Next Article in Journal
Complexity-Effective Joint Detection of Physical Cell Identity and Integer Frequency Offset in 5G New Radio Communication Systems
Next Article in Special Issue
A Novel Gaussian Process Surrogate Model with Expected Prediction Error for Optimization under Constraints
Previous Article in Journal
Extended Sliding Mode Observer-Based Output Feedback Control for Motion Tracking of Electro-Hydrostatic Actuators
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Implementation of Physical Reservoir Computing in a TaOx/FTO-Based Memristor Device

1
Department of Advanced Battery Convergence Engineering, Dongguk University, Seoul 04620, Republic of Korea
2
Division of Electronics and Electrical Engineering, Dongguk University, Seoul 04620, Republic of Korea
*
Authors to whom correspondence should be addressed.
Mathematics 2023, 11(20), 4325; https://doi.org/10.3390/math11204325
Submission received: 15 September 2023 / Revised: 14 October 2023 / Accepted: 16 October 2023 / Published: 17 October 2023
(This article belongs to the Special Issue AI for Brain Science and Brain-Inspired Computing)

Abstract

:
As one of the solutions to overcome the current problems of computing systems, a resistive switching device, the TiN/TaOx/fluorine-doped tin oxide (FTO) stacked device, was fabricated to investigate its capability in embodying neuromorphic computing. The device showed good uniformity during the resistive switching phenomenon under time and cycle-to-cycle dependent switching, which may be due to the oxygen reservoir characteristics of the FTO bottom electrode, storing oxygen ions during resistive switching and enhancing the device property. Based on the uniform switching phenomenon of the TiN/TaOx/FTO device, the pulse applications were performed to seek its ability to mimic the biological brain. It was revealed that the volatile and non-volatile nature of the device can be altered by controlling the pulse stimuli, where strong stimuli result in long-term memory while weak stimuli result in short-term memory. To further investigate the key functions of the biological brain, various learning rules such as paired-pulse facilitation, excitatory postsynaptic current, potentiation and depression, spike-rate dependent plasticity, and spike-time dependent plasticity were tested, with reservoir computing implemented based on the volatile characteristic of the TiN/TaOx/FTO device.

1. Introduction

The current computing system, based on the von Neumann architecture, is encountering limitations due to the bottleneck between the processor and the memory [1,2]. Additionally, the increasing volume of data in modern computer processing has led to the emergence of various alternative architectures. One such architecture, neuromorphic computing, which seeks to emulate the biological brain, is gaining attention due to its energy-efficient parallel data processing capabilities [3,4]. In the pursuit of applications for neuromorphic systems, researchers have explored memristors, including magnetic random-access memory, resistive random-access memory (RRAM), ferroelectric random-access memory, and phase-change random-access memory, to assess their potential for mimicking synaptic behavior [5,6,7,8,9,10]. Among these devices, RRAM stands out as a promising next-generation nonvolatile memory for neuromorphic applications due to its straightforward fabrication process, high scalability, low-power operation, and rapid switching speed [11,12,13,14,15,16,17]. RRAM is a two-terminal device with a metal–insulator–metal structure, where the insulator typically consists of transition metal oxides such as NiOx, HfOx, AlOx, ZrOx, TiOx, and TaOx [18,19,20,21,22,23]. TaOx-based RRAM devices, in particular, have exhibited superior characteristics in previous studies, including high endurance (>1010 cycles), sub-nanosecond switching speeds (<1 ns), a high dielectric constant ( ϵ = 25), and compatibility with current complementary metal-oxide semiconductor processes [24,25,26,27,28,29,30]. Additionally, the low absolute value of Gibbs free energy for the redox reaction of TaOx contributes to its good stability [31].
The resistive switching phenomenon in transition metal oxide-based RRAM devices occurs due to the formation and rupture of a conducting filament [32]. When a voltage is applied to the top electrode, reduction takes place under an electric field, causing the separation of oxygen vacancies (Vo+) and oxygen ions (O2−). The accumulation of oxygen vacancies forms a conducting filament, leading to a significant current flow and a switch from a high-resistance state to a low-resistance state. Conversely, when an opposite bias is applied, oxidation occurs, reuniting the separated oxygen vacancies and ions, causing the filament to rupture, and the device returns to the high-resistance state. Recent studies have indicated that an oxygen reservoir layer, forming at the interface between the transition metal oxide and the metal electrode, can enhance switching properties through its oxygen storage capacity [33,34,35,36,37]. FTO is one of the electrodes acting as an oxygen reservoir, offering the advantages of affordability [38]. Furthermore, research has highlighted the additional benefits of the FTO electrode, including high thermal stability and processability with lower sheet resistance, making it highly suitable for scalable device fabrication and electrode applications [38,39].
To achieve neuromorphic computing, it is crucial that the device emulates the biological brain. The function of the human brain relies on the strength of connections between post- and pre-synapses, where chemical reactions facilitate information exchange [40,41]. Moreover, human memory can be categorized into short-term memory and long-term memory. Short-term memory exhibits synaptic plasticity, lasting from milliseconds to minutes, while long-term memory endures for hours or even days, with transitions occurring from short-term memory to long-term memory through repetitive activity [42,43]. Synaptic behaviors characteristic of long-term memory and short-term memory can be replicated through pulse applications, such as potentiation and depression, excitatory postsynaptic current (EPSC), paired-pulse facilitation (PPF), and more [44,45,46,47]. Furthermore, studies have shown that certain high-order synaptic plasticity states can be monitored using Hebbian learning rules [48,49], including spike-time dependent plasticity (STDP) and spike-rate dependent plasticity (SRDP), adding complexity to synapse emulation [50,51].
Additionally, due to the volatile nature of some memristor devices, reservoir computing is often implemented [52,53,54,55,56]. Reservoir computing is a computational framework suitable for processing temporal and sequential data with low power consumption [54]. Two essential factors for implementing reservoir computing are nonlinearity and volatile characteristics [57]. As the reservoir system’s high-dimensional output depends on time-dependent input without relying on previous inputs, non-linear data conversion within the reservoir layer plays a critical role [58,59].
In this paper, we present the bipolar resistive switching characteristics of the TiN/TaOx/FTO device. The FTO electrode functions as an oxygen reservoir, resulting in good uniformity over 300 DC cycles and 104 s, attributed to its oxygen ion storage capability. Moreover, by applying pulse train-induced potentiation and depression, we tested the Modified National Institute of Standards and Technology database image recognition system, achieving favorable recognition accuracy through training and testing. We also emulated biological synapses using learning tools such as PPF, EPSC, STDP, and SRDP. Finally, we explored additional applications in reservoir computing systems, revealing various potential uses for the TiN/TaOx/FTO memristive device.

2. Materials and Methods

The TiN/TaOx/FTO device was fabricated as follows. Before fabrication, a commercially available 2.2 mm thick FTO electrode on a transparent glass substrate was prepared for its use as a bottom electrode. The surface of the FTO electrode was cleaned using acetone and isopropyl alcohol. Then, a 10 nm thick TaOx film was deposited on the FTO electrode using radio frequency reactive sputtering using a Ta target (99.99% purity). The sputtering gas was a mixture of Ar and O2, at 20 and 6 sccm each. Additionally, the power and the pressure of the fabrication process were 150 W and 5 mTorr. Then, by depositing the TiN top electrode on a circular shadow mask with a diameter of 100 µm, the top electrode TiN was acquired. The TiN top electrode was deposited using DC reactive sputter, with a gas mixture of Ar (19 sccm) and N2 (1 sccm). The pressure and power were 5 mTorr and 350 W. The electrical properties of TiN/TaOx/FTO were investigated through a Keithley 4200-SCS semiconductor parameter analyzer (Keithley Instruments, Cleveland, OH, USA) and a 4225-PMU pulse measuring unit (Keithley Instruments, Cleveland, OH, USA). The top TiN electrode was biased while grounding the bottom FTO electrode.

3. Results and Discussion

Figure 1a shows the fabrication process flow of the TiN/TaOx/FTO device. Figure 1b illustrates the schematic representation of the TiN/TaOx/FTO device fabricated on a glass substrate. The electrical behavior of TiN/TaOx/FTO is elucidated in Figure 1c–f. Before achieving the desired switching, the device undergoes an initial breakdown process termed “forming,” which facilitates the transition from its initial resistance state to the low-resistance state [60]. As depicted in Figure 1c, the forming process occurs at a voltage of −3 V, and notably it does not require any current compliance. This behavior might be attributed to the FTO electrode, as previous research has indicated compliance-free switching in FTO-based RRAM devices [61,62,63]. Following the forming process, the devices exhibit a bipolar resistive switching phenomenon at voltages of −3 V and 3 V. Furthermore, Figure 1d presents the current–voltage (I-V) curves resulting from 100 DC cycle switches in the TiN/TaOx/FTO device. It is evident that the device maintains its self-compliance feature without undergoing a harsh breakdown process, exhibiting good uniformity across cycles. In Figure 1e, we present endurance data, where resistance values are recorded at −0.1 V. Additionally, Figure 1f showcases retention data, with resistance values also measured at −0.1 V. These data illustrate that the device sustains its low-resistance state and high-resistance states for 300 cycles and 104 s, respectively, demonstrating favorable device characteristics and potential relevance in the context of neuromorphic computing emulation.
Building upon the findings presented above and drawing from the conduction mechanisms established in prior investigations of TaOx-based resistive switching [64,65,66,67], we present a conduction model centered on oxygen ion migration, as depicted in Figure 2. When a negative bias is applied to the TiN top electrode, reduction processes unfold, as described in the following Equation (1) [65,68]:
O R e d u c t i o n V O 2 + + O 2
Under the influence of the electric field, the oxygen ions (O2−) and oxygen vacancies (Vo2+) undergo separation, with oxygen ions being repelled away from the top electrode. This repulsion results in the creation of defects, specifically oxygen ions, within the material. These defects accumulate steadily under the continued application of electrical stress (−3 V), culminating in the formation of a conductive filament that establishes a connection between the bottom and top electrodes. This conductive filament facilitates a substantial current flow, thereby reducing the device’s resistance. Consequently, the resistance state of the device transitions from a high-resistance state to a low-resistance state, as illustrated in Figure 2a. Conversely, when a positive bias is applied to the TiN electrode, it induces the attraction of oxygen ions back towards the top electrode. During this migration, an oxidation process takes place, as shown in Equation (2):
O O x i d a t i o n V O 2 + + O 2
In this process, oxygen ions combine with electrons to form oxygen molecules. As a consequence, the previously separated oxygen ions and vacancies recombine. This recombination event leads to the rupture of the conductive filament, causing the device to revert to its high-resistance state, as depicted in Figure 2b.
In our exploration of the potential applications of the TiN/TaOx/FTO device, we employed pulse schemes to examine its conductance response. One of the widely recognized techniques for assessing the current behavior of this device involves conducting experiments related to potentiation and depression [69]. The pulse scheme used to acquire this behavior is illustrated in Figure 3a. As depicted, we applied a sequence of 50 consecutive set and reset pulses to the TiN/TaOx/FTO device in order to observe a gradual change in conductance. The pulse parameters for both set and reset operations included an amplitude of −2.5 V and a width of 50 µs for set pulses, and 2.7 V and 70 µs for reset pulses. After each set and reset pulse, we followed up with read pulses of −0.1 V and 50 µs to assess the conductance change. The results of this pulse application are presented in Figure 3b. Furthermore, during the pulse applications, the device exhibited short-term memory characteristics, as depicted in Figure 3c. When a single pulse of −2.5 V and 500 µs was applied to the device, followed by a 5 ms read pulse, the current value experienced a decrease after a certain period. This decline in current can be interpreted as a process akin to “forgetting” in the human brain, mirroring the tendency of short-term memory in the biological brain to fade over time [70]. Moreover, it is worth noting that the transformation of short-term memory into long-term memory is attainable by regulating the number of input pulses [71]. In Figure 3d, a set pulse sequence of −2.5 V and 500 µs was iterated four times, emulating the rehearsal process observed in the biological brain. Notably, after the fourth set pulse, a diminished current decay was evident, signifying the transformation of short-term memory into long-term memory.
To test further neuromorphic applications, the modified National Institute of Standards and Technology Database pattern recognition system based on deep neural network was tested using Python in Google Collab. For the training and testing process, 28 × 28-pixel handwritten number images were converted from the potentiation and depression data of Figure 3b. In this context, the pixel values increase or decrease compared to the preceding image, following the pattern of potentiation or depression, respectively. Therefore, when asymmetric graphs for potentiation and depression are obtained, training is carried out with uncleared images. Conversely, the more symmetric the potentiation and depression graphs, the clearer the images, and consequently higher accuracies can be achieved. The structure of the deep neural network consists of three layers: the input, hidden, and output layers, as shown in Figure 4a. The calculated recognition accuracy is illustrated in Figure 4b, where after 10 epochs a maximum accuracy of 94.84% was obtained.
PPF is a function utilized to observe short-term memory characteristics in synaptic devices. As illustrated in Figure 5a, twin paired pulses were administered to the device with varying intervals between them. When the interval time was brief, the device retained the memory of the preceding current response, resulting in a more substantial current output upon the second pulse. Conversely, if the interval time was sufficiently long, the device tended to forget the previous response, leading to a similar current response. To assess PPF, twin pulses of −2.5 V and 1 ms were applied, while the interval time ranged from 10 µs to 20 ms, enabling the observation of current responses. The outcomes of PPF are presented in Figure 5b. The PPF index, defined by Equation (3), quantifies the relationship:
P P F   I n d e x % = I 2 I 1 × 100
where I1 and I2 represent the current responses after the first and second pulses, respectively. Additionally, EPSC was explored to investigate the device’s amplitude and interval-dependent characteristics. To obtain EPSC, pulses of varying quantities, ranging from 1 to 50, were administered to the TiN/TaOx/FTO device, as illustrated in Figure 5c. Read pulses of −0.1 V and 1 ms followed the variable set pulses to monitor current changes. The amplitude-dependent EPSC response is displayed in Figure 5d, demonstrating that larger amplitudes result in more significant EPSC responses. The interval-dependent EPSC response is depicted in Figure 5e, revealing that a more substantial EPSC response is achieved with shorter pulse intervals. These EPSC response findings indicate that the devices exhibit favorable input-current responses, making them capable of facile output modulations.
Moreover, we conducted tests involving Hebbian learning rules, one of which is STDP. STDP explores the relationship between synaptic weight (ΔW) and spike timing (Δt). Thanks to the straightforward two-terminal structure of RRAM, it offers the advantage of easily emulating the architectural principles of biological synapses, where the pre-and post-synaptic elements are mimicked by the top and bottom electrodes, as depicted in Figure 6a [72]. In this configuration, synaptic information migrates between the pre- and post-synapse modulated by the synaptic weight, akin to the growth of the conductive filament within the insulating layer. This filament connects the top and bottom electrodes, facilitating the flow of current. In STDP, two primary conditions exist. When the pre-synapse precedes the post-synapse in spike timing (Δt > 0), synaptic weight is enhanced, resulting in potentiation. Conversely, when the post-synapse precedes the pre-synapse in spike timing (Δt < 0), the synaptic weight decreases, leading to depression [65]. The relationships between synaptic weight (ΔW) and spike time (Δt) are formulated in Equations (4) and (5):
Δ t = t p r e t p o s e
Δ W   % = G f G i G i × 100
where tpre, tpost, Gf, and Gi represent the spike firing time pre-synapse, the spike firing time post-synapse, conductance after pulse application, and conductance before pulse application, respectively. To emulate spike firing, a pulse train comprising voltage levels of 1.5 V, −2.5 V, −2.3 V, −2.1 V, −1.9 V, and −1.7 V was applied at different time points, as indicated in the inset of Figure 6b. The results are depicted in Figure 6b, showcasing a gradual change in conductance reminiscent of long-term potentiation (LTP) and long-term depression (LTD). Furthermore, we also conducted experiments involving another Hebbian learning rule known as SRDP. In the context of SRDP, we applied a series of 10 consecutive set pulses, each with a voltage of −2.5 V and a duration of 1 ms, to the TiN/TaOx/FTO device. These pulses were administered with varying pulse intervals between each set, spanning five different conditions: 1 µs, 10 µs, 100 µs, 1 ms, and 10 ms. As illustrated in Figure 6c, the results demonstrate a favorable SRDP response. The SRDP index is mathematically expressed in Equation (6):
S R D P   I n d e x   % = I n I i × 100
where In and Ii represent the current response of the initial state and after pulse applications, respectively.
Finally, based on the coexistence of short-term memory and long-term memory of the TiN/TaOx/FTO device, reservoir computing was implemented. Reservoir computing is a computational paradigm designed for the processing of temporal and sequential data [52]. By implementing physical reservoir computing, complex tasks can be performed, including time-series forecasting, pattern recognition, and pattern generation [73,74,75]. Figure 7a represents the reservoir computing system, which consists of three layers: the input, reservoir, and output layers. The input layer accepts and encodes the incoming time-dependent data into the resistance values of the RRAM cells within the reservoir. Then, the reservoir layer functions as a dynamic and recurrent computational medium, converting input signals into high-dimensional output through nonlinear transformation and information mixing. Lastly, the output layer generates the desired output or prediction based on the information of the reservoir layer. A computational system designed to encode 4-bit data by applying time-dependent pulse train inputs to five devices is illustrated in Figure 7b. When an electrical pulse is applied as input to the different devices, it induces a change in the device’s current state. Hence, due to the short-term memory characteristics of the device, for the next input, the current state of the device will return to its former state, preventing the device from being influenced by previous inputs. The change in current states can be figured out through read pulses following the inputs, enabling its representation of 4-bit data. Figure 7c represents the [1010] state of the device, which is one of sixteen states ranging from [0000] to [1111]. The “1” state of reservoir computing was obtained by applying a set pulse of −2.5 V, 1 ms to the device, with a read pulse of −0.1 V following. For the “0” state, only read pulses were applied. The interval between each following state was modulated as 20 ms, to ensure that the former input had minimal impact on the prior output. Thus, by applying “1”, “0”, “1”, and “0” pulse sequentially with appropriate intervals, state [1010] was obtained. Using “1” and “0” states, 16 different states of reservoir computing are shown in Figure 7d. The “1” and “0” can be distinguished through the current value between 1.82 and 2.59 µA. Furthermore, the letter “K” can be made by combining 4-bit data from five different devices. The black pixel represents “1”, while the white pixel represents “0”. As a result, by merging [1001], [1010], and [1100], the letter “K” consisted of 5 × 4 pixels, as depicted in Figure 7e.

4. Conclusions

In summary, we fabricated a TiN/TaOx/FTO device and investigated its electrical characteristics for neuromorphic applications. The device exhibited bipolar switching for 300 cycles with uniform resistance states. The self-compliance behavior during cycle-repetition may be due to the FTO electrode, which is helpful in avoiding permanent device breakdown. Furthermore, during the pulse applications, the coexistence of short-term memory and long-term memory was observed through current decay and rehearsal. Due to this coexistence, various functions such as potentiation and depression, PPF, EPSC, STDP, and SRDP were mimicked, showing favorable characteristics in emulating biological synapses. Lastly, based on the short-term memory behavior of the device, reservoir computing was implemented. We believe that these results of various implementations of TiN/TaOx/FTO devices depict various applications for the future neuromorphic system.

Author Contributions

D.J.: Conceptualization, Data curation, Formal analysis, Writing—original draft; J.A.: Software; J.H.: Software; S.K.: Funding acquisition, Writing—review and editing, Supervision, Validation; D.C.: Funding acquisition, Writing—review and editing, Supervision, Validation. 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 Korean government (MSIP) under Grant 2021K1A3A1A49098073 and the Korea Institute of Energy Technology Evaluation and Planning (KETEP), and grant funded by the Korean government (MOTIE) under Grant 20224000000020.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yu, S.; Gao, B.; Fang, Z.; Yu, H.; Kang, J.; Wong, H.S.P. Stochastic learning in oxide binary synaptic deivce for neuromorphic system. Front. Neurosci. 2013, 7, 186. [Google Scholar] [CrossRef]
  2. Kang, D.; Jang, J.T.; Park, S.; Ansari, M.H.R.; Bae, J.H.; Choi, S.J.; Kim, D.H.; Kim, C.; Cho, S.; Kim, D.H. Threshold-Variation-Tolerant Coupling-Gate α-IGZO synaptic transistor for more reliably controllable hardware neuromorphic system. IEEE Access 2021, 9, 59345–59352. [Google Scholar] [CrossRef]
  3. Wang, Z.; Joshi, S.; Savel’EV, S.; Song, W.; Midya, R.; Li, Y.; Rao, M.; Yan, P.; Asapu, S.; Zhou, Y.; et al. Fully memristive neural networks for pattern classification with unsupervised learning. Nat. Electron. 2018, 1, 137–145. [Google Scholar] [CrossRef]
  4. Zidan, M.A.; Strachan, J.P.; Lu, W.D. The future of electronics based on memristive systems. Nat. Electron. 2018, 1, 22–29. [Google Scholar] [CrossRef]
  5. Kim, K.; Lee, S. Integration of lead zirconium titanate thin films for high density ferroelectric random access memory. J. Appl. Phys. 2006, 100, 051604. [Google Scholar] [CrossRef]
  6. Burr, G.W.; Breitwisch, M.J.; Franceschini, M.; Garetto, D.; Gopalakrishnan, K.; Jackson, B.; Kurdi, B.; Lam, C.; Lastras, L.A.; Padilla, A.; et al. Phase change memory technology. J. Vac. Sci. Technol. B 2010, 28, 223–262. [Google Scholar] [CrossRef]
  7. Fong, S.W.; Neumann, C.M.; Wong, H.S.P. Phase-change memory-Towards a storage-class memory. IEEE Trans. Electron Devices 2017, 64, 4374–4385. [Google Scholar] [CrossRef]
  8. Khvalkovskiy, A.V.; Apalkov, D.; Watts, S.; Chepulskii, R.; Beach, R.S.; Ong, A.; Tang, X.; Smith, A.D.; Butler, W.H.; Visscher, P.B.; et al. Basic principles of STT-MRAM cell operation in memory arrays. J. Phys. D Appl. Phys. 2013, 46, 074001. [Google Scholar] [CrossRef]
  9. Ryu, H.; Kim, S. Implementation of a reservoir computing system using the short-term effects of Pt/HfO2/TaOX/TiN memristors with self-rectification. Chaos Solit. Fractals 2011, 150, 111223. [Google Scholar] [CrossRef]
  10. Pawar, P.S.; Tikke, R.S.; Patil, V.B.; Mullani, N.B.; Waifalkar, P.P.; Khot, K.V.; Teli, A.M.; Sheikh, A.D.; Dongale, T.D. A low-cost copper oxide thin film memristive device based on successive ionic layer adsorption and reaction method. Mater. Sci. Semicond. Process 2017, 71, 102–108. [Google Scholar] [CrossRef]
  11. Jeong, D.S.; Thomas, R.; Katiyar, R.S.; Scott, J.F.; Kohlstedt, H.; Petraru, A.; Hwang, C.S. Emerging memories: Resistive switching mechanisms and current status. Rep. Prog. Phys. 2012, 75, 076502. [Google Scholar] [CrossRef] [PubMed]
  12. Waser, R.; Aono, M. Nanoionics-based resistive switching memories. Nat. Mater. 2007, 6, 833–840. [Google Scholar] [CrossRef] [PubMed]
  13. Ventra, M.D.; Pershin, Y.V. Memory materials: A unifying description. Mater. Today 2011, 14, 584–591. [Google Scholar] [CrossRef]
  14. Pan, F.; Gao, S.; Chen, C.; Song, C.; Zeng, F. Recent progress in resistive random access memories: Materials, switching mechanisms, and performance. Mater. Sci. Eng. R-Rep. 2014, 83, 1–59. [Google Scholar] [CrossRef]
  15. Kim, D.; Shin, J.; Kim, S. Resistive switching Resistive Switching Characteristics of ZnO-Based RRAM on Silicon Substrate. Metals 2021, 11, 1572. [Google Scholar] [CrossRef]
  16. Bae, D.; Lee, D.; Jung, J.; Kim, S.; Kim, H.D. Self-rectifying resistive switching characteristics of Ti/Zr3N2/p-Si capacitor for array applications. Ceram. Int. 2021, 47, 21943–21949. [Google Scholar] [CrossRef]
  17. Jeon, B.; Kim, S. Effect of ITO electrode on conductance quantization and multi-level cells in TiN/SiOX/ITO devices. Ceram. Int. 2023, 49, 424–430. [Google Scholar] [CrossRef]
  18. Li, Y.; Fang, P.; Fan, X.; Pei, Y. NiO-based memristor with three resistive switching modes. Semicond. Sci. Technol. 2020, 35, 055004. [Google Scholar] [CrossRef]
  19. Wang, Y.; Liu, Q.; Long, S.; Wang, W.; Wang, Q.; Zhang, M.; Zhang, S.; Li, Y.; Zuo, Q.; Yang, J.; et al. Investigation of resistive switching in Cu-doped HfO2 thin film for multilevel non-volatile memory applications. Nanotechnology 2009, 21, 045202. [Google Scholar] [CrossRef]
  20. Lee, Y.; Park, J.; Chung, D.; Lee, K.; Kim, S. Multi-level Cells and Quantized Conductance Characteristics of Al2O3-Based RRAM Device for Neuromorphic System. Nanoscale Res. Lett. 2022, 17, 84. [Google Scholar] [CrossRef]
  21. Lin, C.Y.; Wu, C.Y.; Wu, C.Y.; Lee, T.C.; Yang, F.L.; Hu, C.; Tseng, T.Y. Effect of Top Electrode Material on Resistive Switching Properties of ZrO2 Film Memory Devices. IEEE Electron. Device Lett. 2007, 28, 366–368. [Google Scholar] [CrossRef]
  22. Dongale, T.D.; Desai, N.D.; Khot, K.V.; Volos, C.K.; Bhosale, P.N.; Kamat, R.K. An electronic synapse device based on TiO2 thin film memristor. J. Nanoelectron. Optoelectron. 2018, 13, 68–75. [Google Scholar] [CrossRef]
  23. Yang, Y.; Sheridan, P.; Lu, W. Complementary resistive switching in tantalum oxide-based resistive memory devices. Appl. Phys. Lett. 2012, 100, 203112. [Google Scholar] [CrossRef]
  24. Yang, J.J.; Zhang, M.X.; Strachan, J.P.; Miao, F.; Pickett, M.D.; Kelley, R.D.; Ribeiro, G.M.; Williams, R.D. High switching endurance in TaOX memristive devices. Appl. Phys. Lett. 2010, 97, 232102. [Google Scholar] [CrossRef]
  25. Torrezan, A.C.; Strachan, J.P.; Ribeiro, G.M.; Williams, R.D. Sub-nanosecond switching of a tantalum oxide memristor. Nanotechnology 2011, 22, 485203. [Google Scholar] [CrossRef]
  26. Prakash, A.; Jana, D.; Maikap, S. TaOx -based resistive switching memories: Prospective and challenges. Nanoscale Res. Lett. 2013, 8, 1–17. [Google Scholar] [CrossRef]
  27. 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-X/TaO2-X bilayer structures. Nat. Mater. 2011, 10, 625–630. [Google Scholar] [CrossRef]
  28. Chen, C.; Song, C.; Yang, J.; Zeng, F.; Pan, F. Oxygen migration induced resistive switching effect and its thermal stability in W/TaOX/Pt structure. Appl. Phys. Lett. 2012, 100, 253509. [Google Scholar] [CrossRef]
  29. Bishop, S.M.; Bakhru, H.; Capulong, J.O.; Cady, N.C. Influence of the SET current on the resistive switching properties of tantalum oxide created by oxygen implantation. Appl. Phys. Lett. 2012, 100, 142111. [Google Scholar] [CrossRef]
  30. Gu, T.; Tada, T.; Watanabe, S. Conductive path formation in the Ta2O5 atomic switch: First-principles analyses. ACS Nano 2010, 4, 6477–6482. [Google Scholar] [CrossRef]
  31. Odagawa, A.; Katoh, Y.; Kanzawa, Y.; Wei, Z.; Mikawa, T.; Muraoka, S.; Takagi, T. Electroforming and resistance-switching mechanism in a magnetite thin film. Appl. Phys. Lett. 2007, 91, 133503. [Google Scholar] [CrossRef]
  32. Sawa, A. Resistive switching in transition metal oxides. Mater. Today 2008, 11, 28–36. [Google Scholar] [CrossRef]
  33. Chen, P.H.; Su, Y.T.; Chang, F.C. Stabilizing Resistive Switching Characteristics by Inserting Indium-Tin-Oxide Layer as Oxygen Ion Reservoir in HfO2-Based Resistive Random Access Memory. IEEE. Trans. Electron. Devices 2019, 66, 1276–1280. [Google Scholar] [CrossRef]
  34. Chiang, K.K.; Chen, J.S.; Wu, J.J. Aluminum Electrode Modulated Bipolar Resistive Switching of Al/Fuel-Assisted NiOX/ITO Memory Devices Modeled with a Dual-Oxygen-Reservoir Structure. ACS Appl. Mater. Interfaces 2012, 4, 4237–4245. [Google Scholar] [CrossRef]
  35. Chang, W.Y.; Huang, H.W.; Wang, W.T.; Hou, C.H.; Chueh, Y.L.; He, J.H. High uniformity of resistive switching characteristics in a Cr/ZnO/Pt device. J. Electrochem. Soc. 2012, 159, G29–G32. [Google Scholar] [CrossRef]
  36. Sun, B.; Liu, Y.X.; Liu, L.F.; Xu, N.; Wang, Y.; Liu, X.Y.; Han, R.Q.; Kang, J.F. Highly uniform resistive switching characteristics of TiN/ZrO2/Pt memory devices. J. Appl. Phys. 2009, 105, 061630. [Google Scholar] [CrossRef]
  37. Chen, S.C.; Chang, T.C.; Chen, S.Y.; Chen, C.W.; Chen, S.C.; Sze, S.M.; Tsai, M.J.; Kao, J.K.; Huang, F.Y. Bipolar resistive switching of chromium oxide for resistive random access memory. Solid-State Electron. 2011, 62, 40–43. [Google Scholar] [CrossRef]
  38. Nguyen, H.H.; Ta, H.K.T.; Park, S.; Phan, T.B.; Pham, N.K. Resistive switching effect and magnetic properties of iron oxide nanoparticles embedded-polyvinyl alcohol film. RSC Adv. 2020, 10, 12900–12907. [Google Scholar] [CrossRef]
  39. Bear, M.F.; Conners, B.W.; Paradiso, M.A. Neuroscience: Exploring the Brain; Lippincott Williams & Wilkins: Philadelphia, PA, USA, 2007. [Google Scholar]
  40. Kirkwood, A.; Rioult, M.G.; Bear, M.F. Experience-dependent modification of synaptic plasticity in visual cortex. Nature 1996, 381, 526–528. [Google Scholar] [CrossRef]
  41. Yang, R.; Huang, H.M.; Guo, X. Memristive Synapses and Neurons for Bioinspired Computing. Adv. Electron. Mater. 2019, 5, 1900287. [Google Scholar] [CrossRef]
  42. Nicholls, J.G.; Martin, A.R.; Fuchs, P.A.; Brown, D.A.; Diamond, M.E.; Weisblat, D.A. From Neuron to Brain; Sinauer Associates: Sunderland, MA, USA, 2012. [Google Scholar]
  43. 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] [PubMed]
  44. Moon, K.; Park, S.; Jang, J.; Lee, D.; Woo, J.; Cha, E.; Lee, S.; Park, J.; Song, J.; Koo, Y.; et al. Hardware implementation of associative memory characteristics with analogue-type resistive-switching device. Nanotechnology 2014, 25, 495204. [Google Scholar] [CrossRef] [PubMed]
  45. Ryu, H.; Kim, S. Self-Rectifying Resistive Switching and Short-Term Memory Characteristics in Pt/HfO2/TaOX/TiN Artificial Synaptic Device. Nanomaterials 2020, 10, 2159. [Google Scholar] [CrossRef]
  46. Wang, W.; Covi, E.; Milozzi, A.; Farronato, M.; Ricci, S.; Sbandati, S.; Pedretti, G.; Lemini, D. Neuromorphic motion detection and orientation selectivity by volatile resistive switching memories. Adv. Intell. Syst. 2021, 3, 2000224. [Google Scholar] [CrossRef]
  47. Chen, Z.; Zhang, Y.; Yu, Y.; Li, Y.; Li, Q.; Li, T.; Zhao, H.; Li, Z.; Bing, P.; Yao, J. Resistive switching memory based on polyvinyl alcohol-graphene oxide hybrid material for the visual perception nervous system. Mater. Des. 2022, 223, 111218. [Google Scholar] [CrossRef]
  48. Caporale, N.; Dan, Y. Spike timing–dependent plasticity: A Hebbian learning rule. Annu. Rev. Neurosci. 2008, 31, 25–46. [Google Scholar] [CrossRef]
  49. Munakata, Y.; Pfaffly, J. Hebbian learning and development. Dev. Sci. 2004, 7, 141–148. [Google Scholar] [CrossRef]
  50. Ku, B.; Abbas, Y.; Kim, S.; Sokolov, A.S.; Jeon, Y.R.; Choi, C. Improved resistive switching and synaptic characteristics using Ar plasma irradiation on the Ti/HfO2 interface. J. Alloys Compd. 2019, 797, 277–283. [Google Scholar] [CrossRef]
  51. Ryu, H.; Choi, J.; Kim, S. Voltage Amplitude-Controlled Synaptic Plasticity from Complementary Resistive Switching in Alloying HfOX with AlOX-Based RRAM. Metals 2020, 10, 1410. [Google Scholar] [CrossRef]
  52. Tanaka, G.; Yamane, T.; Heroux, J.B.; Nakane, R.; Kanazawa, N.; Takeda, S.; Numata, H.; Nakano, D.; Hirose, A. Recent advances in physical reservoir computing: A review. Neural Netw. 2019, 115, 100–123. [Google Scholar] [CrossRef]
  53. Du, C.; Cai, F.; Zidan, M.A.; Ma, W.; Lee, S.H.; Lu, W.D. Reservoir computing using dynamic memristors for temporal information processing. Nat. Commun. 2017, 8, 2204. [Google Scholar] [CrossRef]
  54. Mao, J.Y.; Zheng, Z.; Xiong, Z.Y.; Huang, P.; Ding, G.L.; Wang, R.; Wang, Z.P.; Yang, J.Q.; Zhou, Y.; Zhai, T.; et al. Lead-free monocrystalline perovskite resistive switching device for temporal information processing. Nano Energy 2020, 71, 104616. [Google Scholar] [CrossRef]
  55. Milano, G.; Pedretti, G.; Montano, K.; Ricci, S.; Hashemkhani, S.; Boarino, L.; Lemini, D.; Ricciardi, C. In materia reservoir computing with a fully memristive architecture based on self-organizing nanowire networks. Nat. Mater. 2021, 21, 195–202. [Google Scholar] [CrossRef] [PubMed]
  56. Anna, N.M.; Nikita, V.P.; Vsevolod, A.K.; Silvia, B.; Anton, A.M.; Andrey, D.T.; Aleksandr, A.N.; Sergey, A.Z.; Yulia, N.M.; Matteo, P.; et al. Combination of Organic-Based Reservoir Computing and Spiking Neuromorphic Systems for a Robust and Efficient Pattern Classification. Adv. Intell. Syst. 2023, 5, 2200407. [Google Scholar]
  57. Kim, D.; Shin, J.; Kim, S. Implementation of reservoir computing using volatile WOX-based memristor. Appl. Surf. Sci. 2022, 599, 153876. [Google Scholar] [CrossRef]
  58. 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]
  59. Chen, Z.; Li, W.; Fan, Z.; Dong, S.; Chen, Y.; Qin, M.; Zeng, M.; Lu, X.; Zhou, G.; Gao, X.; et al. All-ferroelectric implementation of reservoir computing. Nat. Commun. 2023, 14, 3585. [Google Scholar] [CrossRef] [PubMed]
  60. Yang, Y.; Lu, W. Nanoscale resistive switching devices: Mechanisms and modeling. Nanoscale 2013, 5, 10076–10092. [Google Scholar] [CrossRef]
  61. Munjal, S.; Khare, N. Electroforming free controlled bipolar resistive switching in Al/CoFe2O4/FTO device with self-compliance effect. Appl. Phys. Lett. 2018, 112, 073502. [Google Scholar] [CrossRef]
  62. Pandey, V.; Nehla, P.; Munjal, S. Controlled Self Compliance Filamentary Memory Behavior in Al/NiFe2O4/FTO Resistive Switching Device. Proc. Nat. Acad. Sci. India A 2023, 93, 1–7. [Google Scholar] [CrossRef]
  63. Pandey, V.; Adiba, A.; Ahmad, T.; Nehla, P.; Munjal, S. Forming-free bipolar resistive switching characteristics in Al/Mn3O4/FTO RRAM device. J. Phys. Chem. Solids 2022, 165, 110689. [Google Scholar] [CrossRef]
  64. Huang, Y.J.; Chao, S.C.; Lien, D.H.; Wen, C.Y.; He, J.H.; Lee, S.C. Dual-functional Memory and Threshold Resistive Switching Based on the Push-Pull Mechanism of Oxygen Ions. Sci. Rep. 2016, 6, 23945. [Google Scholar] [CrossRef]
  65. Ju, D.; Kim, J.H.; Kim, S. Highly uniform resistive switching characteristics of Ti/TaOX/ITO memristor devices for neuromorphic system. J. Alloys Compd. 2023, 961, 170920. [Google Scholar] [CrossRef]
  66. Ma, Y.; Li, D.; Herzing, A.A.; Cullen, D.A.; Sneed, B.T.; More, K.L.; Nuhfer, N.T.; Bain, J.A.; Skowronski, M. Formation of the Conducting Filament in TaOX-Resistive Switching Devices by Thermal-Gradient-Induced Cation Accumulation. ACS Appl. Mater. Interfaces 2018, 10, 23187–23197. [Google Scholar] [CrossRef] [PubMed]
  67. Kurnia, F.; Liu, C.; Jung, C.U.; Lee, B.W. The evolution of conducting filaments in forming-free resistive switching Pt/TaOX/Pt structures. Appl. Phys. Lett. 2013, 102, 152902. [Google Scholar] [CrossRef]
  68. Zhu, X.; Su, W.; Liu, Y.; Hu, B.; Pan, L.; Lu, W.; Zhang, J.; Li, R.W. Observation of conductance quantization in oxide-based resistive switching memory. Adv. Mater. 2012, 24, 3941–3946. [Google Scholar] [CrossRef]
  69. Kim, H.J.; Baek, Y.J.; Choi, Y.J.; Kang, C.J.; Lee, H.H.; Kim, M.H.; Kim, K.B.; Yoon, T.S. Digital versus analog resistive switching depending on the thickness of nickel oxide nanoparticle assembly. RCS Adv. 2013, 3, 20978–20983. [Google Scholar] [CrossRef]
  70. Nicholls, J.G.; Martin, A.R.; Wallace, B.G.; Fuchs, P.A. From Neuron to Brain; Oxford University Press: Oxford, UK, 2001; p. 271. [Google Scholar]
  71. Lin, Y.; Zeng, T.; Xu, H.; Wang, Z.; Zhao, X.; Liu, W.; Ma, J.; Liu, Y. Transferable and Flexible Artificial Memristive Synapse Based on WOX Schottky Junction on Arbitrary Substrates. Adv. Electron. Mater. 2018, 4, 1800373. [Google Scholar] [CrossRef]
  72. 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]
  73. Hauser, H.; Ijspeert, A.J.; Füchslin, R.M.; Pfeifer, R.; Maass, W. The role of feedback in morphological computation with compliant bodies. Biol. Cybern. 2012, 106, 595–613. [Google Scholar] [CrossRef]
  74. Wyffels, F.; Schrauwen, B. A comparative study of reservoir computing strategies for monthly time series prediction. Neurocomputing 2010, 73, 1958–1964. [Google Scholar] [CrossRef]
  75. Paquot, Y.; Duport, F.; Smerieri, A.; Dambre, J.; Schrauwen, B.; Haelterman, M.; Massar, S. Optoelectronic reservoir computing. Sci. Rep. 2012, 2, 287. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Fabrication process sequence. (b) Schematic illustration of TiN/TaOx/FTO. (c) I–V characteristics of TiN/TaOx/FTO under DC bias. (d) I–V curves of 102 cycle set and reset process. (e) Endurance for 300 DC sweep. (f) Retention for 104 s.
Figure 1. (a) Fabrication process sequence. (b) Schematic illustration of TiN/TaOx/FTO. (c) I–V characteristics of TiN/TaOx/FTO under DC bias. (d) I–V curves of 102 cycle set and reset process. (e) Endurance for 300 DC sweep. (f) Retention for 104 s.
Mathematics 11 04325 g001
Figure 2. Schematic illustration of the conduction mechanism of the TiN/TaOx/FTO device. (a) Set process and (b) reset process.
Figure 2. Schematic illustration of the conduction mechanism of the TiN/TaOx/FTO device. (a) Set process and (b) reset process.
Mathematics 11 04325 g002
Figure 3. (a) Schematic illustration of the pulse scheme to obtain potentiation and depression. (b) Gained potentiation and depression curves. (c) Short-term memory characteristic of TiN/TaOx/FTO device under pulse application. (d) Conversion of short-term memory to long-term memory demonstrated in TiN/TaOx/FTO device.
Figure 3. (a) Schematic illustration of the pulse scheme to obtain potentiation and depression. (b) Gained potentiation and depression curves. (c) Short-term memory characteristic of TiN/TaOx/FTO device under pulse application. (d) Conversion of short-term memory to long-term memory demonstrated in TiN/TaOx/FTO device.
Mathematics 11 04325 g003
Figure 4. (a) Schematic illustration of the deep neural network simulation framework for modified National Institute of Standards and Technology Database pattern recognition. (b) Pattern recognition accuracy acquired through ten consecutive epochs.
Figure 4. (a) Schematic illustration of the deep neural network simulation framework for modified National Institute of Standards and Technology Database pattern recognition. (b) Pattern recognition accuracy acquired through ten consecutive epochs.
Mathematics 11 04325 g004
Figure 5. (a) Demonstration of PPF pulse scheme. (b) The result of the PPF as a function of time interval. (c) Demonstration of EPSC pulse scheme. (d) EPSC gain obtained under varied pulse amplitude conditions. (e) EPSC gain obtained under varied pulse interval conditions.
Figure 5. (a) Demonstration of PPF pulse scheme. (b) The result of the PPF as a function of time interval. (c) Demonstration of EPSC pulse scheme. (d) EPSC gain obtained under varied pulse amplitude conditions. (e) EPSC gain obtained under varied pulse interval conditions.
Mathematics 11 04325 g005
Figure 6. (a) Schematic illustration of the two-terminal RRAM device emulating biological synapse. (b) The result of STDP as a function of spike time. (c) The result of SRDP as a function of pulse number.
Figure 6. (a) Schematic illustration of the two-terminal RRAM device emulating biological synapse. (b) The result of STDP as a function of spike time. (c) The result of SRDP as a function of pulse number.
Mathematics 11 04325 g006
Figure 7. (a) The schematic illustration of the concept of reservoir computing. (b) A reservoir computing system with a pulse train application, five RRAM devices, and the output node. (c) [1010] state acquired through reservoir computing implementation. (d) Sixteen different states acquired through reservoir computing implementation. (e) Letter “K” implemented using 5 × 4 pixels.
Figure 7. (a) The schematic illustration of the concept of reservoir computing. (b) A reservoir computing system with a pulse train application, five RRAM devices, and the output node. (c) [1010] state acquired through reservoir computing implementation. (d) Sixteen different states acquired through reservoir computing implementation. (e) Letter “K” implemented using 5 × 4 pixels.
Mathematics 11 04325 g007
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.; Ahn, J.; Ho, J.; Kim, S.; Chung, D. Implementation of Physical Reservoir Computing in a TaOx/FTO-Based Memristor Device. Mathematics 2023, 11, 4325. https://doi.org/10.3390/math11204325

AMA Style

Ju D, Ahn J, Ho J, Kim S, Chung D. Implementation of Physical Reservoir Computing in a TaOx/FTO-Based Memristor Device. Mathematics. 2023; 11(20):4325. https://doi.org/10.3390/math11204325

Chicago/Turabian Style

Ju, Dongyeol, Junyoung Ahn, Jungwoo Ho, Sungjun Kim, and Daewon Chung. 2023. "Implementation of Physical Reservoir Computing in a TaOx/FTO-Based Memristor Device" Mathematics 11, no. 20: 4325. https://doi.org/10.3390/math11204325

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