Next Article in Journal
Residual Flexural Capacity of Corroded Prestressed Reinforced Concrete Beams
Next Article in Special Issue
Irregular Resistive Switching Behaviors of Al2O3-Based Resistor with Cu Electrode
Previous Article in Journal
Flotation Separation of Chalcopyrite and Molybdenite Assisted by Microencapsulation Using Ferrous and Phosphate Ions: Part II. Flotation
Previous Article in Special Issue
Liquid-Phase Deposition Synthesis of ZIF-67-Derived Synthesis of Co3O4@TiO2 Composite for Efficient Electrochemical Water Splitting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gradually Tunable Conductance in TiO2/Al2O3 Bilayer Resistors for Synaptic Device

Division of Electronics and Electrical Engineering, Dongguk University, Seoul 04620, Korea
*
Author to whom correspondence should be addressed.
Submission received: 14 February 2021 / Revised: 4 March 2021 / Accepted: 5 March 2021 / Published: 7 March 2021

Abstract

:
In this work, resistive switching and synaptic behaviors of a TiO2/Al2O3 bilayer device were studied. The deposition of Pt/Ti/TiO2/Al2O3/TiN stack was confirmed by transmission electron microscopy (TEM) and energy X-ray dispersive spectroscopy (EDS). The initial state before the forming process followed Fowler-Nordheim (FN) tunneling. A strong electric field was applied to Al2O3 with a large energy bandgap for FN tunneling, which was confirmed by the I-V fitting process. Bipolar resistive switching was conducted by the set process in a positive bias and the reset process in a negative bias. High-resistance state (HRS) followed the trap-assisted tunneling (TAT) model while low-resistance state (LRS) followed the Ohmic conduction model. Set and reset operations were verified by pulse. Moreover, potentiation and depression in the biological synapse were verified by repetitive set pulses and reset pulses. Finally, the device showed good pattern recognition accuracy (~88.8%) for a Modified National Institute of Standards and Technology (MNIST) handwritten digit database in a single layer neural network including the conductance update of the device.

1. Introduction

Resistive switching (memristive effect) was observed from an SiOx layer in the 1960s. Resistance-based memory is considered a computing element and a storage element. Resistance-based memories are largely classified into three types, with each memory type having its advantages and disadvantages. Spin-transfer torque (STT)-magnetic random-access memory (MRAM) [1] has fast switching speed and long endurance. However, its on/off ratio is small and its scalability is inferior to others. Phase-change random-access-memory (PcRAM) [2] has an intermediate character between dynamic random-access memory (DRAM) and NAND flash in terms of latency and memory density. It is used for storage class memory (SCM). Resistive random-access memory (RRAM) [3,4,5,6,7,8,9,10], in which many materials can be used, has various properties. Interface type RRAM has good variability with multi-level cell (MLC). However, it has a slow switching speed [7]. Filamentary type RRAM has fast switching speed and good endurance. Unfortunately, its variation of resistance state is not good [11]. Among a lot of resistive switching materials, such as oxides [12,13,14,15,16,17], nitrides [18,19,20], organic materials [21] and 2D materials [22], oxides such as HfO2, TaOx, TiO2, and Al2O3 are good candidates in terms of the endurance, stability, repeatability and reproducibility. The bilayer structure of oxide can enhance resistive switching performances. For example, HfO2/Al2O3 dielectric stack on silicon can reduce operation current level [23]. The uniformity of cell-to-cell and device-to-device is improved in a TiOx/Al2O3 bilayer dielectric stack [24]. HfO2/Al2O3 bilayer device on TiN electrode can increase the nonlinearity of an I-V curve in the low-resistance state for high-density memory [25]. The stabilization is improved in a ZrO2(Y)/Ta2O5 bilayer device due to the electric field concentration of Ta nanocrystals [26]. TiO2 is one of the important resistive switching materials showing unipolar and bipolar resistive switching behaviors [27]. A lot of resistive switching behaviors including gradual and abrupt switching have been reported in various stacks. Al2O3 with large band gap can be used as a tunnel barrier layer as well as a resistive switching material [24,25].
RRAM devices show long-term and short-term plasticity to realize biological synapse for a neuromorphic system [28,29,30,31]. Many conductance states of memristor devices should be accurately and gradually controlled with the applied voltage bias. The state changed by the pulse can be maintained for a long time for off-chip learning in a neuromorphic system. To meet MLC with fine tuning, gradual set and reset switching in an RRAM device is essential. Many studies have been conducted on MLC improvement at the device level. Especially, optimizing a bilayer device is a powerful way to improve the gradual switching to achieve MLC. For example, synaptic properties are enhanced in a Ta2O5/HfO2 device by controlling the conductance gradually [32]. Recently, similar results have been found for TiOx/Al2O3 [33] and ZrO2/ZTO bilayer devices [34]. An additional insulating layer can act as series resistance so that gradual resistive switching behavior can be obtained during conductance transition.
In this work, a TiO2/Al2O3 bilayer device was fabricated for synaptic device application. Chemical and material compositions were verified by transmission electron microscopy (TEM) and energy-dispersive X-ray spectroscopy (EDS). We investigated resistive switching characteristics and conduction mechanisms, including initial state before forming low-resistance state (LRS) and high-resistance state (HRS). Gradual increase and decrease of conductivity were demonstrated by set pulse and reset pulse. Finally, MNIST pattern recognition accuracy was calculated by neuromorphic simulation composed of a single layer neural network by applying conductance of the device into weight of the network.

2. Materials and Methods

The Ti/TiO2/Al2O3/TiN memory device was prepared as follows—a 100-nm thick TiN bottom was deposited by DC sputtering. The Ti target was reacted with nitrogen (N2: 3 sccm and Ar: 20 sccm) to deposit TiN film in which the pressure was 1 mTorr and the power was 500 W.
A 2-nm thick Al2O3 layer was deposited with an atomic layer deposition (ALD) system. Al(CH3)3 (TMA) and H2O were used as precursors for Al2O3 deposition at a chamber temperature of 200 °C. A 16 nm thick TiO2 layer was then deposited by pulsed sputtering at room temperature. The Ti target was reacted with oxygen in plasma conditions. Flow rates of argon and oxygen were 12 and 8 sccm, respectively. Ti top electrode and Pt capping layer were deposited by DC sputtering and e-bema evaporator after the shadow mask was attached on the TiO2 layer. Each cell was separated by removing the mask.
The electrical properties of basic I-V curves were characterized in DC mode using a Keithley 4200-SCS semiconductor parameter analyzer (SPA, Keithley Instrumnets, Cleveland, OH, USA). Transient characteristics, potentiation, and depression curves were measured in pulse mode using a 4225-PMU ultrafast module. During measurements, a bias voltage and a pulse were applied to the Pt capping layer while maintaining the ground on the TiN bottom electrode.

3. Results and Discussion

Figure 1a shows a high-resolution TEM image of the Pt/Ti/TiO2/Al2O3/TiN memory device. The layer between TiO2 and TiN was thicker than the 2 nm target thickness of Al2O3, indicating that the TiON layer was created as an interfacial layer by reacting Al2O3 and TiN layers during Al2O3 deposition. A TiO2 layer about 16 nm thick as the main switching material, a Ti electrode, and a Pt electrode were clearly observed in the TEM image. Figure 1b shows the EDS line scan to detect elements as a function of the location, confirming the deposition of each layer.
Figure 2a shows the I-V curve of the forming process, which can activate the device before the set and reset processes. The Al2O3 layer has an important role in the conduction mechanism of the Pt/Ti/TiO2/Al2O3/TiN memory device before breakdown of the device. The Al2O3 layer in the initial state can act as a tunnel barrier. In a high field region (>2 V), FN tunneling can be a dominant mechanism in which electrons can tunnel more easily through the triangular barrier induced by a high electric field (Figure 2b). The FN tunneling equation is shown below [35]:
J = q 3 E 2 8 π hq ϕ B exp [ 8 π ( 2 q m * ) 3 h E ] ϕ B 3 / 2
where h is plank constant, q is electronic charge, m* is the effective electron mass, and E is the electric field. Thus, the straight line in ln(I/V2) versus 1/V at more than 2 V means this region follows FN tunneling as shown in Figure 2c.
Figure 3a shows set and reset processes of the Ti/TiO2/Al2O3/TiN memory device after electroforming process. The set process occurred under a positive bias in which the conducting defect (oxygen vacancies) was induced in both layers. Here, the device did not need a specific compliance current for the set process due to self-compliance. The self-compliance occurred with series resistance. In this device, TiON and Al2O3 layers could act as the series resistance. The set process was progressively completed. The oxygen ions with a negative charge moved toward a Ti top electrode with a positive bias. Oxygen vacancies were increased in the TiO2 layer, leading to an increase in the conductance of the Ti/TiO2/Al2O3/TiN system. On the other hand, the reset process happened in a negative bias. The reset process induced a decrease in the conductance gradually. When a negative bias was applied, oxygen vacancies recombined with oxygen ions from the top electrode. To reveal conduction mechanisms of LRS and HRS, I-V fitting processes were conducted. Firstly, we checked FN tunneling for LRS and HRS. However, the fitting accuracy was very low (now shown here). It indicated that large conducting defects existed at both insulating layers after the forming process. Thus, the Al2O3 layer no longer served as a strong tunnel barrier. The possible conduction model of HRS was trap-assisted tunneling (TAT). The TAT model equation is shown below [36]:
J exp [ 4 2 q m ) 3 h E ] ϕ t 3 / 2
where m is the effective mass in the insulator and ϕ t is the energy level of the defect state. The conduction of the electron is strongly dependent on the trap concentration that can be induced by the applied bias. In other words, traps induced in the thin films by the electroforming process can affect this TAT. Figure 3b shows Ln(I) versus 1/V for TAT fitting. It can be seen that the HRS curve fits the TAT well within a voltage range greater than 1.2 V. Figure 3c shows a log–log fitting of I-V in the LRS. Linear fitting means that LRS follows Ohmic conduction and conducting defects could be connected to top and bottom electrodes via two insulators.
Figure 4a shows the transient characteristics of Ti/TiO2/Al2O3/TiN memory device by set and reset pulses. Set and reset pulse voltages were 3 V and −2 V, respectively. The set process occurred during the rising time when the pulse voltage was raised and the reset process was completed after the voltage reached −2 V. Read pulses were added before and after the main switching pulses to check whether a set or reset had occurred. A read voltage of 0.3 V was used so that it would not impact set or reset switching ranges. Increase and decrease of the current were clearly observed after applying set and reset pulses.
Multiple conductances are needed to mimic biological synapses and implement a hardware-based neuromorphic system. Figure 4b shows conductance modulation by pulse inputs to emulate the potentiation and depression of a biological synapse. As the same pulse with a voltage of 4 V was repeatedly applied to the device 49 times, the conductance gradually increased from the HRS to the LRS. Although it tended to decrease after applying the set pulse at a few points, the conductance generally increased and potentiation was successfully emulated. On the other hand, multiple pulse inputs with a voltage of −1.35 V decreased the conductance gradually for the potentiation.
Next, MNIST pattern recognition accuracy was determined when the device’s conductance was used as the weight of the synapse in the neuromorphic system. Figure 5a shows a neural network composed of a single layer network (784 × 10) for the neuromorphic simulation. This neuromorphic simulation was based on off-chip learning. The input neuron was the MINST binary dataset with 28 × 28 pixel images and 10 output neurons to classify 10 classes. Two neuron nodes were fully connected to the synapse. Conductance values from a single cell device were used as weights of synapse including exhibitory and inhibitory ones. Here, it was assumed that all cells in the array had the same characteristics as the measured single cell. The recognition rate might decrease due to variation when an actual array was configured. The weight update of the device was conducted with a vector-matrix multiplication operation. Figure 5b shows the pattern recognition accuracy as a function of the number of training images. As training continued, the overall recognition rate improved. The maximum recognition rate was about 88.8%. The weight of synapse can be optimized in the neural network with training. Thus, an incomplete conductance update will eventually adversely affect the recognition rate.

4. Conclusions

In summary, a TiO2-based RRAM device with a thin Al2O3 layer was fabricated for neuromorphic application. Pt/Ti/TiO2/Al2O3/TiN stack was verified by TEM and EDS analyses. The conduction mechanism of an initial state can be explained by FN tunneling in the high voltage regime. Gradual set and reset switching were obtained under a DC sweep mode. HRS and LRS followed the TAT model and Ohmic conduction. Transient characteristics were well observed by applying set pulse and reset pulse. The conductance is tunable by applying set pulse and reset pulse, which can act as a synaptic device in a hardware based neuromorphic system. Finally, we evaluated the pattern recognition accuracy by constructing a single layer neural network including the measured conductance as synaptic weight. An accuracy of 88.8% was achieved due to a multi-level conductance.

Author Contributions

H.R. conducted electrical measurements and wrote the manuscript; S.K. designed the experiment and supervised the study. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade, Industry & Energy (MOTIE) of the Republic of Korea (no.20194030202320) and was supported by Nano·Material Technology Development Program through the National Research Foundation of Korea(NRF) funded by the Ministry of Science, ICT and Future Planning (2009-0082580).

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. Chien, W.-C.; Chang, Y.-C.; Tsou, Y.-T.; Kuo, S.-Y.; Chang, C.-R. STT-DPSA: Digital PUF-Based Secure Authentication Using STT-MRAM for the Internet of Things. Micromachines 2020, 11, 502. [Google Scholar] [CrossRef]
  2. Park, J. Neuromorphic Computing Using Emerging Synaptic Devices: A Retrospective Summary and an Outlook. Electronics 2020, 9, 1414. [Google Scholar] [CrossRef]
  3. Lanza, M.; Wong, H.-S.P.; Pop, E.; Ielmini, D.; Strukov, D.; Regan, B.C.; Larcher, L.; Villena, M.A.; Yang, J.J.; Goux, L.; et al. Recommended Methods to Study Resistive Switching Devices. Adv. Electron. Mater. 2018, 5, 1800143. [Google Scholar] [CrossRef] [Green Version]
  4. Linn, E.; Rosezin, R.; Kügeler, C.; Waser, R. Complementary resistive switches for passive nanocrossbar memories. Nat. Mater. 2010, 9, 403–406. [Google Scholar] [CrossRef] [PubMed]
  5. Kim, S.; Jung, S.; Kim, M.-H.; Chen, Y.-C.; Chang, T.-C.; Ryoo, K.-C.; Cho, S.; Lee, J.-H.; Park, B.-G. Scaling Effect on Silicon Nitride Memristor with Highly Doped Si Substrate. Small 2018, 14, 1704062. [Google Scholar] [CrossRef]
  6. Zhou, F.; Chang, Y.F.; Fowler, B.; Byun, K.; Lee, J.C. Stabilization of multiple resistance levels by current-sweep in SiOx-based resistive switching memory. Appl. Phys. Lett. 2015, 106, 063508. [Google Scholar]
  7. 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]
  8. Maikap, S.; Banergee, W. In Quest of Nonfilamentary Switching: A Synergistic Approach of Dual Nanostructure Engineering to Improve the Variability and Reliability of Resistive Random-Access-Memory Devices. Adv. Electron. Mater. 2020, 6, 2000209. [Google Scholar] [CrossRef]
  9. 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]
  10. Ryu, H.; Kim, S. Pseudo-Interface Switching of a Two-Terminal TaOx/HfO2 Synaptic Device for Neuromorphic Applications. Nanomaterials 2020, 10, 1550. [Google Scholar] [CrossRef]
  11. Ryu, H.; Kim, S. Synaptic Characteristics from Homogeneous Resistive Switching in Pt/Al2O3/TiN Stack. Nanomaterials 2020, 10, 2055. [Google Scholar] [CrossRef]
  12. Choi, J.; Kim, S. Nonlinear Characteristics of Complementary Resistive Switching in HfAlOx-Based Memristor for High-Density Cross-Point Array Structure. Coatings 2020, 10, 765. [Google Scholar] [CrossRef]
  13. Han, G.; Chen, Y.; Liu, H.; Wang, D.; Qiao, R. Impacts of LaOx Doping on the Performance of ITO/Al2O3/ITO Transparent RRAM Devices. Electronics 2021, 10, 272. [Google Scholar] [CrossRef]
  14. Chen, Y.-C.; Lin, C.-C.; Chang, Y.-F. Post-Moore Memory Technology: Sneak Path Current (SPC) Phenomena on RRAM Crossbar Array and Solutions. Micromachines 2021, 12, 50. [Google Scholar] [CrossRef] [PubMed]
  15. 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]
  16. Zhao, X.; Song, P.; Gai, H.; Li, Y.; Ai, C.; Wen, D. Li-Doping Effect on Characteristics of ZnO Thin Films Resistive Random Access Memory. Micromachines 2020, 11, 889. [Google Scholar] [CrossRef]
  17. Pérez, E.; González Ossorio, Ó.; Dueñas, S.; Castán, H.; García, H.; Wenger, C. Programming Pulse Width Assessment for Reliable and Low-Energy Endurance Performance in Al:HfO2-Based RRAM Arrays. Electronics 2020, 9, 864. [Google Scholar] [CrossRef]
  18. Cho, S.; Kim, S. Emulation of Biological Synapse Characteristics from Cu/AlN/TiN Conductive Bridge Random Access Memory. Nanomaterials 2020, 10, 1709. [Google Scholar] [CrossRef]
  19. Kim, S.; Kim, H.; Jung, S.; Kim, M.H.; Lee, S.; Cho, S.; Park, B.G. Tuning resistive switching parameters in Si3N4-based RRAM for three-dimensional vertical resistive memory applications. J. Alloys Compd. 2016, 663, 419–423. [Google Scholar] [CrossRef]
  20. Jung, J.; Bae, D.; Kim, S.; Kim, H.-D. Reduced Operation Current of Oxygen-Doped ZrN Based Resistive Switching Memory Devices Fabricated by the Radio Frequency Sputtering Method. Coatings 2021, 11, 197. [Google Scholar] [CrossRef]
  21. Patil, H.; Kim, H.; Rehman, S.; Kadam, K.D.; Aziz, J.; Khan, M.F.; Kim, D.-K. Stable and Multilevel Data Storage Resistive Switching of Organic Bulk Heterojunction. Nanomaterials 2021, 11, 359. [Google Scholar] [CrossRef] [PubMed]
  22. Shen, Z.; Zhao, C.; Qi, Y.; Mitrovic, I.Z.; Yang, L.; Wen, J.; Huang, Y.; Li, P.; Zhao, C. Memristive Non-Volatile Memory Based on Graphene Materials. Micromachines 2020, 11, 341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Kim, S.; Chen, J.; Chen, Y.-C.; Kim, M.-H.; Kim, H.; Kwon, M.-W.; Hwang, S.; Ismail, M.; Li, Y.; Miao, X.-S.; et al. Neuronal dynamics in HfO x/AlO y-based homeothermic synaptic memristors with low-power and homogeneous resistive switching. Nanoscale 2019, 11, 237–245. [Google Scholar] [CrossRef] [Green Version]
  24. Banerjee, W.; Xu, X.; Lv, H.; Liu, Q.; Long, S.; Liu, M. Variability Improvement of TiOx/Al2O3 Bilayer Nonvolatile Resistive Switching Devices by Interfacial Band Engineering with an Ultrathin Al2O3 Dielectric Material. ACS Omega 2017, 2, 6888–6895. [Google Scholar] [CrossRef] [PubMed]
  25. Chand, U.; Huang, K.C.; Huang, C.Y.; Tseng, T.Y. Mechanism of Nonlinear Switching in HfO2-Based Crossbar RRAM With Inserting Large Bandgap Tunneling Barrier Layer. IEEE Trans. Electron. Dev. 2015, 62, 3665–3670. [Google Scholar] [CrossRef]
  26. Mikhaylov, A.; Belov, A.; Korolev, D.; Antonov, I.; Kotomina, V.; Kotina, A.; Gryaznov, E.; Sharapov, A.; Koryazhkina, M.; Kryukov, R.; et al. Multilayer Metal-Oxide Memristive Device with Stabilized Resistive Switching. Adv. Mater. Technol. 2020, 5, 1900607. [Google Scholar] [CrossRef]
  27. Jeong, D.S.; Schroeder, H.; Waser, R. Coexistence of Bipolar and Unipolar Resistive Switching Behaviors in a Pt/TiO2/Pt Stack. Electrochem. Solid State Lett. 2007, 10, G51. [Google Scholar] [CrossRef]
  28. Mikhaylov, A.; Pimashkin, A.; Pigareva, Y.; Gerasimova, S.; Gryaznov, E.; Shchanikov, S.; Zuev, A.; Talanov, M.; Lavrov, I.; Demin, V.; et al. Neurohybrid Memristive CMOS-Integrated Systems for Biosensors and Neuroprosthetics. Front. Mol. Neurosci. 2020, 14, 358. [Google Scholar] [CrossRef]
  29. Kuzum, D.; Yu, S.; Wong, H.-S.P. Synaptic electronics: Materials, devices and applications. Nanotechnology 2013, 24, 382001. [Google Scholar] [CrossRef]
  30. Graves, C.E.; Li, C.; Sheng, X.; Miller, D.; Ignowski, J.; Kiyama, L.; Strachan, J.P. In-Memory Computing with Memristor Content Addressable Memories for Pattern Matching. Adv. Mater. 2020, 32, 2003437. [Google Scholar] [CrossRef]
  31. Xia, Q.; Yang, J.J. Memristive crossbar arrays for brain-inspired computing. Nat. Mater. 2019, 18, 309–323. [Google Scholar] [CrossRef] [PubMed]
  32. Ryu, J.H.; Mahata, C.; Kim, S. Long-term and short-term plasticity of Ta2O5/HfO2 memristor for hardware neuromorphic application. J. Alloys Compd. 2021, 850, 156675. [Google Scholar] [CrossRef]
  33. Kim, T.H.; Nili, H.; Kim, M.H.; Min, K.K.; Park, B.G.; Kim, H. Reset-voltage-dependent precise tuning operation of TiOx/Al2O3 memristive crossbar array. Appl. Phys. Lett. 2020, 117, 152103. [Google Scholar] [CrossRef]
  34. Ismail, M.; Abbas, H.; Choi, C.; Kim, S. Controllable analog resistive switching and synaptic characteristics in ZrO2/ZTO bilayer memristive device for neuromorphic systems. Appl. Surf. Sci. 2020, 529, 147107. [Google Scholar] [CrossRef]
  35. Chang, Y.F.; Chen, P.Y.; Fowler, B.; Chen, Y.T.; Xue, F.; Wang, Y.; Zhou, F.; Lee, J.C. Intrinsic SiOx-based unipolar resistive switching memory. II. Thermal effects on charge transport and characterization of multilevel programing. J. Appl. Phys. 2014, 116, 043709. [Google Scholar] [CrossRef]
  36. Yan, X.; Zhou, Z.; Ding, B.; Zhao, J.; Zhang, Y. Superior resistive switching memory and biological synapse properties based on a simple TiN/SiO2/p-Si tunneling junction structure. J. Mater. Chem. C 2017, 5, 2259–2267. [Google Scholar] [CrossRef]
Figure 1. (a) High resolution transmission electron microscopy (TEM) image and (b) Energy-dispersive spectroscopy (EDS) line scan of a Pt/Ti/TiO2/Al2O3/TiN device.
Figure 1. (a) High resolution transmission electron microscopy (TEM) image and (b) Energy-dispersive spectroscopy (EDS) line scan of a Pt/Ti/TiO2/Al2O3/TiN device.
Metals 11 00440 g001
Figure 2. (a) Forming I-V characteristics, (b) Fowler-Nordheim (FN) tunneling illustration in energy band structure, (c) Ln(I/V2) versus 1/V.
Figure 2. (a) Forming I-V characteristics, (b) Fowler-Nordheim (FN) tunneling illustration in energy band structure, (c) Ln(I/V2) versus 1/V.
Metals 11 00440 g002
Figure 3. (a) Set and reset I-V characteristics, (b) Ln(I) versus 1/V fitting for high-resistance state (HRS), (c) Log-log scale for Ohmic conduction in the low-resistance state (LRS).
Figure 3. (a) Set and reset I-V characteristics, (b) Ln(I) versus 1/V fitting for high-resistance state (HRS), (c) Log-log scale for Ohmic conduction in the low-resistance state (LRS).
Metals 11 00440 g003
Figure 4. (a) Set and reset transient characteristics including set pulse, reset pulse, and read pulses; (b) Potentiation and depression curve as a synaptic device.
Figure 4. (a) Set and reset transient characteristics including set pulse, reset pulse, and read pulses; (b) Potentiation and depression curve as a synaptic device.
Metals 11 00440 g004
Figure 5. (a) Single layer neural network including input neuron, synapse, and output layer, (b) accuracy as a function of training epoch.
Figure 5. (a) Single layer neural network including input neuron, synapse, and output layer, (b) accuracy as a function of training epoch.
Metals 11 00440 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ryu, H.; Kim, S. Gradually Tunable Conductance in TiO2/Al2O3 Bilayer Resistors for Synaptic Device. Metals 2021, 11, 440. https://0-doi-org.brum.beds.ac.uk/10.3390/met11030440

AMA Style

Ryu H, Kim S. Gradually Tunable Conductance in TiO2/Al2O3 Bilayer Resistors for Synaptic Device. Metals. 2021; 11(3):440. https://0-doi-org.brum.beds.ac.uk/10.3390/met11030440

Chicago/Turabian Style

Ryu, Hojeong, and Sungjun Kim. 2021. "Gradually Tunable Conductance in TiO2/Al2O3 Bilayer Resistors for Synaptic Device" Metals 11, no. 3: 440. https://0-doi-org.brum.beds.ac.uk/10.3390/met11030440

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