance ratio covering the range of 102–103 with a −1V read voltage is measured. The supplementary view representing successive memory functioning in 100 cycles is observed. Adopted from [91]. Copyright 2015 Elsevier B.V. Schematic representation of the RS mechanism (i) without annealing and (j) annealing treatment in TiN/ZnO/CeO2-x/Pt structures [92]. Copyright 2018 Elsevier B.V.
Figure 11. (a) Typical RS performance of the W/CeO2 (20 nm)/TiN and W/CeO2 (20 nm)/Si (1 nm)/TiN devices, (b) comparison of HRS and LRS during 50 cycles for RRAMs with and without 1-nm Si buffer layer. Device stability window without and with Si (distinguishable window greater than 10). Adopted from [87]. Copyright 2012 Elsevier B.V. (c) SEM photographs, (d) schematic view of Ag/CeO2/La0.5Sr0.5CoO3 decorated over sapphire substrate and analyzed by the focused ion beam technique, and (e) typical I–V measurements under electrical forming for various cell sizes. Adopted from [88]. Copyright 2014 Elsevier B.V. (f) Schematic demonstration of oxygen (O2−) ion movement in the metal/CeO2-x/La0.8Sr0.2MnO3 cell. (g) Graphical representation of the probe scanning over Ag electrodes for CeO2/La0.8Sr0.2MnO3, the forming step is needed to induce the BRS, which is maintained at low voltage operation. (h) Endurance cycling measurement of LSMO structures. The resistance ratio covering the range of 102–103 with a −1V read voltage is measured. The supplementary view representing successive memory functioning in 100 cycles is observed. Adopted from [91]. Copyright 2015 Elsevier B.V. Schematic representation of the RS mechanism (i) without annealing and (j) annealing treatment in TiN/ZnO/CeO2-x/Pt structures [92]. Copyright 2018 Elsevier B.V.
Nanomaterials 13 02443 g011

7. Doped CeO2-Based Memristors

The RS improvement trends in CeO2-based memristors have been analyzed through charge dynamics by using Al as the dopant. Here, results have indicated that Ti/CeO2: Al/Pt device sandwiching, shown in Figure 12a, always offers promising and attractive switching characteristics. In Figure 12b, an Al-doped device in the negative-forming mode is depicted by the left inset, while the right inset shows the variation in endurance. Further, the minimum forming voltage modulated the stability in SET/RESET voltages. In addition, an improved endurance (>104) of switching cycles having window of ROFF/RON > 102 was observed compared to the undoped Ti/CeOx/Pt structure, as shown in Figure 12c,d. The observed precision in the memory switching trend corresponds to a substantial decrease in the formation energy for OVs for increasing OV creation in CeO2 layers corresponding to charge transfer as well as oxygen-obtaining capability of the Al dopant [94]. In this study, the Y3+-doped CeO2 memristor was analyzed for RS characteristics with operating current and voltage. This RRAM device was fabricated by forming Ce1-xYxO2-y ultrathin films using a photochemical metal–organic deposition strategy. It could be obtained with the direct patterning of Ce1-xYxO2-y films on a specific substrate via ultraviolet (UV) irradiation without photoresist and etching processes [95,96].
The photochemical metal–organic deposition procedure is utilized as it is comparatively stable against humidity and temperature owing to the beneficial photochemical metallic–organic precursor (main material), whereas chemical process is primarily initiated under UV irradiation. XRD as well as Raman spectrum results have increasingly revealed the occurrence of lattice shrinkage which induces and increases the structural alteration caused by OV formation with the Y dopant. Ce1-xYxO2-y devices have exhibited bipolar filamentary behavior of RS and consequently both resistance states due to the decline in forming voltage with an increment in Y doping-induced rise in OV density. However, the ON/OFF ratio decreases with an increment in Y density owing to reducing HRSs. Moreover, when the recovery of injecting oxide area associated with HRS switching is rendered, it causes defect sites comprising bounded electrons generated by yttrium. Spectro-microscopy revealed that ceria loses 24% oxygen during LRS, displaying a resistance variation of five orders of magnitude [97]. Three elements (Ca, Zr, and Gd) have been utilized as dopants in CeO2 film sandwiching into Pt/CeO2/Pt devices. In this case, the switching characteristic has been found to be unipolar. The electrical measurement examination for Gd and Zr dopants in the Pt/CeO2/Pt structure exhibited a resistance ratio of 104, improved retention stability with 104 s at 25 °C, and endurance of 1500 sweeping cycles.
By contrast, a Ca-doped device demonstrated the opposite trends. However, all results showed that chemical defects with OV concentration in metallic Gd and Zr have a decisive impact compared to Ca doping in CeO2 films. Further, XPS analysis indicated vital oxygen-containing film activation in forming conducting channels, leading to resistance alteration within the conducting layer [98]. Further, the chemical-bonding energy obtained by XPS analysis revealed comparable undoped components; the Gd/Zr-doped cells possessed lower CeO2-bonding energy, leading to stability in configuration conversion. On the contrary, the Ca-doped components showed higher CeO2-bonding energy, leading to unstable configuration conversion. In conclusion, the said dopants in ultrathin films afforded resistance stability, and TEM results indicated that Zr dopant was much better than Gd dopant for the RS behavior of the CeO2-based RRAM devices. Thus, transition metals as dopants in CeO2 nonvolatile memory have shown improved memory functioning, better HRS/LRS ratio, attractive endurance, and fine retention capabilities. Therefore, with the overview of the experimental results, we can confirm that this memory is a promising candidate for future nonvolatile RS memory devices [99].

8. Neuromorphic Applications

The term “neuromorphic” was coined by Carver Mead to describe analog circuits that emulate biological neurons. Over recent years, the neuromorphic computing field has experienced remarkable advancement, including innovations in neuromorphic chip engineering that aim to replicate specific behaviors exhibited by components of the brain circuitry. Within the realm of neuromorphic circuitry, memristors have gained prominence owing to their resemblance to certain types of neurons [100].

8.1. Single-Layer and Bilayer CeO2-Based Synaptic Devices

Metal oxide-based memristors possess the potential for the advancement of scalable and highly efficient brain-inspired computing systems. Among these systems, oxide-based memristors display inherent electronic analog switching similar to biological synapses. For example, CeO2-based bilayer memristors occupying forming-free, low-voltage potential, and high energy efficiency and reliability play a vital role in biological synapses. Because of pulse measurement, analog behavior may be directly programmed in intermediate resistance states in such devices. Therefore, STDP may be implemented, strictly following spike-dependent Hebbian artificial learning. In this case, the memristive device exhibits substantial variation in the normalized synaptic strength (>30 times) while overlapping the presynaptic and postsynaptic neuron spiking. Accordingly, recent research has demonstrated a decisive role in the physical network with high density and excellent connection in neural networks [101].

8.2. Spike-Timing-Based Learning System

Moreover, CeO2 with an HfOx capping layer exhibits spike-based-learning operations that have shown novel behavior in terms of high efficiency and compactness for managing unstructured information. In these systems, the learning process is necessarily followed by spike-based Hebbian learning, proving that it is STDP, as it clearly describes the synapse strength variation during a time interval corresponding to presynaptic and postsynaptic neuron spiking. To perform these activities, a mean spiking rate of 1 MHz (105 times) is evaluated in the human brain. Moreover, the period is considered as 1 s to refresh the neurons’ existing position while calculating synaptic currents, whereas neuron-spiking probability is taken as 0.01 in the central nervous system. In Figure 13a, the plot of normalized conductance variation versus time difference of the pre/post-spike interval demonstrates synaptic behavior. The fitted data allied with exponentially decaying functions evidences the STDP behavior analogous to the human brain synapse. Further, the data showed a significant variation corresponding to normalizing conductance in this device, which was greater than 30 times in the state of overlapped pre/postsynaptic spikes [101].

8.3. Potentiation and Depression Behaviors

A symmetric electrode, such as a Pt/CeO2/Pt device, has demonstrated synaptic-dependent potentiation and depression behaviors showing polarity-oriented analog switching. Here, the synaptic weight variation is controlled by parameters such as amplitude, width, and pulses, resulting in an artificial synaptic device for neuromorphic computing systems [18]. Figure 13b illustrates the steady and successive current increment at a read voltage of +2 V with repeated pulses of amplitude +10 V and a pulse width of 20 ms with 100 repetitions. This study revealed the analog transition with an HRS-to-LRS ratio > 105, which is better than already reported data. Because of reversible switching, LRS returns to HRS with the application of repeatable −10 V pulses with widths of 1, 10, and 100 ms during 50 repetitions. By contrast, the current with a read voltage of +2 V under low voltage pulsing was 10−10 A, and read disturbance became negligible. The analog decrement and increment in resistance followed the potentiation and depression in synaptic tendency, respectively. More importantly, the optimization in resistance variation, which is ~1 MΩ–200 GΩ), permitted the synapse toward extensive limitation in synaptic weight functioning to generate a signal, thereby establishing learning and memory functions.
The current variation may be controlled by the voltage pulse amplitude along with the width, as illustrated in Figure 13c,d. In addition, the current is generated with a read voltage of +2 V during higher voltage pulsing 30 times. At a specific pulse width of 10 ms, the current dynamics at LRS may increase (~20 to 100 nA) with the pulse amplitude increment of +7–+10 V, as shown in Figure 13c. Likewise, the current may increase by ~10–170 nA when the pulse width is increased from 1 to 50 ms under a specific pulse amplitude of +10 V, as demonstrated in Figure 13d. Moreover, both cases maintain the reversible HRS at negative voltage pulses. The retention mechanism for guessing short-term/long-term synaptic memory with the pulsing operation. With the current increment (tens of times) with positive potentiation pulsing, there may be a decrement in retention with time, leading to memory loss. Interestingly, it decreased speedily in the initial 30 s, whereas it decreased more slowly thereafter. However, the current may be four or five times greater than the primary measurement after 10 min. The result may be more significant/comparable when retention becomes favorable for short-term memory in biotic systems analogous to the human brain, caused by momentary recalling of information during the processing time that usually lasts (seconds to tens of minutes) [102]. On the contrary, improved retention may be mandatory during the long-term memory (LTM) process. In addition, the memory loss in the system may be considered to be explored through irregular electrical charging/recovery (Ce3+ to Ce4+), thereby resulting in the annihilation of OVs when O2− are recombined. The O2− ions must instantaneously stop moving toward the Pt electrode to prevent this recombination. Consequently, the memory loss mechanism and improved retention properties are still questionable [18].

8.4. Potentiation Motion and Synaptic Weight Decay

Figure 14a illustrates the consecutive increment in current at +2 and +6 V during repeated pulses and width of 50 ms with 30 repetitions in potentiation dynamics, and Figure 14b shows normalized memory retention when potentiation is slowly reduced during time t. Therefore, synaptic weight decay is equivalent to the human memory–forgetting curve, the STP, whereas the gradual decay of final synaptic weight denotes LTP [102]. On the contrary, biotic synapses manage the input information and correspondingly rearrange the memory states toward a synapse. Figure 14c illustrates a multistore model system for human psychological memory, as presented in the work of Atkinson and Shiffrin. Furthermore, in this model, the external information may be assessed and later stored for various hierarchy stages, thereby originating from sensual memory, momentary short-term memory (STM), and perpetual LTM by attention learning associated with repetition/rehearsal progressions. In addition, Figure 14 illustrates the enlarged current with memory transition under repeating input pulses to verify the rehearsal scheme, and the trend of STM toward LTM transition with repeated pulsing processes coincides with the multistore model as shown in Figure 14c. Moreover, the STM-to-LTM transition has been constantly detected by increasing the pulse width.
Further, quantitative analysis has revealed the distinguished rehearsal effect to enhance memory retention. The gradual decrement in retention with time and the decaying rate significantly decreases with increasing number of stimulation (N). Herein, the rehearsal mechanism with repeating pulses (N = 70) may significantly enhance the synaptic weight variation as 75–618 nA, while the relaxation time is 24–580 s, and is firmly following the tendency reported by Chang et al. [102], as the resistance variation due to OV diffusion is improved and stabilized under repeated stimulation. In addition, the memory retention curves are closely fitted, which is attributable to the stretching relaxation function exponentially depending upon the relaxation time. Thus, depression is confirmed by the pulsing mode results from a quicker and unexpected decaying current closer to the original state, while the retention loss with time results from more gradual incompleteness in memory loss. Hence, synaptic weight modulation in connection with stable memory is caused by memory loss with time, which may be strengthened by increasing the pulse frequency.

8.5. Short- and Long-Term Plasticity

Analog synaptic weight modulation has been proved to show linearity and symmetric behavior. Therefore, it demonstrates LTS through resistance variation in a Pt/ITO/CeO2/Pt memristor rather than Pt/CeO2/Pt memristive memory, possessing nonlinearity and asymmetric resistance variations. The same resistive memristor displays linearity and symmetric resistance variation trends associated with various voltages with opposite polarities toward synaptic potentiation and depression functions. Additionally, the memristor also exhibits considerable LTS concerning controllable synapse weight with time, which may be extracted from the capping layer (ITO). It behaves like a source of OIs in the CeO2 layer for retaining resistance variation. In the previous literature, biological synapses have exhibited memory-forgetting time depending on input levels. Here, a comparative study of memristors without and with a capping layer may elucidate the proper functioning of the ITO layer in terms of linearity, symmetric nature, and long-term stable behavior in resistance variation in these memristors for application in artificial synapses in neuromorphic computing [19].
By contrast, classic nonvolatile memory may disclose long-term retention of memory levels. Additionally, the current decline over time in the potentiation process for the Pt/CeO2/Pt device and Pt/ITO/CeO2/Pt memristor is considered comparable. The current decay comparison with identical ranges and pulses of various amplitudes is performed in these memristors. Primarily, the reference memristor demonstrates voltage pulses as +6–+9 V with a corresponding width of 50 ms applied for 10 cycles, and subsequently, the read current generated by +2 V during 10 cycles in 10 s. Further, to analyze the current decay during 100 s, all steps are repeated again 10 times. The above procedure repeats itself in memristor functioning by utilizing a pulse amplitude of +8–+11 V. Further, with the current increment associated with potentiation pulses, both memristors display current-decay functioning during specific intervals, representing forgetting operations based on biotic memory. Moreover, the quick current decay, analogous to STP, shows a gradual decline toward LTP [102]. After decay for 100 s, further increment in the current is observed with the following pulse, corresponding to PPF features, which are characterized by additional synaptic weight by applying two close consecutive input pulses. Application of a high pulse amplitude will result in less current decay associated with enhanced PPF. Furthermore, the retention characteristics are then investigated by measuring current in the absence of potentiation pulses.
In Figure 15a, the reference Pt/CeO2/Pt memristor displays current increment values as 7.5–49 nA under +9 V pulses and later decreases to 11 nA with a retention time of 1000 s. By contrast, the Pt/ITO/CeO2/Pt memristor achieved high in retention performance. In particular, the current increases from 6.9 to 37 nA during the potentiation process, which subsequently declines to 16 nA during 1000 s, as depicted in Figure 15b. Surprisingly, retention the results of both memristors are prominently connected with the resistance variation mechanism, which is thus difficult to reveal owing to instantaneous resistance change by applying a voltage that in turn decays over time. In this respect, retention operation in favor of the double-layer Pt/ITO/CeO2/Pt device is enhanced by increasing the pulse repetition number n, as demonstrated in Figure 15c,d. For instance, the current-increase trend shows an ΔI(0) of ~140 nA generated with 3000 repetitions, followed by a decaying trend, with ΔI(t) = ~50 nA during 500 s. On the contrary, ΔI(0) becomes ~3.3 nA during 100 repetitions with the pulse rate, and ΔI(t) = ~0.8 nA during 500 s. Finally, speedy decay with a smaller n value is attributed to STM features.
Moreover, better retention with increased n values consistent with LTM, significantly mimicking identities during the transition from STM to LTM, is realized in several memory hierarchy stages in the natural brain. Figure 15c demonstrates the linear scaling for the repeating pulses caused by a larger ΔI, resulting in speedy decay at the preliminary level despite the retained window becoming larger. Subsequently, it contracts the gap among the current stages corresponding to repeating pulse number. Despite the decaying process, the current stages are distinct, confirming the multilevel probability or analog synaptic weight modulation concerning repeating pulse numbers. Thus, analog characteristics such as linear behavior, symmetric structure, and prolonged stability in synaptic weight modulation are successfully investigated for the memristive device under discussion. In addition, more parameters such as operating speed, voltage level, and power consumption are considered critical factors in the synaptic procedure. The appropriate engineering of the aforementioned factors may be critically performed in applications for artificial synapses and neuromorphic computing [19]. Artificial synapses with analog resistance variations are actively investigated because they are fundamental elements for brain-inspired neuromorphic computing.
Recent information has suggested interface-state-oriented artificial synapse modulation. In particular, analog behavior, linear trend, and symmetric synaptic weight in Pt/CeO2/Pt devices are based on subsequent deposition on a heated CeO2 layer. Further, memristive characteristics have been studied at optimized (400 °C and 500 °C) annealing temperatures, showing consistent and analog-nature resistance decrement at positive polarity and reverse at negative polarity. This critical resistance alteration is extensively responsible for linearity and symmetry during potentiation and depression functioning, which may be tunable through variation in pulse number with amplitude. In addition, STP and LTP have also been investigated under different pulse conditions. In addition, conducting channels are tailored through the Schottky conducting phenomena caused by energy barrier junctions. Moreover, minor resistance variations have been investigated in the reference memristor, but with the controlled insertion of SiO2 interfacial layers at both electrodes concerning the Pt-CeO2 junctions, thereby fabricating Pt/SiO2/CeO2/SiO2/Pt structure. Here, it has been concluded that the resistance variations are strictly voltage-driven, thereby tuning the Pt/CeO2 junctions. Necessarily, controllable interfacial layers are considered to be decisive toward synergetic synaptic operations [26].

8.6. Pavlov’s Dog Experiment-Based Synoptic Study

More recently, an artificial neural network has been considered the hottest research spot for advanced science and technology. The CeO2 exhibits excellent performance among various oxide-based materials, such as longer retention with healthier stability. However, its use in artificial neural synapses is still questionable. In this regard, a Ag/CeO2/Pt memristive-type device has been realized toward the selected target. Herein, the artificial synaptic functions are successfully explored, and a sequential neuromorphic system simulation has been updated. In addition, the relations among the pulse arrangement parameters involving resistance states owing to synapse devices have also been investigated.
A study tailored an electrically generated signal simulation during Pavlov’s dog experiments. These findings demonstrate the device’s significance for realizing important applications in artificial neural synaptic simulations [39]. To achieve these targets, bionic synaptic appearances are analyzed afterward. The biological synapses show that excitatory neurotransmitters are released through activation of the presynaptic sheath, which produces current reverting into the postsynaptic sheath. Here, Ag acts as the neurotransmitter to connect both electrodes (top and bottom) under application of a biased voltage signal. Then, the current is abruptly increased by using the voltage signal at the TE. When the write voltage of 0.275 V is removed, the current returns to the starting level prior to stimulation during a fixed time, thereby exhibiting STP trend, as the excitatory postsynaptic current is produced at that level.
To demonstrate the further relationship between time and the attenuation owing to excitatory postsynaptic current, Figure 16a,b shows that two consecutive pulse signals are intended to simulate PPF, significantly exploring STP movement in the neuroscience literature. Moreover, to simulate potentiation, two consecutive positive pulses are operated; thereby, 0.35 V is achieved in the writing process. Furthermore, to simulate the depression effect, two negative pulses must be applied while achieving a V-write of −0.25 V. Moreover, both simulated synaptic activations indicate that the current is read out through a small voltage amplitude, which is called V-read. The aforementioned synaptic trend usually occurs when double-stimulated signals are generated during a short time. These studied synaptic weights are given by Equation (1). The curve is fitted using Equation (2). Here, pulse stimulations are illustrated in Figure 16c,d, where weight updates are.
ΔW = (G2G1)/G1,
PPF = (G2G1)/G1 × 100% = C1exp(−t/τ1) + C2exp(−t/τ2).
G1 and G2 denote the first and second pulse conductance obtained from the current measurements (A1, A2), as shown in Figure 16a,b. Moreover, t denotes the pulse interval, and C1 and C2 indicate the first facilitation values for the concerning phases; τ1, τ2 represent conductance magnitudes for the first and second pulses as a representative of synaptic weights. At the forward signal, τ1 and τ2 values are observed as 2.74 and 15.94 ms, respectively. At a negative signal, the achieved values of τ1 and τ2 are 12.7 and 12.71 ms, respectively. The logic behind the large gap between τ1 and τ2 (under positive pulses) is the easier formation of CFs than fractures, resulting in fast activation in the synaptic study. To increase the response current, the pulse interval is reduced so that successful simulation of the impact of consecutive signals and signal duration law in neurosynaptic trends may be achieved. As in the previous literature, synaptic plasticity may be characterized as STP and LTP. STP may merely be continued for a short time after pulse signals. On the contrary, LTP may extend to several hours/days without resistance variation. Figure 16c,d demonstrates device conversion (STP to LTP) when a pulse cluster with 0.1-V amplitude, 20-ms interval, with 5-ms duration is applied. The response current abruptly enhances when a pulse is applied and quickly returns to the previous state. Again, when a pulse group with 0.3-V amplitude and 5-ms interval and 20-ms duration is applied, there may be a gradual decrease in the response current. Upon removing the pulse, the response current still shows stability, whereas the resistance shows a lower value than the initial level; conversion from STP to LTP is also accomplished.
Further, the associations among pulse train variables, such as duration, amplitude, and interval, correspond to resistance states. The device resistance is examined by using these controllable variables, as depicted in Figure 17a–c, showing their effect on resistance modulation for forward stimulation; further, Figure 17d–f demonstrates the effects under negative stimulation. The impact of speed and degree of resistance variation is positively linked with the with pulse amplitude in specific intervals, whereas they are negatively attached in pulse interval trains. These results mimic learning by human intelligence. For instance, when an individual tries to recall some memory, the same individual may easily forget if the recitation intensity is inadequate.
On the contrary, LTM may be comprehended when the recitation intensity may strengthen. Consequently, the resistance level may vary intensely upon the application of a forward pulse. Interestingly, only a few pulses are required to implement the transition of the resistance state, which is caused by Ag atoms not returning to the TE during the reversing reset voltage. Thus, STP, LTP, training learning, and complex learning with forgetting functioning in the biotic brain may be simulated. In this regard, forgetting roles analogous to those in the natural brain are used to understand the forgetting process. First, a learning pulse is applied 25 times, whereby the optimum conductance after the initial learning process is proposed to be 100% for the memory intensity and is analogous to human learning behavior, identical to first-time learning as shown in Figure 18a. Then the conductance is read every 5 ms after removing the learning pulse. It is significantly achieved by memory intensity detection, which decays quickly at first, indicating a similar trend as that in human STM. Nevertheless, memory intensity will stabilize gradually during increasing time intervals, finally maintaining 75% retention intensity, attributed to human LTM. Afterward, the identical pulse learning signal is repeated.
When the learning pulse is removed in an identical time interval, the device’s memory intensity attenuation becomes lower than the initial level. By contrast, stabilized learning intensity becomes higher than the initial level. This is analogous to the human brain, as it could attain stronger memory than in early times by repeating learning processes caused by forgetting during specific periods. Figure 18b demonstrates retention array information, which could be achieved by increasing the pulse amplitude, whereas writing data intensity may be stronger than the previous level. This way, the device is realized to find short-term and long-term neuro status toward storage and memory fields.
In addition, bionic learning (Figure 18a) of the device is impossible by employing pulse training experiments intended to simulate familiarity testing (Pavlov’s dog). Figure 18b shows that the pulses group with 0.1-V amplitude is utilized for simulating ringing stimulation, while the pulses group with 0.3-V amplitude is utilized for simulating food stimulation; the excitatory postsynaptic current is used for the simulation of the dog’s response stimulation. Here, a current of 0.9 mA is set as a reaction boundary; the response current may be produced through a stimulation greater than 0.9 mA. Then, the reaction of the device is considered to be drooling for the dog learning.
Further, upon application of the bell-pulse stimulation, the memory does not show an optimum responsive state. However, the device exhibits a higher response state upon using food pulse stimulation. Then, ring pulse as well as food pulse are executed to simulate dog training course. This time, the device exhibits a higher responsive state. Furthermore, in training, when a bell-pulse stimulus is used, the device reaches a higher responsive state and maintains the running state during subsequent 20 bells stimulation, as shown in Figure 18c. Consequently, the desired memory positively simulates the standard Pavlov’s dog experiment comprising highly learned efficiency with excellent memory features [39].

8.7. Latest Development in Ceria-Based Neuromorphic Computing

In this section, we explore potentiation and depression behaviors that exhibit remarkable linearity and symmetry. These characteristics arise owing to the incorporation of a lower concentration of OVs within the Gd-doped ceria (GDC) layer when combined with CeO2. This leads to oxygen vacancy redistribution and unique synaptic weight redistribution behaviors. The distinct trends in synaptic weight update for polarity-dependent potentiation and depression voltages are attributed to the stacking order of CeO2 and GDC as depicted in Figure 19. We demonstrate how potentiation and depression functions can be studied using +4V−4V as potentiation/depression voltages for the GDC-top device and +5V−4V for the CeO2-top device, each with distinct consecutive functions at corresponding voltages. These behaviors are further explained through voltage-controlled energy barrier modulation at the junction of CeO2 and GDC or at the Pt electrode junction, which is influenced by oxygen vacancy redistribution.
Furthermore, we showcase a theoretical simulation that employs an improved handwritten digit patterning approach via a bilayer perceptron neural network. This simulation yields an average accuracy of 88%, encompassing dynamic range, linearity, symmetry, and accuracy of states. The synaptic properties explored are also demonstrated through 32 × 32 crossbar arrays of GDC-top memristors, underscoring the potential of two-layer memristors in pioneering the implementation of artificial synapse networks for novel applications in neuromorphic computing [103].
For biological synapses, one of the most crucial neuromorphic characteristics is STDP. STDP demonstrates that when the pre-spike signal precedes the post-spike signal (Δt > 0), it leads to long-term potentiation (LTP), resulting in an increase in synaptic weight (w). Conversely, if the post-spike follows the pre-spike (Δt < 0), long-term depression (LTD) occurs, leading to a decrease in synaptic weight (w). Figure 20a depicts the waveform designed to induce the STDP phenomenon using continuous single pulses. Notably, the first time slot is occupied by a negative pulse, while a positive pulse with reduced amplitude appears in the subsequent time slot. When pre-spike and post-spike signals overlap, the programming pulse corresponds to the necessary amplitude for resistance modulation. Additionally, the voltage drop across the device arises from the disparity between pre-spike and post-spike voltages. When spike timings are closely aligned, the negative pulse often overlaps with the positive pulse, yielding a higher amplitude and consequently more significant resistance variation. If the pre-spike precedes the post-spike, a positively programmed pulse emerges (left side of Figure 20a), whereas the opposite results in a negatively programmed pulse (right side of Figure 20a).
Furthermore, Figure 20b showcases the simulation of the STDP learning rule applied to the current device. For Δt > 0 with Δw > 0, LTP is observed, whereas for Δt < 0 with Δw < 0, LTD is significantly induced. However, tighter spike timings result in more substantial resistance modulation. Notably, experimental learning data following the STDP rule reveals an impressive fit with exponential functions, supported by fundamental fitting parameters, as also demonstrated in Figure 6b. The equation capturing this relationship is provided as follows:
Δ W = A e x p Δ t τ + Δ w 0 ,  
In this context, the symbol ΔW represents the updates in synaptic weight. A is employed as a scaling factor, and τ signifies the time constant. Moreover, the device’s adjustable conductance, synaptic saturation, and simulation of the STDP learning rules collectively demonstrate the considerable potential of Pt/Ti/AlOx/CeOx/Pt storage devices for adaptive features within neuromorphic systems. Consequently, the device exhibits synaptic characteristics, including adjustable conductance, synaptic saturation, and the ability to simulate STDP learning rules. These aspects highlight its significant potential for application in artificial neuromorphic computing [104]. Copyright © 2022 Wenqing Wang et al.
Following this comprehensive discussion on recent advancements in the field of neuromorphic computing based on ceria, we aim to delve into the domain of neuromorphic dynamics linked with NWNs. This area has also explored brain-inspired neuromorphic systems, wherein synapse-like interfaces establish neural networks to facilitate artificial learning processes, as elucidated by Nakayama et al. [105] Notably, NWNs epitomize an exemplary neuromorphic system. These networks can be categorized as metal/metal-oxide NWNs, atomic switch NWNs, and metal/polymer NWNs. Each of these systems comprises a network and exhibits switching behavior that hinges on the choice of materials or schemes employed [105].
In conclusion, it is imperative to acknowledge that the abovementioned work offers a level of flexibility, particularly in the context of ceria-based neuromorphic nanowires. This flexibility serves as an encouraging factor for researchers exploring this intriguing field.

9. Conclusions and Perspective

High-k oxide-based materials are extensively suggested as promising memristors; however, they are still challenging in neuromorphic studies. CeO2 has become more efficient and advanced in RRAM technology, as it covers the maximum gap left behind in the memory field. Therefore, CeO2-based memristors have been widely used because they show substantial behaviors in the filamentary mechanism and classification, optimum switching behaviors (unipolar, bipolar, and threshold) depending upon the nature of electrodes (symmetric and asymmetric), and switching layers (single with capping layer, and doped-insulating layer). In this context, various RS devices have been discussed in detail in this review. Further, because of their highly efficient behavior in the RS mechanism, CeO2-based synaptic devices have been suggested in this work, exhibiting unique neuromorphic characteristics compared to other high-k oxide-based memristors. These devices afford optimum RS behavior and improved outcomes in their synaptic behavior for neuromorphic artificial systems. Therefore, this review strictly emphasizes CeO2-based artificial synaptic devices. They are based on spike-based learning, potentiation, and depression, STP and LTP, potentiation motion, and weight decay. Finally, Pavlov’s dog experiments for the advanced synaptic study of CeO2-based memristors are systematically focused on, revealing recent advances in synaptic devices. High-k oxide materials may be the best alternative to compensate for the weak areas indicated in this review.

Author Contributions

S.A., M.A.U. and A.R. Literature, Ideas, and Writing—original draft. M.W.I., M.F.K., M.R. and M.I.: Writing—review and editing. S.K.: Writing—review and editing, Supervision, Resources, Project administration, and Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korea government (MOTIE) under Grant 20224000000020 by the National Research Foundation of Korea (NRF) grant funded by the Ministry of Science and ICT (2021K1A3A1A49098073).

Data Availability Statement

Not applicable.

Acknowledgments

The authors are thankful to all individuals and organizations permitted to republish their figures and other relevant information.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Kim, K.M.; Jeong, D.S.; Hwang, C.S. Nanofilamentary resistive switching in binary oxide system; a review on the present status and outlook. Nanotechnology 2011, 22, 254002. [Google Scholar] [CrossRef] [PubMed]
  2. Hermes, C.; Wimmer, M.; Menzel, S.; Fleck, K.; Bruns, G.; Salinga, M.; Bottger, U.; Bruchhaus, R.; Schmitz-Kempen, T.; Wuttig, M.; et al. Analysis of Transient Currents During Ultrafast Switching of TiO2 Nanocrossbar Devices. IEEE Electron Device Lett. 2011, 32, 1116–1118. [Google Scholar] [CrossRef]
  3. Kwon, D.-H.; Kim, K.M.; Jang, J.H.; Jeon, J.M.; Lee, M.H.; Kim, G.H.; Li, X.-S.; Park, G.-S.; Lee, B.; Han, S.; et al. Atomic structure of conducting nanofilaments in TiO2 resistive switching memory. Nat. Nanotechnol. 2010, 5, 148–153. [Google Scholar] [CrossRef] [PubMed]
  4. Snider, G.S. Spike-timing-dependent learning in memristive nanodevices. In Proceedings of the 2008 IEEE International Symposium on Nanoscale Architectures, Anaheim, CA, USA, 12–13 June 2008; pp. 85–92. [Google Scholar]
  5. Pershin, Y.V.; La Fontaine, S.; Di Ventra, M. Memristive model of amoeba learning. Phys. Rev. E 2009, 80, 021926. [Google Scholar] [CrossRef] [PubMed]
  6. Prezioso, M.; Merrikh Bayat, F.; Hoskins, B.; Likharev, K.; Strukov, D. Self-Adaptive Spike-Time-Dependent Plasticity of Metal-Oxide Memristors. Sci. Rep. 2016, 6, 21331. [Google Scholar] [CrossRef] [PubMed]
  7. Panda, D.; Tseng, T.-Y. Perovskite Oxides as Resistive Switching Memories: A Review. Ferroelectrics 2014, 471, 23–64. [Google Scholar] [CrossRef]
  8. Yan, Z.B.; Liu, J.M. Resistance switching memory in perovskite oxides. Ann. Phys. 2015, 358, 206–224. [Google Scholar] [CrossRef]
  9. Luo, Z.; Lau, H.K.; Chan, P.K.L.; Leung, C.W. Resistive Switching in Perovskite-Oxide Capacitor-Type Devices. IEEE Trans. Magn. 2014, 50, 1–4. [Google Scholar] [CrossRef]
  10. Wong, H.S.P.; Lee, H.Y.; Yu, S.; Chen, Y.S.; Wu, Y.; Chen, P.S.; Lee, B.; Chen, F.T.; Tsai, M.J. Metal–Oxide RRAM. Proc. IEEE 2012, 100, 1951–1970. [Google Scholar] [CrossRef]
  11. Hochstetter, J.; Zhu, R.; Loeffler, A.; Diaz-Alvarez, A.; Nakayama, T.; Kuncic, Z. Avalanches and edge-of-chaos learning in neuromorphic nanowire networks. Nat. Commun. 2021, 12, 4008. [Google Scholar] [CrossRef]
  12. Yang, N.; Belianinov, A.; Strelcov, E.; Tebano, A.; Foglietti, V.; Di Castro, D.; Schlueter, C.; Lee, T.-L.; Baddorf, A.P.; Balke, N.; et al. Effect of Doping on Surface Reactivity and Conduction Mechanism in Samarium-Doped Ceria Thin Films. ACS Nano 2014, 8, 12494–12501. [Google Scholar] [CrossRef]
  13. Foglietti, V.; Yang, N.; Aruta, C.; Di Pietrantonio, F.; Cannatà, D.; Benetti, M.; Balestrino, G. High plasticity reversible resistive switching in heteroepitaxial metal/CeO2−x/Nb:SrTiO3/Ti/Pt structures. Nanotechnology 2016, 27, 375705. [Google Scholar] [CrossRef] [PubMed]
  14. Sun, Y.; Kiewra, E.W.; Souza, J.P.d.; Bucchignano, J.J.; Fogel, K.E.; Sadana, D.K.; Shahidi, G.G. High-Performance In0.7Ga0.3As -Channel MOSFETs With High-κ Gate Dielectrics and α-Si Passivation. IEEE Electron. Device Lett. 2009, 30, 5–7. [Google Scholar] [CrossRef]
  15. Hsieh, C.-C.; Roy, A.; Rai, A.; Chang, Y.-F.; Banerjee, S.K. Characteristics and mechanism study of cerium oxide based random access memories. Appl. Phys. Lett. 2015, 106, 173108. [Google Scholar] [CrossRef]
  16. Ismail, M.; Huang, C.-Y.; Panda, D.; Hung, C.-J.; Tsai, T.-L.; Jieng, J.-H.; Lin, C.-A.; Chand, U.; Rana, A.M.; Ahmed, E.; et al. Forming-free bipolar resistive switching in nonstoichiometric ceria films. Nanoscale Res. Lett. 2014, 9, 45. [Google Scholar] [CrossRef]
  17. Lin, C.-Y.; Lee, D.-Y.; Wang, S.-Y.; Lin, C.-C.; Tseng, T.-Y. Reproducible resistive switching behavior in sputtered CeO2 polycrystalline films. Surf. Coat. Technol. 2008, 203, 480–483. [Google Scholar] [CrossRef]
  18. Kim, H.J.; Zheng, H.; Park, J.-S.; Hun Kim, D.; Jung Kang, C.; Tae Jang, J.; Hwan Kim, D.; Yoon, T.-S. Artificial synaptic characteristics with strong analog memristive switching in a Pt/CeO2/Pt structure. Nanotechnology 2017, 28, 285203. [Google Scholar] [CrossRef]
  19. Kim, H.J.; Kim, M.; Beom, K.; Lee, H.; Kang, C.J.; Yoon, T.-S. A Pt/ITO/CeO2/Pt memristor with an analog, linear, symmetric, and long-term stable synaptic weight modulation. APL Mater. 2019, 7, 071113. [Google Scholar] [CrossRef]
  20. Kim, H.J.; Park, D.; Yang, P.; Beom, K.; Kim, M.J.; Shin, C.; Kang, C.J.; Yoon, T.-S. Synaptic characteristics with strong analog potentiation, depression, and short-term to long-term memory transition in a Pt/CeO2/Pt crossbar array structure. Nanotechnology 2018, 29, 265204. [Google Scholar] [CrossRef]
  21. Kim, H.; Gilmore, C.M.; Piqué, A.; Horwitz, J.S.; Mattoussi, H.; Murata, H.; Kafafi, Z.H.; Chrisey, D.B. Electrical, optical, and structural properties of indium–tin–oxide thin films for organic light-emitting devices. J. Appl. Phys. 1999, 86, 6451–6461. [Google Scholar] [CrossRef]
  22. Yan, X.; Zhao, Q.; Chen, A.P.; Zhao, J.; Zhou, Z.; Wang, J.; Wang, H.; Zhang, L.; Li, X.; Xiao, Z.; et al. Vacancy-Induced Synaptic Behavior in 2D WS2 Nanosheet–Based Memristor for Low-Power Neuromorphic Computing. Small 2019, 15, 1901423. [Google Scholar] [CrossRef]
  23. Rajendran, B.; Sebastian, A.; Schmuker, M.; Srinivasa, N.; Eleftheriou, E. Low-Power Neuromorphic Hardware for Signal Processing Applications: A Review of Architectural and System-Level Design Approaches. IEEE Signal Process. Mag. 2019, 36, 97–110. [Google Scholar] [CrossRef]
  24. Burr, G.W.; Shelby, R.M.; Sebastian, A.; Kim, S.; Kim, S.; Sidler, S.; Virwani, K.; Ishii, M.; Narayanan, P.; Fumarola, A.; et al. Neuromorphic computing using non-volatile memory. Adv. Phys. X 2017, 2, 89–124. [Google Scholar] [CrossRef]
  25. Guo, S.; Arwin, H.; Jacobsen, S.N.; Järrendahl, K.; Helmersson, U. A spectroscopic ellipsometry study of cerium dioxide thin films grown on sapphire by rf magnetron sputtering. J. Appl. Phys. 1995, 77, 5369–5376. [Google Scholar] [CrossRef]
  26. Park, K.; Chung, P.H.; Sahu, D.P.; Yoon, T.-S. Interface state-dependent synaptic characteristics of Pt/CeO2/Pt memristors controlled by post-deposition annealing. Mater. Sci. Semicond. Process. 2022, 147, 106718. [Google Scholar] [CrossRef]
  27. Wang, W.; Gao, S.; Li, Y.; Yue, W.; Kan, H.; Zhang, C.; Lou, Z.; Wang, L.; Shen, G. Artificial Optoelectronic Synapses Based on TiNxO2–x/MoS2 Heterojunction for Neuromorphic Computing and Visual System. Adv. Funct. Mater. 2021, 31, 2101201. [Google Scholar] [CrossRef]
  28. Seo, S.; Kang, B.-S.; Lee, J.-J.; Ryu, H.-J.; Kim, S.; Kim, H.; Oh, S.; Shim, J.; Heo, K.; Oh, S.; et al. Artificial van der Waals hybrid synapse and its application to acoustic pattern recognition. Nat. Commun. 2020, 11, 3936. [Google Scholar] [CrossRef]
  29. Lan, J.; Cao, G.; Wang, J.; Yan, X. Artificial nociceptor based on TiO2 nanosheet memristor. Sci. China Mater. 2021, 64, 1703–1712. [Google Scholar] [CrossRef]
  30. Yang, Y.C.; Pan, F.; Liu, Q.; Liu, M.; Zeng, F. Fully Room-Temperature-Fabricated Nonvolatile Resistive Memory for Ultrafast and High-Density Memory Application. Nano Lett. 2009, 9, 1636–1643. [Google Scholar] [CrossRef] [PubMed]
  31. Das, M.; Kumar, A.; Singh, R.; Htay, M.T.; Mukherjee, S. Realization of synaptic learning and memory functions in Y2O3 based memristive device fabricated by dual ion beam sputtering. Nanotechnology 2018, 29, 055203. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, W.; Zhang, B.; Zhao, H. Forming-free bipolar and unipolar resistive switching behaviors with low operating voltage in Ag/Ti/CeO2/Pt devices. Results Phys. 2020, 16, 103001. [Google Scholar] [CrossRef]
  33. Rana, A.M.; Akbar, T.; Ismail, M.; Ahmad, E.; Hussain, F.; Talib, I.; Imran, M.; Mehmood, K.; Iqbal, K.; Nadeem, M.Y. Endurance and Cycle-to-cycle Uniformity Improvement in Tri-Layered CeO2/Ti/CeO2 Resistive Switching Devices by Changing Top Electrode Material. Sci. Rep. 2017, 7, 39539. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, L.N.; Zang, C.H.; Wang, B.; Su, W.; Xiao, H.Y.; Zhang, D.M.; Zhang, Y.S. Ceria thin film memristive device by magnetron sputtering method. Vacuum 2020, 173, 109128. [Google Scholar] [CrossRef]
  35. Li, X.; Zhang, L.; Guo, R.; Chen, J.; Yan, X. A Flexible Transient Biomemristor Based on Hybrid Structure HfO2/BSA:Au Double Layers. Adv. Mater. Technol. 2020, 5, 2000191. [Google Scholar] [CrossRef]
  36. Liu, Q.; Sun, J.; Lv, H.; Long, S.; Yin, K.; Wan, N.; Li, Y.; Sun, L.; Liu, M. Real-Time Observation on Dynamic Growth/Dissolution of Conductive Filaments in Oxide-Electrolyte-Based ReRAM. Adv. Mater. 2012, 24, 1844–1849. [Google Scholar] [CrossRef] [PubMed]
  37. Zheng, H.; Kim, H.J.; Yang, P.; Park, J.-S.; Kim, D.W.; Lee, H.H.; Kang, C.J.; Yoon, T.-S. Forming-free resistive switching characteristics of Ag/CeO2/Pt devices with a large memory window. Semicond. Sci. Technol. 2017, 32, 055006. [Google Scholar] [CrossRef]
  38. Li, Y.; Long, S.; Liu, Q.; Lv, H.; Liu, M. Resistive Switching Performance Improvement via Modulating Nanoscale Conductive Filament, Involving the Application of Two-Dimensional Layered Materials. Small 2017, 13, 1604306. [Google Scholar] [CrossRef] [PubMed]
  39. Li, H.; Liu, T.; Wang, Y.; Geng, S.; Xu, T.; Cao, M.; Fan, S.; Liu, T.; Su, J. Electronic synaptic characteristics and simulation application of Ag/CeO2/Pt memristor. Ceram. Int. 2022, 48, 13754–13760. [Google Scholar] [CrossRef]
  40. Dearnaley, G.; Stoneham, A.M.; Morgan, D.V. Electrical phenomena in amorphous oxide films. Rep. Prog. Phys. 1970, 33, 1129–1191. [Google Scholar] [CrossRef]
  41. Sawa, A. Resistive switching in transition metal oxides. Mater. Today 2008, 11, 28–36. [Google Scholar] [CrossRef]
  42. Akinaga, H.; Shima, H. Resistive Random Access Memory (ReRAM) Based on Metal Oxides. Proc. IEEE 2010, 98, 2237–2251. [Google Scholar] [CrossRef]
  43. Jiang, Z.; Wu, Y.; Yu, S.; Yang, L.; Song, K.; Karim, Z.; Wong, H.S.P. A Compact Model for Metal–Oxide Resistive Random Access Memory With Experiment Verification. IEEE Trans. Electron Devices 2016, 63, 1884–1892. [Google Scholar] [CrossRef]
  44. Sun, Z.G.; Mizuguchi, M.; Manago, T.; Akinaga, H. Magnetic-field-controllable avalanche breakdown and giant magnetoresistive effects in Gold∕semi-insulating-GaAs Schottky diode. Appl. Phys. Lett. 2004, 85, 5643–5645. [Google Scholar] [CrossRef]
  45. Wagenaar, J.J.T.; Morales-Masis, M.; van Ruitenbeek, J.M. Observing “quantized” conductance steps in silver sulfide: Two parallel resistive switching mechanisms. J. Appl. Phys. 2012, 111, 014302. [Google Scholar] [CrossRef]
  46. Oligschlaeger, R.; Waser, R.; Meyer, R.; Karthäuser, S.; Dittmann, R. Resistive switching and data reliability of epitaxial (Ba,Sr)TiO3 thin films. Appl. Phys. Lett. 2006, 88, 042901. [Google Scholar] [CrossRef]
  47. Wang, H.; Yan, X. Overview of Resistive Random Access Memory (RRAM): Materials, Filament Mechanisms, Performance Optimization, and Prospects. Phys. Status Solidi (RRL)—Rapid Res. Lett. 2019, 13, 1900073. [Google Scholar] [CrossRef]
  48. Nagarajan, L.; De Souza, R.A.; Samuelis, D.; Valov, I.; Börger, A.; Janek, J.; Becker, K.-D.; Schmidt, P.C.; Martin, M. A chemically driven insulator–metal transition in non-stoichiometric and amorphous gallium oxide. Nat. Mater. 2008, 7, 391–398. [Google Scholar] [CrossRef] [PubMed]
  49. Hickmott, T.W. Low-Frequency Negative Resistance in Thin Anodic Oxide Films. J. Appl. Phys. 1962, 33, 2669–2682. [Google Scholar] [CrossRef]
  50. Zhang, P.; Gao, C.; Xu, B.; Qi, L.; Jiang, C.; Gao, M.; Xue, D. Structural Phase Transition Effect on Resistive Switching Behavior of MoS2-Polyvinylpyrrolidone Nanocomposites Films for Flexible Memory Devices. Small 2016, 12, 2077–2084. [Google Scholar] [CrossRef] [PubMed]
  51. Jung, K.; Seo, H.; Kim, Y.; Im, H.; Hong, J.; Park, J.-W.; Lee, J.-K. Temperature dependence of high- and low-resistance bistable states in polycrystalline NiO films. Appl. Phys. Lett. 2007, 90, 052104. [Google Scholar] [CrossRef]
  52. Yang, Y.; Gao, P.; Gaba, S.; Chang, T.; Pan, X.; Lu, W. Observation of conducting filament growth in nanoscale resistive memories. Nat. Commun. 2012, 3, 732. [Google Scholar] [CrossRef]
  53. Tsuruoka, T.; Hasegawa, T.; Terabe, K.; Aono, M. Conductance quantization and synaptic behavior in a Ta2O5-based atomic switch. Nanotechnology 2012, 23, 435705. [Google Scholar] [CrossRef]
  54. Muenstermann, R.; Menke, T.; Dittmann, R.; Waser, R. Coexistence of Filamentary and Homogeneous Resistive Switching in Fe-Doped SrTiO3 Thin-Film Memristive Devices. Adv. Mater. 2010, 22, 4819–4822. [Google Scholar] [CrossRef]
  55. Park, G.-S.; Kim, Y.B.; Park, S.Y.; Li, X.S.; Heo, S.; Lee, M.-J.; Chang, M.; Kwon, J.H.; Kim, M.; Chung, U.I.; et al. In situ observation of filamentary conducting channels in an asymmetric Ta2O5−x/TaO2−x bilayer structure. Nat. Commun. 2013, 4, 2382. [Google Scholar] [CrossRef]
  56. Lee, S.; Sohn, J.; Jiang, Z.; Chen, H.-Y.; Philip Wong, H.S. Metal oxide-resistive memory using graphene-edge electrodes. Nat. Commun. 2015, 6, 8407. [Google Scholar] [CrossRef]
  57. Lee, J.; Yoo, B.; Lee, H.; Cha, G.D.; Lee, H.-S.; Cho, Y.; Kim, S.Y.; Seo, H.; Lee, W.; Son, D.; et al. Ultra-Wideband Multi-Dye-Sensitized Upconverting Nanoparticles for Information Security Application. Adv. Mater. 2017, 29, 1603169. [Google Scholar] [CrossRef]
  58. Munjal, S.; Khare, N. Valence Change Bipolar Resistive Switching Accompanied With Magnetization Switching in CoFe2O4 Thin Film. Sci. Rep. 2017, 7, 12427. [Google Scholar] [CrossRef]
  59. Ielmini, D.; Wong, H.S.P. In-memory computing with resistive switching devices. Nat. Electron. 2018, 1, 333–343. [Google Scholar] [CrossRef]
  60. Lastras-Montaño, M.A.; Cheng, K.-T. Resistive random-access memory based on ratioed memristors. Nat. Electron. 2018, 1, 466–472. [Google Scholar] [CrossRef]
  61. Terabe, K.; Hasegawa, T.; Nakayama, T.; Aono, M. Quantized conductance atomic switch. Nature 2005, 433, 47–50. [Google Scholar] [CrossRef] [PubMed]
  62. Tian, H.; Zhao, L.; Wang, X.; Yeh, Y.-W.; Yao, N.; Rand, B.P.; Ren, T.-L. Extremely Low Operating Current Resistive Memory Based on Exfoliated 2D Perovskite Single Crystals for Neuromorphic Computing. ACS Nano 2017, 11, 12247–12256. [Google Scholar] [CrossRef]
  63. Younis, A.; Chu, D.; Li, S. Evidence of Filamentary Switching in Oxide-based Memory Devices via Weak Programming and Retention Failure Analysis. Sci. Rep. 2015, 5, 13599. [Google Scholar] [CrossRef]
  64. Kumar, M.; Kim, H.-S.; Park, D.Y.; Jeong, M.S.; Kim, J. Compliance-Free Multileveled Resistive Switching in a Transparent 2D Perovskite for Neuromorphic Computing. ACS Appl. Mater. Interfaces 2018, 10, 12768–12772. [Google Scholar] [CrossRef] [PubMed]
  65. Gao, S.; Zeng, F.; Chen, C.; Tang, G.; Lin, Y.; Zheng, Z.; Song, C.; Pan, F. Conductance quantization in a Ag filament-based polymer resistive memory. Nanotechnology 2013, 24, 335201. [Google Scholar] [CrossRef]
  66. Cheng, X.-F.; Hou, X.; Zhou, J.; Gao, B.-J.; He, J.-H.; Li, H.; Xu, Q.-F.; Li, N.-J.; Chen, D.-Y.; Lu, J.-M. Pseudohalide-Induced 2D (CH3NH3)2PbI2(SCN)2 Perovskite for Ternary Resistive Memory with High Performance. Small 2018, 14, 1703667. [Google Scholar] [CrossRef]
  67. Lim, Z.X.; Cheong, K.Y. Nonvolatile Memory Device Based on Bipolar and Unipolar Resistive Switching in Bio-Organic Aloe Polysaccharides Thin Film. Adv. Mater. Technol. 2018, 3, 1800007. [Google Scholar] [CrossRef]
  68. Pfleger, J. Electronic Memory Devices Based on Solution-Processable Nanostructured Materials. In Solution-Processable Components for Organic Electronic Devices; Willey Online Library: Hoboken, NJ, USA, 2019; pp. 591–626. [Google Scholar] [CrossRef]
  69. Mohammad, B.; Jaoude, M.A.; Kumar, V.; Homouz, D.M.A.; Nahla, H.A.; Al-Qutayri, M.; Christoforou, N. State of the art of metal oxide memristor devices. Nanotechnol. Rev. 2016, 5, 311–329. [Google Scholar] [CrossRef]
  70. Younis, A.; Chu, D.; Li, S. Oxygen level: The dominant of resistive switching characteristics in cerium oxide thin films. J. Phys. D Appl. Phys. 2012, 45, 355101. [Google Scholar] [CrossRef]
  71. Younis, A.; Chu, D.; Mihail, I.; Li, S. Interface-Engineered Resistive Switching: CeO2 Nanocubes as High-Performance Memory Cells. ACS Appl. Mater. Interfaces 2013, 5, 9429–9434. [Google Scholar] [CrossRef] [PubMed]
  72. Sun, B.; Liang, D.; Li, X.; Chen, P. Nonvolatile bio-memristor fabricated with natural bio-materials from spider silk. J. Mater. Sci. Mater. Electron. 2016, 27, 3957–3962. [Google Scholar] [CrossRef]
  73. Michaelson, H.B. The work function of the elements and its periodicity. J. Appl. Phys. 1977, 48, 4729–4733. [Google Scholar] [CrossRef]
  74. Hussain, F.; Imran, M.; Cai, Y.Q.; Ullah, H.; Shakoor, A.; Rashid, M.; Ahmad, E.; Ahmad, S.A. Electronic properties of two-dimensional ZnO atomically sheet on Cu substrate: A first-principles study. Mod. Phys. Lett. B 2014, 28, 1450204. [Google Scholar] [CrossRef]
  75. Firdos, M.; Hussain, F.; Imran, M.; Ismail, M.; Rana, A.M.; Arshad Javid, M.; Majid, A.; Arif Khalil, R.M.; Ullah, H. Ab initio study of ceria films for resistive switching memory applications. Mater. Res. Express 2017, 4, 106301. [Google Scholar] [CrossRef]
  76. Ismail, M.; Rana, A.M.; Nisa, S.U.; Hussain, F.; Imran, M.; Mahmood, K.; Talib, I.; Ahmed, E.; Bao, D.H. Effect of annealing treatment on the uniformity of CeO2/TiO2 bilayer resistive switching memory devices. Curr. Appl. Phys. 2017, 17, 1303–1309. [Google Scholar] [CrossRef]
  77. Ismail, M.; Ullah, R.; Hussain, R.; Talib, I.; Rana, A.M.; Hussain, M.; Mahmood, K.; Hussain, F.; Ahmed, E.; Bao, D. Influence of argon and oxygen pressure ratio on bipolar-resistive switching characteristics of CeO2−x thin films deposited at room temperature. Appl. Phys. A 2018, 124, 89. [Google Scholar] [CrossRef]
  78. Ismail, M.; Ahmad, A.; Mahmood, K.; Akbar, T.; Rana, A.M.; Lee, J.; Kim, S. Room temperature deposited oxygen-deficient CeO2−x layer for multilevel resistive switching memory. Appl. Surf. Sci. 2019, 483, 803–810. [Google Scholar] [CrossRef]
  79. Sahu, D.P.; Park, K.; Han, J.; Yoon, T.-S. Improvement of forming-free threshold switching reliability of CeO2-based selector device by controlling volatile filament formation behaviors. APL Mater. 2022, 10, 051111. [Google Scholar] [CrossRef]
  80. Ismail, M.; Rana, A.M.; Talib, I.; Tsai, T.-L.; Chand, U.; Ahmed, E.; Nadeem, M.Y.; Aziz, A.; Shah, N.A.; Hussain, M. Room-temperature fabricated, fully transparent resistive memory based on ITO/CeO2/ITO structure for RRAM applications. Solid State Commun. 2015, 202, 28–34. [Google Scholar] [CrossRef]
  81. Wang, B.; Zang, C.H. Negative differential resistance effect induced by bulk and interface defect states in Pt/CeO2-x/Pt resistive memory devices. Vacuum 2022, 202, 111171. [Google Scholar] [CrossRef]
  82. Zhou, G.; Ren, Z.; Wang, L.; Sun, B.; Duan, S.; Song, Q. Artificial and wearable albumen protein memristor arrays with integrated memory logic gate functionality. Mater. Horiz. 2019, 6, 1877–1882. [Google Scholar] [CrossRef]
  83. Xie, S.; Pei, L.; Li, M.; Zhu, Y.; Cheng, X.; Ding, H.; Xiong, R. Light-controlled resistive switching and voltage-controlled photoresponse characteristics in the Pt/CeO2/Nb:SrTiO3 heterostructure. J. Alloys Compd. 2019, 778, 141–147. [Google Scholar] [CrossRef]
  84. Jabeen, S.; Ismail, M.; Rana, A.M.; Ahmed, E. Impact of work function on the resistive switching characteristics of M/ZnO/CeO2/Pt devices. Mater. Res. Express 2017, 4, 056401. [Google Scholar] [CrossRef]
  85. Kakushima, K.; Okamoto, K.; Koyanagi, T.; Kouda, M.; Tachi, K.; Kawanago, T.; Song, J.; Ahmet, P.; Tsutsui, K.; Sugii, N.; et al. Selection of rare earth silicates for highly scaled gate dielectrics. Microelectron. Eng. 2010, 87, 1868–1871. [Google Scholar] [CrossRef]
  86. Shubhakar, K.; Pey, K.L.; Kushvaha, S.S.; O’Shea, S.J.; Raghavan, N.; Bosman, M.; Kouda, M.; Kakushima, K.; Iwai, H. Grain boundary assisted degradation and breakdown study in cerium oxide gate dielectric using scanning tunneling microscopy. Appl. Phys. Lett. 2011, 98, 072902. [Google Scholar] [CrossRef]
  87. Dou, C.; Kakushima, K.; Ahmet, P.; Tsutsui, K.; Nishiyama, A.; Sugii, N.; Natori, K.; Hattori, T.; Iwai, H. Resistive switching behavior of a CeO2 based ReRAM cell incorporated with Si buffer layer. Microelectron. Reliab. 2012, 52, 688–691. [Google Scholar] [CrossRef]
  88. Velichko, A.; Boriskov, P.; Savenko, A.; Grishin, A.; Khartsev, S.; Yar, M.A.; Muhammed, M. Memory resistive switching in CeO2-based film microstructures patterned by a focused ion beam. Thin Solid Film. 2014, 556, 520–524. [Google Scholar] [CrossRef]
  89. Hu, Q.; Kang, T.S.; Abbas, H.; Lee, T.S.; Lee, N.J.; Park, M.R.; Yoon, T.-S.; Kang, C.J. Resistive switching characteristics of Ag/MnO/CeO2/Pt heterostructures memory devices. Microelectron. Eng. 2018, 189, 28–32. [Google Scholar] [CrossRef]
  90. Li, J.-C.; Cui, H.-P.; Hou, X.-Y. Effect of p-n interface on resistive switching of NiO/CeO2 thin films. J. Alloys Compd. 2018, 752, 247–252. [Google Scholar] [CrossRef]
  91. Ortega-Hernandez, R.; Coll, M.; Gonzalez-Rosillo, J.; Palau, A.; Obradors, X.; Miranda, E.; Puig, T.; Suñe, J. Resistive switching in CeO2/La0.8Sr0.2MnO3 bilayer for non-volatile memory applications. Microelectron. Eng. 2015, 147, 37–40. [Google Scholar] [CrossRef]
  92. Ismail, M.; Jabeen, S.; Akber, T.; Talib, I.; Hussain, F.; Rana, A.M.; Hussain, M.; Mahmood, K.; Ahmed, E.; Bao, D. Elimination of endurance degradation by oxygen annealing in bilayer ZnO/CeO2-x thin films for nonvolatile resistive memory. Curr. Appl. Phys. 2018, 18, 924–932. [Google Scholar] [CrossRef]
  93. Yao, G.; Qu, Z.; Li, C.; Peng, Y.; Li, Q.; Zeng, Z.; Dong, J.; Zhang, B. Improvement of the Parameter Distribution of TiN/HfO2/CeOx/TiN Memristors by Rapid Thermal Annealing. J. Nanomater. 2022, 2022, 6522007. [Google Scholar] [CrossRef]
  94. Ismail, M.; Ahmed, E.; Rana, A.M.; Hussain, F.; Talib, I.; Nadeem, M.Y.; Panda, D.; Shah, N.A. Improved Endurance and Resistive Switching Stability in Ceria Thin Films Due to Charge Transfer Ability of Al Dopant. ACS Appl. Mater. Interfaces 2016, 8, 6127–6136. [Google Scholar] [CrossRef] [PubMed]
  95. Tradel, S.; Li, G.; Zhang, X.; Hill, R.H. Positive and Negative Lithography by Photochemical Metalorganic Deposition from Metal 2-ethylhexanoates. J. Photopolym. Sci. Technol. 2006, 19, 467–475. [Google Scholar] [CrossRef]
  96. Buono-Core, G.E.; Cabello, G.; Klahn, A.H.; Lucero, A.; Nuñez, M.V.; Torrejón, B.; Castillo, C. Growth and characterization of molybdenum oxide thin films prepared by photochemical metal–organic deposition (PMOD). Polyhedron 2010, 29, 1551–1554. [Google Scholar] [CrossRef]
  97. Kim, S.-M.; Moon, H.G.; Lee, H.-S. Resistive switching characteristics of directly patterned Y-doped CeO2 by photochemical organic-metal deposition. Ceram. Int. 2020, 46, 22831–22836. [Google Scholar] [CrossRef]
  98. Rehman, S.; Kim, H.; Farooq Khan, M.; Hur, J.-H.; Lee, A.D.; Kim, D.-K. Tuning of ionic mobility to improve the resistive switching behavior of Zn-doped CeO2. Sci. Rep. 2019, 9, 19387. [Google Scholar] [CrossRef] [PubMed]
  99. Be Lan, T.T.; Li, Y.-T.; Aidan Sun, A.-C.; Lu, H.-C.; Wang, S.-F. Chemical defect-dependent resistive switching characterization in CeO2 thin films. Mater. Sci. Semicond. Process. 2022, 137, 106177. [Google Scholar] [CrossRef]
  100. Caravelli, F.; Carbajal, J.P. Memristors for the curious outsiders. Technologies 2018, 6, 118. [Google Scholar] [CrossRef]
  101. Hsieh, C.-C.; Roy, A.; Chang, Y.-F.; Shahrjerdi, D.; Banerjee, S.K. A sub-1-volt analog metal oxide memristive-based synaptic device with large conductance change for energy-efficient spike-based computing systems. Appl. Phys. Lett. 2016, 109, 223501. [Google Scholar] [CrossRef]
  102. Chang, T.; Jo, S.-H.; Lu, W. Short-Term Memory to Long-Term Memory Transition in a Nanoscale Memristor. ACS Nano 2011, 5, 7669–7676. [Google Scholar] [CrossRef]
  103. Moon, S.; Park, K.; Chung, P.H.; Sahu, D.P.; Yoon, T.S. Tunable voltage polarity-dependent resistive switching characteristics by interface energy barrier modulation in ceria-based bilayer memristors for neuromorphic computing. J. Alloys Compd. 2023, 963, 171211. [Google Scholar] [CrossRef]
  104. Wang, W.; Zhao, H.; Zhang, B.; Tu, H. Forming-Free Pt/Ti/AlOx/CeOx/Pt Multilayer Memristors with Multistate and Synaptic Characteristics. J. Nanomater. 2022, 2022, 1370919. [Google Scholar] [CrossRef]
  105. Kuncic, Z.; Nakayama, T. Neuromorphic nanowire networks: Principles, progress and future prospects for neuro-inspired information processing. Adv. Phys. X 2021, 6, 1894234. [Google Scholar] [CrossRef]
Figure 2. Schematic illustration of (a) original state containing electrodes as top Al and bottom FTO, (b) electroforming mechanism, (c) LRS-containing conducting cobalt ferrite film and oxygen diffusion layer toward the Al electrode, (d) reset state showing O−2 ions returning and originating from oxygen spreading layer toward the cobalt ferrite film, (e,f) HRS with the set process. Adopted from [58]. Copyright 2017 Nature Publishing Group.
Figure 2. Schematic illustration of (a) original state containing electrodes as top Al and bottom FTO, (b) electroforming mechanism, (c) LRS-containing conducting cobalt ferrite film and oxygen diffusion layer toward the Al electrode, (d) reset state showing O−2 ions returning and originating from oxygen spreading layer toward the cobalt ferrite film, (e,f) HRS with the set process. Adopted from [58]. Copyright 2017 Nature Publishing Group.
Nanomaterials 13 02443 g002
Figure 4. Schematic of (I–V) switching performance of memristors for (a) unipolar type, (b) bipolar type, and for compliance current (CC). Adopted from [69]. Copyright 2016 Walter de Gruyter GmbH.
Figure 4. Schematic of (I–V) switching performance of memristors for (a) unipolar type, (b) bipolar type, and for compliance current (CC). Adopted from [69]. Copyright 2016 Walter de Gruyter GmbH.
Nanomaterials 13 02443 g004
Figure 5. (a) Typical semilogarithmic I–V curves of the Au/CeO2/ITO device. Adopted from [70]. Copyright 2012 IOP Publishing. (b) Typical DC sweep I–V characteristics of the Al/CeO2/Au cell. The order of the arrows discloses the sweeping path with numbers. The value of 1 μm indicates the diameter of each device lying between Au (30 nm), CeOx (13 nm), and Au (30 nm). Adopted from [15]. Copyright 2015 American Institute of Physics. Typical bipolar-mode I–V curves indicating the RS trend for the Zr/CeOx/Pt structure containing CeOx layers with different thicknesses: (c) 25 and (d) 14 nm. Adopted from [16], Copyright 2014 Springer Nature. (e) Typical I–V graphs for the Pt/CeO2/Pt device at the semilogarithmic scale and inset showing the scanning electron microscopy (SEM) image of the CeO2/Pt/Ti/SiO2 construction. Adopted from [17]. Copyright 2008 Elsevier B.V. (f) Semilogarithmic I–V curves under voltage sweep. Adopted from [18]. Copyright 2017 IOP Publishing. (g) Schematic depiction of self-assembled CeO2 nano cubes for the functioning of the RS device with the I–V measurement scheme, (i) temperature dependence of ROFF and RON for the device to indicate semiconducting performance. The inset shows the OFF-state Arrhenius plot, (ii) Change in carrier density under an electric field, and simulation of OFF/ON ratio under variation in film thickness. Adopted from [71]. Copyright 2013 American Chemical Society.
Figure 5. (a) Typical semilogarithmic I–V curves of the Au/CeO2/ITO device. Adopted from [70]. Copyright 2012 IOP Publishing. (b) Typical DC sweep I–V characteristics of the Al/CeO2/Au cell. The order of the arrows discloses the sweeping path with numbers. The value of 1 μm indicates the diameter of each device lying between Au (30 nm), CeOx (13 nm), and Au (30 nm). Adopted from [15]. Copyright 2015 American Institute of Physics. Typical bipolar-mode I–V curves indicating the RS trend for the Zr/CeOx/Pt structure containing CeOx layers with different thicknesses: (c) 25 and (d) 14 nm. Adopted from [16], Copyright 2014 Springer Nature. (e) Typical I–V graphs for the Pt/CeO2/Pt device at the semilogarithmic scale and inset showing the scanning electron microscopy (SEM) image of the CeO2/Pt/Ti/SiO2 construction. Adopted from [17]. Copyright 2008 Elsevier B.V. (f) Semilogarithmic I–V curves under voltage sweep. Adopted from [18]. Copyright 2017 IOP Publishing. (g) Schematic depiction of self-assembled CeO2 nano cubes for the functioning of the RS device with the I–V measurement scheme, (i) temperature dependence of ROFF and RON for the device to indicate semiconducting performance. The inset shows the OFF-state Arrhenius plot, (ii) Change in carrier density under an electric field, and simulation of OFF/ON ratio under variation in film thickness. Adopted from [71]. Copyright 2013 American Chemical Society.
Nanomaterials 13 02443 g005
Figure 7. (a) Schematic depiction of the Ti/CeO2−x/Pt device, indicating the addition of an insulating layer between two electrodes. Arrows display the I–V measurements. Typical I–V curves show the initial electroforming process with initial sweeping for a bipolar RS trend in Ti/CeO2−x/Pt cells containing different Ar/O2 amounts for (b) specimen A, (c) specimen B, (d) specimen C, and (e) forming voltage with SET/RESET processes involving the Ar/O2 ratio in the CeO2 layer depositing chamber. Adopted from [77]. Copyright 2018 Springer Nature. (f) Schematic illustration of the Ni/CeO2/ITO/glass structures, (g) I–V curves for the Ni/CeO2−x/ITO/glass device measured at different RESET-stop voltages ranging from −1.5 to −2.0 V, and (h) statistical distribution of the SET and RESET voltages. Adopted from [78]. Copyright 2019 Elsevier B.V.
Figure 7. (a) Schematic depiction of the Ti/CeO2−x/Pt device, indicating the addition of an insulating layer between two electrodes. Arrows display the I–V measurements. Typical I–V curves show the initial electroforming process with initial sweeping for a bipolar RS trend in Ti/CeO2−x/Pt cells containing different Ar/O2 amounts for (b) specimen A, (c) specimen B, (d) specimen C, and (e) forming voltage with SET/RESET processes involving the Ar/O2 ratio in the CeO2 layer depositing chamber. Adopted from [77]. Copyright 2018 Springer Nature. (f) Schematic illustration of the Ni/CeO2/ITO/glass structures, (g) I–V curves for the Ni/CeO2−x/ITO/glass device measured at different RESET-stop voltages ranging from −1.5 to −2.0 V, and (h) statistical distribution of the SET and RESET voltages. Adopted from [78]. Copyright 2019 Elsevier B.V.
Nanomaterials 13 02443 g007
Figure 8. I–V characteristics displaying threshold behavior for (a) Ag/CeO2/Pt cells and (b) Ag/CeO2/SiO2/Pt cells. Endurance performance of (c) Ag/CeO2/Pt device and (d) Ag/CeO2/SiO2/Pt device under pulse operation mode. Endurance of the Ag/CeO2/SiO2/Pt device, (e) threshold switching behavior of the Ag/CeO2/SiO2/Pt device with variable CC, and (f) increment in ON-state resistance along with CC. Adapted from [79] Copyright 2022 American Institute of Physics.
Figure 8. I–V characteristics displaying threshold behavior for (a) Ag/CeO2/Pt cells and (b) Ag/CeO2/SiO2/Pt cells. Endurance performance of (c) Ag/CeO2/Pt device and (d) Ag/CeO2/SiO2/Pt device under pulse operation mode. Endurance of the Ag/CeO2/SiO2/Pt device, (e) threshold switching behavior of the Ag/CeO2/SiO2/Pt device with variable CC, and (f) increment in ON-state resistance along with CC. Adapted from [79] Copyright 2022 American Institute of Physics.
Nanomaterials 13 02443 g008
Figure 10. Fitting results of I–V curves at three levels: (a) prior to forming, (b) transition in resistance states, and (c) at LRS. Adopted from [39]. Copyright Elsevier B.V. (d) Ce 3d level XPS spectra of the CeO2/NSTO device. (e) Retention/endurance efficiency at LRS, IRS1, IRS2, and during illumination. (f) Isc and Voc of the device during illumination at various applied voltage pulses. Adopted from [83]. Copyright 2019 Elsevier B.V.
Figure 10. Fitting results of I–V curves at three levels: (a) prior to forming, (b) transition in resistance states, and (c) at LRS. Adopted from [39]. Copyright Elsevier B.V. (d) Ce 3d level XPS spectra of the CeO2/NSTO device. (e) Retention/endurance efficiency at LRS, IRS1, IRS2, and during illumination. (f) Isc and Voc of the device during illumination at various applied voltage pulses. Adopted from [83]. Copyright 2019 Elsevier B.V.
Nanomaterials 13 02443 g010
Figure 12. (a) Schematic illustration of Al-doped CeO2 structures, (b) Al-doped device behavior in negative-forming type with left- inset showing negative forming and right inset depicting the scattering in endurance. I–V characterization with endurance performances: (c) undoped and (d) doped devices. Adopted from [94] Copyright 2016 American Chemical Society.
Figure 12. (a) Schematic illustration of Al-doped CeO2 structures, (b) Al-doped device behavior in negative-forming type with left- inset showing negative forming and right inset depicting the scattering in endurance. I–V characterization with endurance performances: (c) undoped and (d) doped devices. Adopted from [94] Copyright 2016 American Chemical Society.
Nanomaterials 13 02443 g012
Figure 13. (a) Application of STDP learning process by utilizing HfOx/CeOx memristors. Adopted from [101]. Copyright 2016 American Institute of Physics. Read current variation with repetition: (b) +10-V pulses of width 20 ms during 100 repetitions, and consecutive −10-V pulses of width 1, 10, and 100 ms during 50 repetitions, (c) pulses with increment in ±V (with fixed width), and (d) fixed ±10-V pulses (with varied width). Adopted from [18]. Copyright 2017 IOP Publishing.
Figure 13. (a) Application of STDP learning process by utilizing HfOx/CeOx memristors. Adopted from [101]. Copyright 2016 American Institute of Physics. Read current variation with repetition: (b) +10-V pulses of width 20 ms during 100 repetitions, and consecutive −10-V pulses of width 1, 10, and 100 ms during 50 repetitions, (c) pulses with increment in ±V (with fixed width), and (d) fixed ±10-V pulses (with varied width). Adopted from [18]. Copyright 2017 IOP Publishing.
Nanomaterials 13 02443 g013
Figure 14. (a) Read current at +2 V) with pulsing at +6 V, width of 50 ms, and time interval of 2 s for 30 cycles; (b) the subsequent normalized memory retention ((ΔI(t)/ΔI(0)) × 100(%)) along with time fitting with stretching relaxation function exponentially; and (c) the read current at +2 V with time in the form of repeated application of voltage pulses with different amplitudes (+5–+11 V), 50-ms width, and 100-s time interval. Adopted from [20]. Copyright 2018 IOP Publishing.
Figure 14. (a) Read current at +2 V) with pulsing at +6 V, width of 50 ms, and time interval of 2 s for 30 cycles; (b) the subsequent normalized memory retention ((ΔI(t)/ΔI(0)) × 100(%)) along with time fitting with stretching relaxation function exponentially; and (c) the read current at +2 V with time in the form of repeated application of voltage pulses with different amplitudes (+5–+11 V), 50-ms width, and 100-s time interval. Adopted from [20]. Copyright 2018 IOP Publishing.
Nanomaterials 13 02443 g014
Figure 15. Schematic display of read current dependence on the decaying function during subsequent potentiation, with PPF, and retention stability in 1000 s for (a) Pt/CeO2/Pt reference device and (b) bilayer Pt/ITO/CeO2/Pt cell. Schematic of read current at +2 V potential for to Pt/ITO/CeO2/Pt on subsequent application of +8 V pulses, along with growing pulse repetitions from 100 to 3000. (c) Semilogarithmic scale and (d) histogram at specific times with a 100-s interval. Adopted from [19]. Copyright 2019 American Institute of Physics.
Figure 15. Schematic display of read current dependence on the decaying function during subsequent potentiation, with PPF, and retention stability in 1000 s for (a) Pt/CeO2/Pt reference device and (b) bilayer Pt/ITO/CeO2/Pt cell. Schematic of read current at +2 V potential for to Pt/ITO/CeO2/Pt on subsequent application of +8 V pulses, along with growing pulse repetitions from 100 to 3000. (c) Semilogarithmic scale and (d) histogram at specific times with a 100-s interval. Adopted from [19]. Copyright 2019 American Institute of Physics.
Nanomaterials 13 02443 g015
Figure 16. Pair of pulse waves with (a) positive signal simulation and (b) negative signal simulation. Fitting consecutive outcomes with (c) positive and (d) negative pulses. Adopted from [39]. Copyright 2022 Elsevier B.V.
Figure 16. Pair of pulse waves with (a) positive signal simulation and (b) negative signal simulation. Fitting consecutive outcomes with (c) positive and (d) negative pulses. Adopted from [39]. Copyright 2022 Elsevier B.V.
Nanomaterials 13 02443 g016
Figure 17. Relationship among three pulse train variables with variation in the resistance state with (a) positive pulse train amplitude, (b) time interval, (c) duration, (d) negative pulse train amplitude, (e) time interval, and (f) with duration. Adopted from [39]. Copyright 2022 Elsevier B.V.
Figure 17. Relationship among three pulse train variables with variation in the resistance state with (a) positive pulse train amplitude, (b) time interval, (c) duration, (d) negative pulse train amplitude, (e) time interval, and (f) with duration. Adopted from [39]. Copyright 2022 Elsevier B.V.
Nanomaterials 13 02443 g017
Figure 18. (a) Schematic of the simulation of multiple learning along with forgetting responses in the natural brain, whereas * indicates maximum learning behavior to be achieved (b) patterning the memory of the synaptic device array corresponding to the memristor’s derivatives, (c) Pavlov’s dog experimental simulation by training the device with consecutive pulse signals. Adopted from [39]. Copyright 2022 Elsevier B.V.
Figure 18. (a) Schematic of the simulation of multiple learning along with forgetting responses in the natural brain, whereas * indicates maximum learning behavior to be achieved (b) patterning the memory of the synaptic device array corresponding to the memristor’s derivatives, (c) Pavlov’s dog experimental simulation by training the device with consecutive pulse signals. Adopted from [39]. Copyright 2022 Elsevier B.V.
Nanomaterials 13 02443 g018
Figure 19. Proposed bilayer memristors containing a stacking structure of CeO2 and GDC layers and their linear, symmetric, and analog synaptic weight update characteristics depending upon their stacking order. The synaptic characteristics/functions in the form of potentiation and depression outcomes for the (a) GDC-top device and (b) CeO2-top device for consecutive potentiation/depression functions [103]. Copyright 2023 Elsevier B.V.
Figure 19. Proposed bilayer memristors containing a stacking structure of CeO2 and GDC layers and their linear, symmetric, and analog synaptic weight update characteristics depending upon their stacking order. The synaptic characteristics/functions in the form of potentiation and depression outcomes for the (a) GDC-top device and (b) CeO2-top device for consecutive potentiation/depression functions [103]. Copyright 2023 Elsevier B.V.
Nanomaterials 13 02443 g019
Figure 20. Demonstration of (a) the STDP realization schemes with pulse amplitude modulation. The pulse amplitudes for the pre-spike reveal the values −0.5, 0.45, 0.43, 0.41, 0.39, 0.37, 0.35, 0.33, 0.31, and 0.29 V. For the post-spike, the values are −0.25, 0.45, 0.4, 0.35, 0.3, 0.25, 0.2, 0.15, 0.1, and 0.05 V. The pulse width of the pulses is 50 ms. The pulse interval between consecutive pulses is 150 ms. (b) The application of STDP learning for the Pt/Ti/AlOx/CeOx/Pt memristor synapse [104].
Figure 20. Demonstration of (a) the STDP realization schemes with pulse amplitude modulation. The pulse amplitudes for the pre-spike reveal the values −0.5, 0.45, 0.43, 0.41, 0.39, 0.37, 0.35, 0.33, 0.31, and 0.29 V. For the post-spike, the values are −0.25, 0.45, 0.4, 0.35, 0.3, 0.25, 0.2, 0.15, 0.1, and 0.05 V. The pulse width of the pulses is 50 ms. The pulse interval between consecutive pulses is 150 ms. (b) The application of STDP learning for the Pt/Ti/AlOx/CeOx/Pt memristor synapse [104].
Nanomaterials 13 02443 g020
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ali, S.; Ullah, M.A.; Raza, A.; Iqbal, M.W.; Khan, M.F.; Rasheed, M.; Ismail, M.; Kim, S. Recent Advances in Cerium Oxide-Based Memristors for Neuromorphic Computing. Nanomaterials 2023, 13, 2443. https://doi.org/10.3390/nano13172443

AMA Style

Ali S, Ullah MA, Raza A, Iqbal MW, Khan MF, Rasheed M, Ismail M, Kim S. Recent Advances in Cerium Oxide-Based Memristors for Neuromorphic Computing. Nanomaterials. 2023; 13(17):2443. https://doi.org/10.3390/nano13172443

Chicago/Turabian Style

Ali, Sarfraz, Muhammad Abaid Ullah, Ali Raza, Muhammad Waqas Iqbal, Muhammad Farooq Khan, Maria Rasheed, Muhammad Ismail, and Sungjun Kim. 2023. "Recent Advances in Cerium Oxide-Based Memristors for Neuromorphic Computing" Nanomaterials 13, no. 17: 2443. https://doi.org/10.3390/nano13172443

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop