Next Article in Journal
A Study on the Interfacial Reactions between Gallium and Cu/Ni/Au(Pd) Multilayer Metallization
Next Article in Special Issue
Amorphous BN-Based Synaptic Device with High Performance in Neuromorphic Computing
Previous Article in Journal
ESD Research of SCR Devices under Harsh Environments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Double-Forming Mechanism of TaOx-Based Resistive Memory Device and Its Synaptic Applications

Division of Electronics and Electrical Engineering, Dongguk University, Seoul 04620, Republic of Korea
*
Author to whom correspondence should be addressed.
Materials 2023, 16(18), 6184; https://doi.org/10.3390/ma16186184
Submission received: 3 August 2023 / Revised: 12 September 2023 / Accepted: 12 September 2023 / Published: 13 September 2023
(This article belongs to the Special Issue Advanced Semiconductor/Memory Materials and Devices)

Abstract

:
The bipolar resistive switching properties of Pt/TaOx/InOx/ITO-resistive random-access memory devices under DC and pulse measurement conditions are explored in this work. Transmission electron microscopy and X-ray photoelectron spectroscopy were used to confirm the structure and chemical compositions of the devices. A unique two-step forming process referred to as the double-forming phenomenon and self-compliance characteristics are demonstrated under a DC sweep. A model based on oxygen vacancy migration is proposed to explain its conduction mechanism. Varying reset voltages and compliance currents were applied to evaluate multilevel cell characteristics. Furthermore, pulses were applied to the devices to demonstrate the neuromorphic system’s application via testing potentiation, depression, spike-timing-dependent plasticity, and spike-rate-dependent plasticity.

1. Introduction

Because processing and memory components are physically separated, the conventional von Neumann architecture employed in computers encounters processing issues. Furthermore, modern CMOS-based electronic devices reach their scaling constraints owing to Moore’s law [1,2,3]. To overcome these problems, bioinspired neuromorphic computing is attracting great attention because of its high efficiency, low power consumption, and parallel data-processing feature [4,5]. The main focus of a neuromorphic computing system is to emulate the human brain’s synapses, in which large amounts of information move from one neuron to another. Many solid-state devices have been researched to mimic this system, and emerging nonvolatile devices are applicable candidates, including ferroelectric random-access memory (RAM) [6,7,8], magnetic RAM [9,10], phase-change RAM [11,12], and resistant RAM (RRAM) [13,14,15,16,17,18]. Among these, RRAM devices have benefits such as simple fabrication, high switching speeds, outstanding scalability, and high endurance, making them one of the most promising choices [19,20,21,22]. Moreover, the simple two-terminal structure of RRAMs, comprising a switching layer sandwiched between the top and bottom electrodes, most closely emulates the structure of a biological synapse [23]. Furthermore, applying different biases with different polarities causes a phenomenon termed the electro-resistance effect, where the resistance condition changes between a low-resistance state (LRS) and a high-resistance state (HRS), and which information is stored at 0 s and 1 s, respectively [24,25]. Various transition metal oxides have been employed as resistive switching insulators, including HfO2 [26], TiO2 [27,28,29], TaOx [30], Al2O3 [31,32], and ZnO [30]. Extensive research has been conducted on TaOx, revealing it as a promising candidate for the resistive switching layer [33]. In addition, previous studies have indicated that TaOx exhibits superior memory characteristics owing to its high endurance (>1010) [34], fast switching speed (<1 ns) [35], and good scalability (<30 nm) [36].
The resistive switching phenomena, which is the basis of RRAM, occurs due to the change in resistance states under an applied bias. For instance, in a valence change memory (VCM) device, the generation and rupture of the conducting filament is a key function in switching resistance [26,27,28]. The applied bias separates oxygen ions (O2−) and oxygen vacancies (Vo+). Then, due to the migration of oxygen ions under the applied electric field, the generated oxygen vacancies create a conductive filament that connects the top and bottom electrodes. Through the filament, a large current flows; thus, the resistance state is switched from a high resistance state (HRS) to a low resistance state (LRS). On the other hand, when the opposite bias is applied, reoxidation occurs in the conductive filament, and the filament is ruptured. Consequently, the filament ruptures and the device returns to HRS. The RRAM device stores memory in these two states, HRS and LRS, which can be consequently reproduced by applying sufficient bias [30]. To increase storage density, research has indicated that by applying multilevel cell (MLC) characteristics, a high storage density could be achieved due to multiple stable states between HRS and LRS. Research has indicated that multilevel cell (MLC) characteristics are evident in resistive switching devices, and they are key features that result in high storage density. This functionality allows devices to save data in the HRS, LRS, and between these two states by simply altering the compliance current (CC) and reset voltage [28,29,37]. Furthermore, the modulation of the memristive device’s conductance is both controllable and incremental, emulating the biological synapse. Here, the strength of the connection between the presynaptic and postsynaptic neurons is incrementally increased or decreased through input spikes by maintaining a history-dependent synaptic weight update [38,39]. Additionally, various synaptic functions can be emulated using pulse responses to assess the application of RRAM as a neuromorphic computing device. These functions include the potentiation and depression of short- and long-term memory (STM and LTM, respectively), spike-rate-dependent plasticity, and spike-time-dependent plasticity (STDP). Controllable conductance and synaptic weight changes can be monitored using these methods [40,41,42], while complex tests (such as pattern recognition systems through handwritten Modified National Institute of Standards and Technology (MNIST) datasets) are often conducted to evaluate the use of memristors as artificial synapses [43,44].
In this work, we studied a Pt/TaOx/InOx/indium tin oxide (ITO) device to investigate its potential for mimicking biological synapses. Bipolar gradual and uniform-resistive change behaviors were achieved with the MLC characteristic. Additionally, during the RF-sputtered deposition of the TaOx layer, the diffusion of oxygen toward the ITO bottom electrode occurred, creating an InOx layer. Due to the additional layer, the device exhibited unique forming behavior (termed double forming [45]) and self-compliance. Uniform switching during cycles (>102) and retention (>104) was also examined with gradual changes in potentiation and depression. The result of potentiation and depression was inserted into PRS using MNIST handwritten figures. Finally, synaptic functions such as long-term potentiation (LTP), long-term depression (LTD), STDP, and SRDP were emulated.

2. Experimental Section

The Pt/TaOx/InOx/ITO RRAM device was prepared using the following procedure. The bottom electrode was a commercially available ITO with a 30 nm thickness on a glass substrate (ITO/glass). Isopropyl alcohol and acetone were used to clean the surface, after which radio frequency (RF) reactive sputtering with a power of 150 W was used to deposit a 5 nm TaOx layer on an ITO/glass substrate. The Ta source target was sputtered at room temperature with Ar (20 sccm) and O2 (6 sccm) at a pressure of 5 mTorr. An oxygen-rich InOx layer of 3 nm thickness was produced by the reactive–sputtering process owing to oxygen migration from TaOx to ITO. Subsequently, a Ti adhesion layer of 1.5 nm thickness was formed by an e-beam evaporator. Finally, the e-beam evaporator was used to deposit Pt with a thickness of 100 nm, a deposition rate of 3 Å/s, and a pressure of 3.7 Torr. The liftoff process was performed to create patterned RRAM cells with a diameter of 100 μm. The electrical properties of Pt/TaOx/InOx/ITO were investigated using a Keithley 4200-SCS semiconductor parameter analyzer in the DC mode and a 4225-PMU ultrafast current–voltage (I–V) pulse module in the pulse mode. Furthermore, bias was applied to the top electrode (Pt), while the bottom electrode (ITO) was grounded at room temperature. The device properties (including cross-section analysis and elemental profiles) were determined using field emission transmission electron microscopy (TEM, JEOL JEMF200,Tokyo, Japan)) and X-ray photoelectron spectroscopy (XPS).

3. Results and Discussion

Figure 1a displays a schematic of the Pt/TaOx/InOx/ITO device. A TEM image was also used to verify the thickness of the device, as displayed in Figure 1b. Pt (100 nm thickness) and ITO (30 nm thickness) were verified using the TEM image. Furthermore, TaOx and InOx layers were observed between the Ti adhesion layer (1 nm thickness) and the ITO bottom electrode. Figure 1c depicts the elemental distribution of each layer, which was validated by an energy-dispersive X-ray spectroscopy line scan.
The chemical properties of the RRAM device are displayed in Figure 2. The insulating TaOx and InOx films were investigated in the XPS depth-profile mode. Figure 2a,b display the XPS spectra of Ta 4f and O 1s, respectively, for the TaOx layer as the first insulator at 4 s. Two peaks of Ta4f7/2 and Ta4f5/2 were located around the binding energies of 22.52 and 25.13 eV, representing the Ta–O bonds. Additionally, the O 1s peak position of bulk TaOx was located around 530.2 eV, indicating the existence of the TaOx thin film. Furthermore, the second insulating layer (InOx) at an etch time of 20 s is depicted in Figure 2c,d. The spectral peaks of In3d5/2 and O 1s were located at approximately 444.81 and 530.5 eV, representing In–O bonding and the existence of an oxygen-distributed InOx layer, respectively. In addition, the oxygen vacancy concentration of each insulating layer was inspected, as displayed in the insets of Figure 2b,d. The peak at 532.2 eV corresponds to the oxygen vacancies in the InOx and TaOx layers. Because of oxygen migration during the RF sputtering process, the percentage of oxygen vacancies in the TaOx layer was 23.21% compared to 45.89% in the InOx layer. Accordingly, fewer oxygen vacancies were stored in the TaOx layer compared to the InOx layer [45].
Figure 3 displays the electrical characteristics of Pt/TaOx/InOx/ITO under DC sweep conditions. In particular, Figure 3a displays the I–V curve, including the double-forming switching phenomenon, which is unlike the conventional RRAM operation. The “forming” process [46] (also known as dielectric soft breakdown) has been reported to occur once under applied bias, transforming the resistance state of the device from its initial state to LRS [47,48]. However, the RRAM device in this work required additional forming processes in the opposite bias to switch the resistance condition to LRS. The double-forming process can be divided into six steps: (1) the first forming process, (2) the medium state, (3) the second forming process, (4) LRS, (5) reset process, and (6) HRS, as displayed in Figure 3a [45]. By applying a voltage bias of −4 V and a CC of 100 μA to protect the device from hard breakdown, the first conducting filament was formed, and the device turned from its initial state into a medium state. Then, upon the application of a set voltage of 3 V and a reset voltage of −3 V, the device switched from the medium state to LRS and from LRS to HRS owing to the rupture and creation of a conduction path. The bipolar set and reset operations did not require CC, demonstrating self-compliance properties that could be implemented because of the ITO electrode [49,50]. Figure 3b depicts the successful resistive switching characteristic for 275 cycles. Under self-compliance conditions, a modest change in the current was detected, and the device maintained its original HRS and LRS without any significant degradation. As demonstrated in Figure 3a, there was an abrupt jump in the current at 2.1 V, similar to the first forming process at –3.8 V. However, a gradual change in current was exhibited in the set process shown in Figure 3b, and its window decreased from 4.55 to 2.21 at a read voltage of 0.3 V. Therefore, we applied the switching phenomena (3) and (4) as the second forming process with a forming voltage of 3 V. Figure 3c displays the endurance of the device with a read voltage of –0.3 V, where the device maintained its HRS and LRS. Additionally, we investigated the data retention capability of the device, as displayed in Figure 3d. These results indicate that the device maintained its HRS and LRS for 104 s without degradation, demonstrating its good nonvolatile properties.
Figure 4 depicts the conduction mechanism of the Pt/TaOx/InOx/ITO device, where the white dot in the figure represents the implementation of lattice oxygen vacancy. Lin et al. reported that the physical size of the conductive filament could be determined by the existence of lattice oxygen vacancy, where the size of the conducting filament in the oxygen-vacancy-deficient layer was narrower than that in an oxygen-vacancy-rich layer [51]. Furthermore, Huang et al. explained the double-forming mechanism of bilayer-structured resistive switching devices using an asymmetric conductive filament [45]. Based on these previous studies, it can be implied that implementing different amounts of existing lattice oxygen vacancies in the TaOx and InOx layers constructs two different sizes of conducting filaments in each insulating layer. Through the connection of two asymmetric filaments, the device switches to LRS, observing the double-forming mechanism. A thick filament was formed in the oxygen-vacancy-rich InOx layer, whereas a relatively narrow filament was formed in the oxygen-vacancy-deficient TaOx layer. When negative voltage was applied to the top electrode, the first forming process occurred. Redox reactions then caused the separation of oxygen vacancies (Vo+) and oxygen ions (O2−). These separated oxygen ions were then repelled away from the top electrode due to an applied bias. Thus, the generated defects (oxygen vacancies) accumulated and formed conductive filaments of different sizes in the InOx and TaOx layers, as depicted in Figure 4b. However, CC limited the thickening of the conduction path in the TaOx layer, implying that the current decreased in step (2) of Figure 3a. Then, under a positive voltage applied to the top electrode, oxygen ions migrated toward the bottom electrode and were repelled back toward the top electrode due to the electric field. Due to the additional generation of oxygen vacancies in this process, a conductive filament in TaOx, with its thickest part formed at the interface of InOx and TaOx could be completed. Consequently, the asymmetric shape of the conductive filament (Figure 4c) could be constructed. In this second forming process, the device finally changed to LRS (as in step (4)), and repeatable resistive switching phenomena could be observed in the device. The reset process occurred when a negative voltage was applied to Pt. Oxygen ions drifted toward the TaOx layer to recombine with the oxygen vacancies. The weakest part of the filament, at the interface of Pt/TaOx, was ruptured due to the recombination of the oxygen ion and vacancy, assisted with local joule heating [52,53]. The schematic of the reset process is shown in Figure 4d, where the rupture of the weakest part of the conductive filament could be observed. On the other hand, when a positive voltage was applied to the top electrode, oxygen ions migrated toward the top electrode due to the applied electric field, leaving the oxygen vacancies. Thus, defects again accumulated, and the reconstruction of conduction paths occurred. Thus, a large current flowed through the filament, altering the device state to LRS.
Further, we investigated the MLC characteristics of the device. For the application of the device as a synaptic device, MLC characteristics are important for implementing the multiple weights of each synapse in an artificial neural network [54,55]. Additionally, it features high storage density and introduces multiple data storage areas. Two types of voltage bias schemes are usually applied to investigate MLC characteristics: (i) controlling the reset voltage and (ii) controlling the CC.
Figure 5 presents the method for controlling the reset voltage. In Figure 5a, the I–V curve of a Pt/TaOx/InOx/ITO device with four reset voltages is displayed, which reveals that adjusting the reset voltage results in multiple HRSs while maintaining the LRS as constant. This effect is caused by the varying rate of ruptured filaments under different reset voltages [56]. As the reset voltage increases, more oxygen ions are repelled from the top electrode, changing the degree of recombination rate. Therefore, the gap between the Pt top electrode and the conducting filament increases, resulting in decreased IHRS [Figure 5c]. Figure 5b depicts the cycling endurance with varying reset voltages. As stated previously, increasing the reset voltages increased HRS, resulting in a lower IHRS, while LRS remained fairly constant.
Another method of obtaining MLC characteristics was controlling CC during the set operation. Figure 6a displays the I–V curve of a Pt/TaOx/InOx/ITO device with six CC settings. Here, CC was adjusted from 200 μA to 1 mA, and the reset voltage was maintained at −2.5 V. Evidently, increasing CC reduced LRS while maintaining a constant HR. This phenomenon was considered to be induced by an increase in CC, which caused an increase in the current flow in the LRS state. Figure 6c presents a filament schematic of the controlled CC. As the CC increased, the width of the conducting filament also increased. Then, more electrons could move through the enlarged conducting path, resulting in a larger ILRS [44,57]. Consequently, ILRS increased, while IHRS remained unchanged. Figure 6b illustrates a 10-cycle endurance test while applying different CCs. Here, the resistance of the LRS decreased while CC increased.
Next, a scheme of 100 consecutive pulses was applied to evaluate the synaptic characteristics of Pt/TaOx/InOx/ITO. The pulse train comprised 50 identical potentiation and depression pulses each. For potentiation, the pulse width and amplitude were 100 μs and 2 V, respectively, compared to 50 μs and −2.7 V for depression. Figure 7a depicts the result of the applied pulse scheme, which demonstrated a linear rise and decay of conductance. Then, to test the synaptic reproducibility, LTP and LTD were explored using 10-cycle potentiation and depression pulse methods [58]. The results are depicted in Figure 7b, where identical conductance changes could be observed. In other words, the conductance levels after each pulse application favorably maintained their states under repetitive operation, proving the applicability of the device to mimic the human brain. Additionally, a pattern recognition test using handwritten digits from an MNIST dataset was conducted to check the further application of the device as a synaptic device. The training was conducted with unclear images, and the gradual and symmetric conductance changes in potentiation and depression resulted in clearer images with higher accuracy [59,60]. As shown in Figure 7c, the deep neural network (DNN) comprised 784 input neurons, 3 hidden layers, and 10 output neurons. Each of these three hidden layers had 128, 64, and 32 neurons, and the backpropagation method was employed to improve accuracy. To determine linearity and accuracy, the potentiation and depression depicted in Figure 7a were converted into an MNIST handwritten number of 28 × 28 pixels and applied to the neural network. The result of the number recognition is illustrated in Figure 7d, where the highest obtained accuracy was 94.21%.
Finally, the Hebbian learning rules of synapses and neurons were tested on a Pt/TaOx/InOx/ITO device to investigate its ability to mimic a biological synapse [61]. Figure 8a illustrates the RRAM device mimicking the human synapse. Here, the top and bottom electrodes mimic pre- and post-spikes, while synaptic information between the neurons varied due to the conducting filament connecting the top and bottom electrodes [62,63]. STDP is composed of two parts of synaptic variance: LTP and LTD. For example, LTP occurs when the pre-spike exceeds the post-spike, while LTD occurs when the post-spike exceeds the pre-spike. These synaptic investigations were conducted by applying the same pulse train to the pre-and post-spikes. A pulse train with a pulse width of 100 μs was applied to Pt/TaOx/InOx/ITO, as illustrated in Figure 8b. The pulse application was conducted with a difference in the time interval, which is termed spike timing difference Δt (Δt = tpre − tpost). During LTP (Δt > 0), a set of positive pulse trains was applied, which induced a decrease in the resistance of the devices. Further, negative pulse trains were applied during LTD (Δt < 0), inducing an increase in the resistance of the devices. The conductance was acquired because the pulse application was input into the formula below to convert it into synaptic weight (ΔW), which represented the spike connection:
Δ W = G f G i G i × 100 % ,
where Gi and Gf represent the conductance of the initial value before and the final value after applying the pulse trains, respectively. Figure 8c presents the experimental data of STDP obtained from the device. For Δt > 0, the weight change increased continuously with decreasing time intervals, and LTP was obtained. By contrast, for Δt < 0, the weight change decreased, and LTD was obtained. This result proves the successful experimental demonstration of the STDP learning rule with synaptic weight changes at different spike times using the proposed Pt/TaOx/InOx/ITO memristor device, favorably mimicking a biological synapse [23]. Additionally, another Hebbian learning rule, SRDP, was tested to obtain the device’s frequency-dependent characteristics [64]. Ten consequent pulses were applied to the Pt/TaOx/InOx/ITO device with the same pulse width and amplitude of 100 μs and 2 V. These pulse intervals varied from 1 μs to 1 ms, as depicted in Figure 8d. The term SRDP Index is calculated as:
S R D P I n d e x = I n I i × 100 %
where In and Ii represent the current of the initial value before and the current after applying consequent pulse trains. The result indicates that when the pulse interval is small, the device response rapidly increases, successfully emulating SRDP behavior.

4. Conclusions

In this study, Pt/TaOx/InOx/ITO exhibited bipolar resistive switching characteristics under double-forming conditions. The device also exhibited acceptable endurance (>102) and retention (>104) properties, with multilevel characteristics, which were demonstrated by varying the reset voltage and CC. Furthermore, pulse trains were applied to reveal their neuromorphic system characteristics. The results of potentiation and depression, MNIST pattern recognition, STDP, and SRDP proved their potential for future applications as a neuromorphic device.

Author Contributions

Software, S.K. (Sunghun Kim); Formal analysis, D.J. and S.L.; Investigation, D.J. and S.K. (Sunghun Kim); Writing—original draft, D.J.; Writing—review & editing, S.K. (Sungjun Kim); Supervision, S.K. (Sungjun Kim); Funding acquisition, S.K. (Sungjun Kim) All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the National Research Foundation of Korea (NRF) grant funded by the Ministry of Science and ICT (2021R1C1C1004422) and by Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korea government (MOTIE) under Grant 20224000000020.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Baek, M.-H.; Kim, H. Polysilicon-Channel Synaptic Transistors for Implementation of Short- and Long-Term Memory Characteristics. Biomimetics 2023, 8, 368. [Google Scholar] [CrossRef] [PubMed]
  2. Li, H.; Gao, B.; Chen, Z.; Zhao, Y.; Huang, P.; Ye, H.; Liu, L.; Liu, X.; Kang, J. A learnable parallel processing architecture towards unity of memory and computing. Sci. Rep. 2015, 5, 13330. [Google Scholar] [CrossRef] [PubMed]
  3. Milo, V.; Malavena, G.; Compagnoni, C.M.; Ielmini, D. Memristive and CMOS devices for neuromorphic computing. Materials 2020, 13, 166. [Google Scholar] [CrossRef] [PubMed]
  4. Woo, J.; Moon, K.; Song, J.; Lee, S.; Kwak, M.; Park, J.; Hwang, H. Improved synaptic behavior under identical pulses using AlOx/HfO2 bilayer RRAM array for neuromorphic systems. IEEE Electron. Device Lett. 2016, 37, 994–997. [Google Scholar] [CrossRef]
  5. Wu, W.; Wu, H.; Gao, B.; Deng, N.; Yu, S.; Qian, H. Improving Analog Switching in HfOx-Based Resistive Memory with a Thermal Enhanced Layer. IEEE Electron. Device Lett. 2017, 38, 1019–1022. [Google Scholar] [CrossRef]
  6. Hiroshi, I. Ferroelectric Random Access Memories. J. Nanosci. Nanotechnol. 2012, 12, 7619–7627. [Google Scholar]
  7. Yoshihiro, A.; Ishiwara, H. Current Status of Ferroelectric Random-Access memory. MRS Bull. 2004, 29, 823–828. [Google Scholar]
  8. Mikolajick, T.; Slesazeck, S.; Park, M.H.; Schroeder, U. Ferroelectric hafnium oxide for ferroelectric random-access memories and ferroelectric field-effect transistors. MRS Bull. 2018, 43, 340–346. [Google Scholar] [CrossRef]
  9. Krzysteczko, P.; Münchenberger, J.; Schäfers, M.; Reiss, G.; Thomas, A. The memristive magnetic tunnel junction as a nanoscopic synapse-neuron system. Adv. Mater. 2012, 24, 762–766. [Google Scholar] [CrossRef]
  10. Vincent, A.F.; Larroque, J.; Locatelli, N.; Romdhane, N.B.; Bichler, O.; Gamrat, C.; Zhao, W.S.; Klein, J.O.; Galdin-Retailleau, S.; Querlioz, D. Spin-transfer torque magnetic memory as a stochastic memristive synapse for neuromorphic systems. IEEE Trans. Biomed. Circuits Syst. 2015, 9, 166–174. [Google Scholar] [CrossRef]
  11. Kuzum, D.; Jeyasingh, R.G.D.; Lee, B.; Wong, H.S.P. Nanoelectronic programmable synapses based on phase change materials for brain-inspired computing. Nano Lett. 2012, 12, 2179–2186. [Google Scholar] [CrossRef]
  12. Burr, G.W.; Shelby, R.M.; Sidler, S.; Di Nolfo, C.; Jang, J.; Boybat, I.; Shenoy, R.S.; Narayanan, P.; Virwani, K.; Giacometti, E.U.; et al. Experimental Demonstration and Tolerancing of a Large-Scale Neural Network (165,000 Synapses) Using Phase-Change Memory as the Synaptic Weight Element. IEEE Trans. Electron Devices 2015, 62, 3498–3507. [Google Scholar] [CrossRef]
  13. Gao, B.; Bi, Y.; Chen, H.Y.; Liu, R.; Huang, P.; Chen, B.; Liu, L.; Liu, X.; Yu, S.; Wong, H.S.P.; et al. Ultra-low-energy three-dimensional oxide-based electronic synapses for implementation of robust high-accuracy neuromorphic computation systems. ACS Nano 2014, 8, 6998–7004. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, I.T.; Chang, C.C.; Chiu, L.W.; Chou, T.; Hou, T.H. 3D Ta/TaOx/TiO2/Ti synaptic array and linearity tuning of weight update for hardware neural network applications. Nanotechnology 2016, 27, 365204. [Google Scholar] [CrossRef]
  15. Ambrogio, S.; Balatti, S.; Milo, V.; Carboni, R.; Wang, Z.Q.; Calderoni, A.; Ramaswamy, N.; Ielmini, D. Neuromorphic Learning and Recognition with One-Transistor-One-Resistor Synapses and Bistable Metal Oxide RRAM. IEEE Trans. Electron Devices 2016, 63, 1508–1515. [Google Scholar] [CrossRef]
  16. Wang, H.; Yan, X. Overview of Resistive Random Access Memory (RRAM): Materials, Filament Mechanisms, Performance Optimization, and Prospects. PhysStatus Solidi-Rapid Res. Lett. 2019, 13, 1900073. [Google Scholar] [CrossRef]
  17. Li, C.; Hu, M.; Li, Y.; Jiang, H.; Ge, N.; Montgomery, E.; Zhang, J.; Song, W.; Dávila, N.; Graves, C.E.; et al. Analogue signal and image processing with large memristor crossbars. Nat. Electron. 2018, 1, 52–59. [Google Scholar] [CrossRef]
  18. Wang, Z.; Joshi, S.; Savel’Ev, S.; Song, W.; Midya, R.; Li, Y.; Rao, M.; Yan, P.; Asapu, S.; Zhuo, Y.; et al. Fully memristive neural networks for pattern classification with unsupervised learning. Nat. Electron. 2018, 1, 137–145. [Google Scholar] [CrossRef]
  19. Park, J.; Kwak, M.; Moon, K.; Woo, J.; Lee, D.; Hwang, H. TiOx-Based RRAM Synapse with 64-Levels of Conductance and Symmetric Conductance Change by Adopting a Hybrid Pulse Scheme for Neuromorphic Computing. IEEE Electron Device Lett. 2016, 37, 1559–1562. [Google Scholar] [CrossRef]
  20. Moon, K.; Kwak, M.; Park, J.; Lee, D.; Hwang, H. Improved Conductance Linearity and Conductance Ratio of 1T2R Synapse Device for Neuromorphic Systems. IEEE Electron Device Lett. 2017, 38, 1023–1026. [Google Scholar] [CrossRef]
  21. Mehonic, A.; Kenyon, A.J. Emulating the electrical activity of the neuron using a silicon oxide RRAM cell. Front. Neurosci. 2016, 10, 57. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, J.; Yang, H.; Ma, Z.; Chen, K.; Zhang, X.; Huang, X.; Oda, S. Characteristics of multilevel storage and switching dynamics in resistive switching cell of Al2O3/HfO2/Al2O3 sandwich structure. J. Phys. D Appl. Phys. 2017, 51, 025102. [Google Scholar] [CrossRef]
  23. Park, J.; Park, H.; Chung, D.; Kim, S. Dynamic and Static Switching in ITO/SnOx/ITO and Its Synaptic Application. Int. J. Mol. Sci. 2022, 23, 9995. [Google Scholar] [CrossRef] [PubMed]
  24. Wei, X.; Huang, H.; Ye, C.; Wei, W.; Zhou, H.; Chen, Y.; Zhang, R.; Zhang, L.; Xia, Q. Exploring the role of nitrogen incorporation in ZrO2 resistive switching film for enhancing the device performance. J. Alloys Compd. 2019, 775, 1301–1306. [Google Scholar] [CrossRef]
  25. Liu, H.C.; Tang, X.G.; Liu, Q.X.; Jiang, Y.P.; Li, W.H.; Guo, X.B.; Tang, Z.H. Bipolar resistive switching behavior and conduction mechanisms of composite nanostructured TiO2/ZrO2 thin film. Ceram. Int. 2020, 46, 21196–21201. [Google Scholar] [CrossRef]
  26. Puglisi, F.M.; Pavan, P.; Padovani, A.; Larcher, L. A study on HfO2 RRAM in HRS based on I-V and RTN analysis. Solid-State Electron. 2014, 102, 69–75. [Google Scholar] [CrossRef]
  27. Kim, K.M.; Park, T.H.; Hwang, C.S. Dual conical conducting filament model in resistance switching TIO2 thin films. Sci. Rep. 2015, 5, 7844. [Google Scholar] [CrossRef] [PubMed]
  28. Cho, H.; Kim, S. Short-Term Memory Dynamics of TiN/Ti/TiO2/SiOx/Si Resistive Random Access Memory. Nanomaterials 2020, 10, 1821. [Google Scholar] [CrossRef]
  29. Sedghi, N.; Li, H.; Brunell, I.F.; Dawson, K.; Potter, R.J.; Guo, Y.; Gibbon, J.T.; Dhanak, V.R.; Zhang, W.D.; Zhang, J.F.; et al. The role of nitrogen doping in ALD Ta2O5 and its influence on multilevel cell switching in RRAM. Appl. Phys. Lett. 2017, 110, 102902. [Google Scholar] [CrossRef]
  30. Park, M.; Jeon, B.; Park, J.; Kim, S. Memristors with Nociceptor Characteristics Using Threshold Switching of Pt/HfO2/TaOx/TaN Devices. Nanomaterials 2022, 12, 4206. [Google Scholar] [CrossRef]
  31. Ryu, H.; Kim, S. Synaptic Characteristics from Homogeneous Resistive Switching in Pt/Al2O3/TiN Stack. Nanomaterials 2020, 10, 2055. [Google Scholar] [CrossRef] [PubMed]
  32. Asapu, S.; Maiti, T. Multifilamentary conduction modeling in transition metal oxide-based rram. IEEE Trans. Electron Devices 2017, 64, 3145–3150. [Google Scholar] [CrossRef]
  33. Lee, M.H.; Kim, K.M.; Kim, G.H.; Seok, J.Y.; Song, S.J.; Yoon, J.H.; Hwang, C.S. Study on the electrical conduction mechanism of bipolar resistive switching TiO2 thin films using impedance spectroscopy. Appl. Phys. Lett. 2010, 96, 152909. [Google Scholar] [CrossRef]
  34. Yang, J.J.; Zhang, M.X.; Strachan, J.P.; Miao, F.; Pickett, M.D.; Kelley, R.D.; Medeiros-Ribeiro, G.; Williams, R.S. High switching endurance in TaOx memristive devices. Appl. Phys. Lett. 2010, 97, 232102. [Google Scholar] [CrossRef]
  35. Torrezan, A.C.; Strachan, J.P.; Medeiros-Ribeiro, G.; Williams, R.S. Sub-nanosecond switching of a tantalum oxide memristor. Nanotechnology 2011, 22, 485203. [Google Scholar] [CrossRef]
  36. Lee, M.J.; Lee, C.B.; Lee, D.; Lee, S.R.; Chang, M.; Hur, J.H.; Kim, Y.B.; Kim, C.J.; Seo, D.H.; Seo, S.; et al. A fast, high-endurance and scalable non-volatile memory device made from asymmetric Ta2O5-xx/TaO2-xbilayer structures. Nat. Mater. 2011, 10, 625–630. [Google Scholar] [CrossRef] [PubMed]
  37. Deuermeier, J.; Kiazadeh, A.; Klein, A.; Martins, R.; Fortunato, E. Mulit-level cell properties of a bilayer Cu2O/Al2O3 resistive switching device. Nanomaterials 2019, 9, 289. [Google Scholar] [CrossRef] [PubMed]
  38. Malenka, R.C. Synaptic plasticity in the hippocampus: LTP and LTD. Cell 1994, 78, 535–538. [Google Scholar] [CrossRef]
  39. Malenka, R.C.; Francisco, S. Review: LTP and LTD: Dynamic and interactive processes of synaptic plasticity. Neuroscientist 1995, 1, 35–42. [Google Scholar] [CrossRef]
  40. ASokolov, S.; Jeon, Y.R.; Ku, B.; Choi, C. Ar ion plasma surface modification on the heterostructured TaOx/InGaZnO thin films for flexible memristor synapse. J. Alloys Compd. 2020, 822, 153625. [Google Scholar] [CrossRef]
  41. Kim, M.K.; Lee, J.S. Short-Term Plasticity and Long-Term Potentiation in Artificial Biosynapses with Diffusive Dynamics. ACS Nano 2018, 12, 1680–1687. [Google Scholar] [CrossRef] [PubMed]
  42. Prezioso, M.; Bayat, F.M.; Hoskins, B.; Likharev, K.; Strukov, D. Self-Adaptive Spike-Time-Dependent Plasticity of Metal-Oxide Memristors. Sci. Rep. 2016, 6, 21331. [Google Scholar] [CrossRef]
  43. Zhang, Y.; Li, Y.; Wang, X.; Friedman, E.G. Synaptic Characteristics of Ag/AgInSbTe/Ta-Based Memristor for Pattern Recognition Applications. IEEE Trans. Electron Devices 2017, 64, 1806–1811. [Google Scholar] [CrossRef]
  44. Klimo, M.; Such, O.; Skvarek, O.; Fratrik, M. Memristor-based pattern matching. Semicond. Sci. Technol. 2014, 29, 104007. [Google Scholar] [CrossRef]
  45. Huang, C.Y.; Huang, C.Y.; Tsai, T.L.; Lin, C.A.; Tseng, T.Y. Switching mechanism of double forming process phenomenon in ZrOx/HfOy bilayer resistive switching memory structure with large endurance. Appl. Phys. Lett. 2014, 104, 062901. [Google Scholar] [CrossRef]
  46. Degraeve, R.; Fantini, A.; Raghavan, N.; Goux, L.; Clima, S.; Govoreanu, B.; Belmonte, A.; Lineyen, D.; Jurczak, M. Causes and consequences of the stochastic aspect of filamentary RRAM. Microelectron. Eng. 2015, 147, 171–175. [Google Scholar] [CrossRef]
  47. Oh, I.; Pyo, J.; Kim, S. Resistive Switching and Synaptic Characteristics in ZnO/TaON-Based RRAM for Neuromorphic System. Nanomaterials 2022, 12, 2185. [Google Scholar] [CrossRef]
  48. Athena, F.F.; West, M.P.; Hah, J.; Hanus, R.; Graham, S.; Vogel, E.M. Towards a better understanding of the forming and resistive switching behavior of Ti-doped HfOx RRAM. J. Mater. Chem. 2022, 10, 5896–5904. [Google Scholar] [CrossRef]
  49. Ye, C.; Zhan, C.; Tsai, T.M.; Chang, K.C.; Chen, M.C.; Chang, T.C.; Deng, T.; Wang, H. Low-power bipolar resistive switching TiN/HfO2/ITO memory with self-compliance current phenomenon. Appl. Phys. Express 2014, 7, 034101. [Google Scholar] [CrossRef]
  50. Lin, C.Y.; Chang, K.C.; Chang, T.C.; Tsai, T.M.; Pan, C.H.; Zhang, R.; Liu, K.H.; Chen, H.M.; Tseng, Y.T.; Hung, Y.C.; et al. Effects of varied negative stop voltages on current self-compliance in indium tin oxide resistance random access memory. IEEE Electron. Device Lett. 2015, 36, 564–566. [Google Scholar] [CrossRef]
  51. Lin, M.H.; Wu, M.C.; Huang, C.Y.; Lin, C.H.; Tseng, T.Y. High-speed and localized resistive switching characteristics of double-layer SrZrO3 memory devices. J. Phys. D Appl. Phys. 2010, 43, 295404. [Google Scholar] [CrossRef]
  52. Zhao, L.; Chen, H.; Wu, S.; Jiang, Z.; Yu, S.; Hou, T.; Wong, H.S.; Nishi, Y. Improved multi-level control of RRAM using pulse-train programming. In Proceedings of the Technical Program—2014 International Symposium on VLSI Technology, Systems and Application (VLSI-TSA), Hsinchu, Taiwan, 28–30 April 2014; pp. 1–2. [Google Scholar]
  53. Tsai, T.L.; Chang, H.Y.; Lou, J.J.C.; Tseng, T.Y. A high performance transparent resistive switching memory made form ZrO2/AlON bilayer structure. Appl. Phys. Lett. 2016, 108, 153505. [Google Scholar] [CrossRef]
  54. Ling, H.; Yi, M.; Nagai, M.; Xie, L.; Wang, L.; Hu, B.; Huang, W. Controllable Organic Resistive Switching Achieved by One-Step Integration of Cone-Shaped contact. Adv. Mater. 2017, 29, 1701333. [Google Scholar] [CrossRef] [PubMed]
  55. Bai, Y.; Wu, H.; Wu, R.; Zhang, Y.; Deng, N.; Yu, Z.; Qian, H. Study of multi-level characteristics for 3D vertical resistive switching memory. Sci. Rep. 2014, 4, 5780. [Google Scholar] [CrossRef]
  56. Milo, V.; Zambelli, C.; Olivo, P.; Pérez, E.; Mahadevaiah, M.K.; Ossorio, O.G.; Wenger, C.; Ielmini, D. Multilevel HfO2-based RRAM devices for low-power neuromorphic networks. APL Mater. 2019, 7, 081120. [Google Scholar]
  57. Prakash, A.; Hwang, H. Multilevel cell storage and resistance variability in resistive random access memory. Phys. Sci. Rev. 2016, 1, 20160010. [Google Scholar]
  58. Wang, Z.; Yin, M.; Zhang, T.; Cai, Y.; Wang, Y.; Yang, Y.; Huang, R. Engineering incremental resistive switching in TaOx based memristors for brain-inspired computing. Nanoscale 2016, 8, 14015–14022. [Google Scholar] [CrossRef]
  59. Jena, A.K.; Sahu, M.C.; Mohanan, K.U.; Mallik, S.K.; Sahoo, S.; Pradhan, G.K.; Sahoo, S. Bipolar Resistive Switching in TiO2 Artificial Synapse Mimicking Pavlov’s Associative Learning. ACS Appl. Mater. Interfaces 2023, 15, 3574–3585. [Google Scholar] [CrossRef]
  60. Ju, D.; Kim, J.; Kim, S. Highly uniform resistive switching characteristics of Ti/TaOx/ITO memeristor devices for neuromorphic system. J. Alloys Compd. 2023, 961, 170920. [Google Scholar] [CrossRef]
  61. Caporale, N.; Dan, Y. Spike timing–dependent plasticity: A Hebbian learning rule. Annu. Rev. Neurosci. 2008, 31, 25–46. [Google Scholar] [CrossRef]
  62. Zhang, S.R.; Zhou, L.; Mao, J.Y.; Ren, Y.; Yang, J.Q.; Yang, G.H.; Zhu, X.; Han, S.T.; Roy, V.A.L.; Zhou, Y. Artificial Synapse Emulated by Charge Trapping-Based Resistive Switching Device. Adv. Mater. Technol. 2019, 4, 1800342. [Google Scholar] [CrossRef]
  63. Wang, Z.; Wu, H.; Burr, G.W.; Hwang, C.S.; Wang, K.L.; Xia, Q.; Yang, J.J. Resistive switching materials for information processing. Nat. Rev. Mater. 2020, 5, 173–195. [Google Scholar] [CrossRef]
  64. Huang, F.; Ke, C.; Li, J.; Chen, L.; Yin, J.; Li, X.; Wu, Z.; Zhang, C.; Wu, F.; Wu, Y.; et al. Controllable Resistive Switching in ReS2/WS2 Heterostructure for Nonvolatile Memory and Synaptic Simulation. Adv. Sci. 2023, 2023, 2302813. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Schematic illustration of the device’s structure. (b) Typical cross-sectional TEM image of the Pt/TaOx/InOx/ITO structure. (c) Component distribution: Pt, Ta, O, In, and Sn.
Figure 1. (a) Schematic illustration of the device’s structure. (b) Typical cross-sectional TEM image of the Pt/TaOx/InOx/ITO structure. (c) Component distribution: Pt, Ta, O, In, and Sn.
Materials 16 06184 g001
Figure 2. XPS spectra of (a) Ta 4f (b) and O 1s at an etch time of 4 s. (c) In 3d and (d) O 1s at an etch time of 20 s.
Figure 2. XPS spectra of (a) Ta 4f (b) and O 1s at an etch time of 4 s. (c) In 3d and (d) O 1s at an etch time of 20 s.
Materials 16 06184 g002
Figure 3. (a) Typical I–V curve. (b) Bipolar resistive switching ran 275 cycles. (c) Endurance test. (d) Retention test.
Figure 3. (a) Typical I–V curve. (b) Bipolar resistive switching ran 275 cycles. (c) Endurance test. (d) Retention test.
Materials 16 06184 g003
Figure 4. Schematic description of the conduction mechanism of the double-forming process of the Pt/TaOx/InOx/ITO RRAM device in the (a) Initial state, (b) First forming process, (c) Second forming process, (d) Reset process, and (e) Set process.
Figure 4. Schematic description of the conduction mechanism of the double-forming process of the Pt/TaOx/InOx/ITO RRAM device in the (a) Initial state, (b) First forming process, (c) Second forming process, (d) Reset process, and (e) Set process.
Materials 16 06184 g004
Figure 5. (a) MLC obtained by controlling the reset voltage and (b) DC endurance performance. (c) Filament schematics.
Figure 5. (a) MLC obtained by controlling the reset voltage and (b) DC endurance performance. (c) Filament schematics.
Materials 16 06184 g005
Figure 6. (a) MLC obtained by the controlling compliance current and (b) DC endurance performance. (c) Filament schematics.
Figure 6. (a) MLC obtained by the controlling compliance current and (b) DC endurance performance. (c) Filament schematics.
Materials 16 06184 g006
Figure 7. (a) Potentiation and depression. (b) Potentiation and depression run for 10 cycles. (c) Schematic illustration of a DNN for numerical number recognition containing the input (784 neurons), hidden (3 layers), and output (10 neurons) layers. (d) Simulated recognition accuracy using the MNIST numerical datasets of raining images, with approximately 94% recognition accuracy for the Pt/TaOx/InOx/ITO memristor.
Figure 7. (a) Potentiation and depression. (b) Potentiation and depression run for 10 cycles. (c) Schematic illustration of a DNN for numerical number recognition containing the input (784 neurons), hidden (3 layers), and output (10 neurons) layers. (d) Simulated recognition accuracy using the MNIST numerical datasets of raining images, with approximately 94% recognition accuracy for the Pt/TaOx/InOx/ITO memristor.
Materials 16 06184 g007
Figure 8. (a) Schematic illustration of the human synaptic neural structure. (b) Pulse schematic. (c) Result of the STDP measurement. (d) Result of the SRDP measurement.
Figure 8. (a) Schematic illustration of the human synaptic neural structure. (b) Pulse schematic. (c) Result of the STDP measurement. (d) Result of the SRDP measurement.
Materials 16 06184 g008
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

Ju, D.; Kim, S.; Lee, S.; Kim, S. Double-Forming Mechanism of TaOx-Based Resistive Memory Device and Its Synaptic Applications. Materials 2023, 16, 6184. https://doi.org/10.3390/ma16186184

AMA Style

Ju D, Kim S, Lee S, Kim S. Double-Forming Mechanism of TaOx-Based Resistive Memory Device and Its Synaptic Applications. Materials. 2023; 16(18):6184. https://doi.org/10.3390/ma16186184

Chicago/Turabian Style

Ju, Dongyeol, Sunghun Kim, Subaek Lee, and Sungjun Kim. 2023. "Double-Forming Mechanism of TaOx-Based Resistive Memory Device and Its Synaptic Applications" Materials 16, no. 18: 6184. https://doi.org/10.3390/ma16186184

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