Next Article in Journal
CeO2 Nanomaterials from Diesel Engine Exhaust Induce DNA Damage and Oxidative Stress in Human and Rat Sperm In Vitro
Previous Article in Journal
Magnetic Ionic Liquid Nanocatalyst to Improve Mechanical and Thermal Properties of Epoxy Nanocomposites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Neuro-Transistor Based on UV-Treated Charge Trapping in MoTe2 for Artificial Synaptic Features

1
Department of Electrical Engineering, Sejong University, 209 Neungdong-ro, Gwangjin-gu, Seoul 05006, Korea
2
School of electronics Engineering, Chungbuk National University, Cheongju 28644, Korea
3
Department of Convergence Engineering for Intelligent Drone, Sejong University, Seoul 05006, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2020, 10(12), 2326; https://doi.org/10.3390/nano10122326
Submission received: 4 November 2020 / Revised: 18 November 2020 / Accepted: 21 November 2020 / Published: 24 November 2020
(This article belongs to the Section 2D and Carbon Nanomaterials)

Abstract

:
The diversity of brain functions depend on the release of neurotransmitters in chemical synapses. The back gated three terminal field effect transistors (FETs) are auspicious candidates for the emulation of biological functions to recognize the proficient neuromorphic computing systems. In order to encourage the hysteresis loops, we treated the bottom side of MoTe2 flake with deep ultraviolet light in ambient conditions. Here, we modulate the short-term and long-term memory effects due to the trapping and de-trapping of electron events in few layers of a MoTe2 transistor. However, MoTe2 FETs are investigated to reveal the time constants of electron trapping/de-trapping while applying the gate-voltage pulses. Our devices exploit the hysteresis effect in the transfer curves of MoTe2 FETs to explore the excitatory/inhibitory post-synaptic currents (EPSC/IPSC), long-term potentiation (LTP), long-term depression (LTD), spike timing/amplitude-dependent plasticity (STDP/SADP), and paired pulse facilitation (PPF). Further, the time constants for potentiation and depression is found to be 0.6 and 0.9 s, respectively which seems plausible for biological synapses. In addition, the change of synaptic weight in MoTe2 conductance is found to be 41% at negative gate pulse and 38% for positive gate pulse, respectively. Our findings can provide an essential role in the advancement of smart neuromorphic electronics.

Graphical Abstract

1. Introduction

Various attempts have been made to emulate the key functions of biological neurons in artificial intelligence and neuromorphic electronic devices for the next generation [1,2,3,4]. In a complementary metal–oxide–semiconductor (MOS) transistor, the von-Neumann architecture is insufficient to control the conventional constraints of computational difficulty and associated power consumptions [5]. Meanwhile, recent neuromorphic technology faces the challenges of high circuitry complexity, low efficiency, and difficulties of low-dimensional size of devices. However, recent developments in imaging techniques have led to the discovery of several significant and important details about information processing inside the human nervous systems which is precisely considered in electronic devices to develop more genuine and proficient human brain-inspired machine interfaces [6]. The scientists and physicians are encouraged to diagnose and treat various brain diseases and disorders by using these artificial machines. The main purpose of neuromorphic engineering and the human-brain scheme is to realize the function of brain which manages billions (~1011) of information processing units connected through trillions (~1015) of synapses. Neurons and synapses are basic units of brain memory and information processing which emulate neuromorphic responses via nano-electronic devices to achieve the objectives of machine learning, robust computational development, and short-term and long-term plasticity. It can also be helpful to design the artificial sensory systems which demonstrate the growth and advancement of diseases and disorders and likely effects of drugs, and autonomous and intelligent robots that will ultimately simplify difficult surgical techniques. To emulate the analog computing arrangement of one synapse, at least 10 complementary MOS transistors are justified [7]. Considering these architectures demand strongly interlinked elements, the scalability of these systems becomes a primary challenge [8]. Furthermore, compacting with the aforementioned issue, a large amount of power (>1000 W) is utilized by such platforms to execute tasks that the human brain completes with <4 W [9]. Therefore, this emphasizes the urgent demand to reveal the electronic devices that can mimic the synaptic or neuronal response which are attuned with the human brain. It is evident that the information collected through different human body parts/organs transform into electrical impulses. Later, these impulses are coped and processed by release of neurotransmitters inside the human brain, an intricate network of neurons [10]. Thereafter, the decisive results of the neural information processing is categorized into sort-term action (logic operation) and long-term changes (memory storage) [11]. Intentionally, it can be associated with two different mechanisms having different time constants in order to mimic both the short-term and long-term plasticity of the human brain [12,13]. However, the time constant for logic operation is principally controlled by the transit time of charge carriers (electrons/holes) from the source to the drain of FETs. The transit time of carriers is critical to determine the memory retention which mainly depends upon the conductance of channel material (MoTe2), charging and dis-charging of carriers and gate oxide. Several techniques have been adopted to mimic the neuromorphic transmission and human behavior in artificial devices. As such, transistors, memristors, and other devices are being used for this purpose and bi-state spin valve junctions could be the best choice for computing purposes [14,15,16,17,18,19]. Although the two-terminal devices have commendably demonstrated the synaptic systems with a facile rout of fabrication, they face some issues when performing weight training and signal transmission [20,21,22].
To account for these issues, the three-terminal devices like FETs have been introduced to mimic the synapse called “synaptic transistor”. Numerous oxide materials, carbon nanotubes, nanoparticles, and ferroelectrics have been investigated to emulate the neural electronics [23,24,25,26]. Recently, promising two-dimensional (2D) transition metal dichalcogenides (TMDCs) have shown a huge potential to be utilized in synaptic transistors and in a recent study a 2D-TMDCs disulphide group (MoS2 and ReS2) was adopted for neuromorphic computing [27,28]. However, less attention has been paid to the telliride (Te) group of 2D-TMDCs. Generally, for neuromorphic transistors researchers use special high-k dielectrics (HfO2, Al2O3 and TaOx) substrates for charge trapping purpose. However, we used simple Si/SiO2 substrates and subsequently the 2D material (MoTe2) is treated with UV in air to enhance the trapping mechanism. Our technique is proposed to be easy and novel for electron trapping in neuro-transistors.

2. Materials and Methods

First, we made large patterns on p++-Si/SiO2 substrates by photolithography and deposited Cr/Au (5/30 nm) by thermal evaporation to obtain big patterns. Initially, MoTe2 flaks were exfoliated in on scotch tape and transferred on polydimethylsiloxane (PDMS) stamp. Later, on PDMS MoTe2 was treated with DUV for 10 sec in ambient condition to encourage trapping states and finally transferred on SiO2 substrates with the help of a micromanipulator. At this moment, the treated side of MoTe2 was coincided with SiO2. Subsequently, using electron beam lithography, EL-9 and PMMA were spin-coated on the devices to make fine source and drain terminals. Lastly, Cr/Au, 7/80 nm, were deposited in thermal evaporation chamber and the lift-off was completed in acetone/methanol. Electrical measurements were performed by using Keithley-2400 (Keithley Instruments Inc., Solon, OH, USA) and Picoammeter Keithley-6485 (Keithley Instruments Inc., Solon, OH, USA). The pulsed measurements were performed using the Keithley-4200 (Keithley Instruments Inc., Solon, OH, USA) semiconductor characterization system (SCS) with additional pulse measurement units (PMU). The Raman spectrum was acquired using a Renishaw microspectrometer (Renishaw plc., Wotton-under-Edge, UK) with a laser excitation wavelength of 514 nm. Moreover, DUV light of 220 nm was used with an optical power of 11 mW cm−2.

3. Results and Discussion

The schematic representation and final optical microscope image of MoTe2 field effect transistor (FET) are demonstrated in Figure 1a,b, respectively. The MoTe2 transistor is supposed to replicate the neuro-biological dynamics in a chemical synapse. In Figure 1b, the MoTe2 flake is transferred on a SiO2 substrate and the Cr/Au electrodes are deposited by thermal evaporation. Further detailed information is given in the experimental section. Raman spectroscopy is the best tool to characterize the band and crystal structure of materials. In Figure 1c, we showed the Raman spectrum of MoTe2 which demonstrated the few frequency modes at different positions. We characterized Raman spectroscopy MoTe2 with laser light having wavelength of 514 nm. The power intensity of laser light is kept at 1 mW/cm2. Thereby, three modes are observed, namely, A1g, E2g1 and B2g1 at ~174.6 cm−1, ~235.9 cm−1, and ~291.2 cm−1, respectively. Thus, our results are almost consistent with previous reports and confirmed that MoTe2 flake is multilayer in thickness [29]. For further confirmation we performed the atomic force microscopy (AFM) and investigate the exact thickness (~15 nm) of MoTe2 flake. The AFM image and height profile of the flake is shown in Figure S1a,b. The synapses have a fundamental role in the propagation of information in the central nervous system. In this type of structures, plasma membrane of the presynaptic neuron (signal-carrying neuron) comes very close to the plasma membrane of the postsynaptic neuron (target neuron). The schematic presentation of synaptic response where release and absorption of neurotransmitter are controlled by action potentials is shown in Figure 1d.
We performed our all electrical measurements in ambient conditions where the drain current acts as the EPSC, which is supposed to be the device conductance as a function of time. Moreover, the pre- and post-synaptic pulses have been applied to the back-gate terminal (p-Si/SiO2) and the MoTe2 channel with the trapped states behaves as an analog to the synaptic fluid where neurotransmitters are released and absorbed. However, in Figure 2a we measured the transfer characteristics of MoTe2 transistor in forward and revers sweep from −35 V to +35 V and then from +35 V to −35 V. Here, we found n-type semiconductor by fixing Vds = 500 mV owing gate voltage gap size near 1 V. It is evident from Figure 2a that we observed a promising hysteresis loop in transfer characteristics at 0 ms holding/delay time of the gate sweep, which is commonly attributed to the donor like trap/defect states formed at MoTe2/SiO2 interface and also due to top surface adsorbents/water dipoles on MoTe2 [30,31]. Also, as we increased the holding/delay time in transfer characteristics the size of hysteresis loop becomes large, as shown in Figure 2a. It is noted in a previous report that at short delay time, the charges are trapped only in shallow levels and the hysteresis is found to be minor. However, after increasing the delay time, the charges start being trapped in deep levels due to which a large number of electrons are trapped in both the shallow and deep levels. Therefore, the hysteresis loop become significant at large delay time [32]. We proposed that by increasing the delay time, we may improve the effect of trapping states on charge carries. Therefore, to see in depth we calculate the mobility of MoTe2 transistor at different holding/delay times. The mobility of MoTe2 device is obtained using the following Equation (1):
μ = L WC g V ds dI ds dV g
where dI ds / dV g is the slope of the transfer curve in the linear region, Cg is the gate capacitance (~115 aF μm−2) of Si/SiO2 substrates Vds is the drain voltage, L is the channel length, and W is the width. In Figure 2b, the mobility is decreased from (6.9, 5.5, 4.8 and 4.2 cm−2/Vs) as the delay time of gate sweep is increased from (0, 100, 300, and 500 ms) which endorsed the enhanced influence of trapping states with increased delay time. To estimate the equivalent trap density (NTrap) at the MoTe2/SiO2 interface the following Equation (2) is used:
N(Trap) = Cg[Vth(Forward) − Vth(Reverse)]/q
where Cg = ~115 aF μm−2 is the gate capacitance of back gate, Vth is the threshold voltage (Forward and Reverse) and q is the electric charge. The NTrap of n-FET based MoTe2 is 1.07 × 1012, 1.2 × 1012, 1.5 × 1012 and 1.95 × 1012 cm−2 for 0, 100, 300, and 500 ms holding time by gate weep as shown in Figure 2c. To get additional understanding into the charge trapping mechanism, the hysteresis loops of the transistor are also measured at different back gate sweeping ranges (−35 V to +35 V, −25 V to +25 V and −15 V to +15 V) as shown in Figure S2a (Supplementary Information). The variation in the hysteresis loops at different sweep rates designates the existence of slow charge trapping in defect states. The observed change in ION of the transistor is usually due to the random trapping and de-trapping of the charge carriers in the MoTe2 or SiO2. Also, we demonstrated the effect of UV on the bottom side of MoTe2 flake by measuring the transfer curves of FET. We can see in Figure 2d that the hysteresis loop becomes large compared to the pristine one. In the pristine device the hysteresis loop is ΔV = ~5 V but after UV treatment the loop is improved to ΔV = ~15 V.
To explain further, the trapping and de-trapping of the charge carriers are demonstrated by the energy band diagram as shown in Figure 3. According to Figure 3a when the applied back gate voltage is negative (Vg > 0, off state), the induced electric field in SiO2 (dielectric) will encourage the trapped electrons to be released from trapped states into the n-semiconductor MoTe2 channel and produce a large current density, and as a result the conductivity of MoTe2, will be enhanced. On the other hand, in Figure 3b, if we applied positive pulse of back gate voltage (Vg < 0, on state), then the induced electrical field will deplete the electrons from the MoTe2, thereby reducing the channel conductivity of MoTe2. This situation give rise to the shift in threshold voltage of the transistor, which is consistent with the previous report [28]. The variation of channel conductivity can be interlinked with the synaptic plasticity of neurons. Our work is mainly based on the trapping and de-trapping of charge carriers in MoTe2 FETs where we need to examine their behavior to emulate the human synaptic functions and learning rules like LTP/LTD and various learning mechanisms such as SADP and STDP. These activities of neuromorphic function are based on different single, double, and successive positive and negative back gat pulses. Therefore, the back gate behaves like the pre-synaptic terminal and controls the conductance of the MoTe2, which acts like a synapse.
Subsequently, the drain acts as the post-synaptic terminal. Hence, the hysteresis effect in the transfer characteristics of the MoTe2 transistors is used to mimic the vibrant plasticity. Also, we can control the gate pulses to trap and de-trap the electrons from the MoTe2 and modify the channel conductance to describe the EPSC. As given in Figure 4a, a significant change is observed in the threshold voltage of MoTe2 transistor before and after applying a train of 50 pulses to the back-gate terminal. The amplitude and width of the pulse are −15 V and 300 ms, respectively, for potentiation (red). In addition, to recover the initial position of the threshold voltage, a train of 50 pulses are applied with same width and amplitude in opposite polarity +10 V to get a depression (black). However, this kind of measurement, when the pulse is applied to the gate terminal, can control the different repeated and reversible conductance states of MoTe2. Moreover, when a single voltage pulse (single spike response) is given the synapse can be either excitatory (negative gate pulse) or inhibitory (positive gate pulse) in nature. Nevertheless, it depends upon the tendency of increments in the post-synaptic current to threshold or decrement in the post-synaptic current away from the threshold. For only single gate voltage pulse, we give the height of gate pulse is −8 V by keeping the pulse width of 25 ms at 500 mV bias voltage. At this moment, the EPSC current reaches up to ~15 μA and then gradually comes back to the original position after the pulse ends, as shown in Figure 4b. The increase of EPSC current in MoTe2 channel is due to the release of trapped electrons into the MoTe2 channel from trapping states in the SiO2/MoTe2 interface. Here, the amplitude of the PSC triggered by the presynaptic spike is related to the synaptic strength/weight in the biological synapse. Similarly, when a positive gate voltage pulse is applied with height pulse = +8 V reduces the IPSC current reduced to ~−17 μA, attributed to the injected electrons from MoTe2 channel to the SiO2/MoTe2 interface which can emulate the IPSC behavior of biological synapse, as shown in Figure 4c.
The merit features of synaptic plasticity are that of long-term potentiation (LTP) and long-term depression (LTD). These features can play essential role in the information processing of the biological nervous system [33,34]. In neuro-electronic synapse, LTP and LTD can be mimicked by using multiple pulses (spikes) of high frequency and in this way the synaptic transmission can be increased or decreased, respectively as shown in Figure S2b,c. Here, a train of 80 gate voltage pulses are given to the gate as pre-synaptic input for the investigation of drain current (post- synaptic current). It is found that a positive gate pulse heights (+5, +10 and +15 V) and a negative gate pulse height (−5, −10 and −15 V) generate negative and positive exponential changes in post-synaptic current, respectively, which indicate the LTP and LTD actions of a synapse. Further, the time constants for potentiation (at Vg = −4 V) and depression (at Vg = +4 V) responses are calculated from the experimental data which is found to be 2.9 and 2.7 s, respectively, as shown in Figure 4d. The conductance variation of MoTe2 channel is asymmetric with sudden increase and decrease in early potentiation and depression spikes and then it became saturated. We have also investigated the post-synaptic current by applying the different heights of gate voltage pulses, thereby the post-synaptic current experiences an increase/decrease in the channel conductance of MoTe2 with an increase/decrease in the gate height pulse, respectively, as shown in Figure 5a. Figure 5a describes the emulation of spike amplitude-dependent (SADP) of the biological synapse, since when the height applied gate voltage increases, more trapping states become activated and start to trap more electrons at the interface. The STDP in neuromorphic systems is another important function for the learning and memory process of a human brain. Thus, by increasing the time between two consecutive action potentials of pre-synaptic and post-synaptic spikes, the synaptic weight strength between two neurons can be changed [35]. Therefore, in our work the transistors mimicked the indirect STDP with a 41% change in MoTe2 channel conductance for negative gate pulse (Vg = −4 V) and a 38% change for positive gate pulse (Vg = +4 V), as shown in Figure 5b. Hence, this change is attained by paired gate pulses with a specific time period interval [36]. It is also important to note that as Δt increases, the change in channel conductance of MoTe2 become saturated because trapped electrons take a lot of time to be de-trapped. We studied the time constants for potentiation and depression pulses, which are 0.6 and 0.9 s, respectively, which seems plausible with the response times of biological synapses [35]. The time constant of relaxation for the potentiation and depression pulses are estimated by an exponential fit to the STDP data by using the following Equation (3):
Δ W α { exp ( Δ t τ + ) ,   if   Δ t 0 exp ( Δ t τ ) ,   if   Δ t 0
Furthermore, we observed the effect of double spike response of the MoTe2 synaptic transistor through the application of pair pre-synaptic pulses on the gate terminal. The paired pulsed facilitation/depression (PPF/PPD) is analog short-term synaptic plasticity, which is important to elaborate the time-based information in auditory or visual signals [37].
Therefore, EPSC or IPSC response can be observed at drain when two consecutive pre-synaptic spikes are applied to the gate terminal. As such, the enhancement or depression of PSC will occur because the back diffusion of the electron is not completed before the arrival of the second gate pulse. Thus, the post-synaptic current due to second pulse is lower than the first gate pulse. Substantially, as time increases the change in MoTe2 conductance is reduced due to the second pulse A2 or becomes equal to the first pulse A1 for higher time interval as shown in Figure 5c. Hereafter, we describe the PPF index as:
[(A2 − A1)/A1] × 100(%)
Here, we have plotted the PPF index (%) versus interval from a positive gate voltage (Vg = 4 V) and pulse width = 25 ms) as shown in Figure 5d. Therefore, it is found that as we increased the time interval (Δt) between the pair of applied gate pulses the A2/A1 (PPF ratio) decreased. It is fitted by the exponential relation:
PPF = 1 A 1 e Δ t τ 1
where A1 is the initial facilitation value and τ1 is the time relaxation of the phase. The calculation of the plasticity, PPF and STDP related time constants of relaxation can be used to initiate circuit level simulations for spiking neural networks. The ability of electrons trapping/de-trapping at interface can open a new avenues in future neuromorphic applications.

4. Conclusions

In summary, we demonstrated the tailoring of hysteresis loops in n-type MoTe2 FETs to emulate the biological synapses. We investigated that by modulating the amplitude and polarity of gate voltage pulses, we can precisely capture the excitatory or inhibitory nature of neurotransmitter. The trapping and de-trapping of electron from trapping states, formed at SiO2/MoTe2 interface, exhibit relaxation time constants and allow for the emulation of both STP and LTP which is an essential characteristic of learning and memory in biological synapses. In addition, we discussed the STDP, PFF, potentiation, and depression to replicate the human brain for advanced neuromorphic computing. The use of 2D MoTe2 as channel material on SiO2 substrate is a next-generation semiconducting material which offers outstanding device dimensions down to a single atomic thickness with diverse synaptic applications.

Supplementary Materials

Supplementary materials can be found at https://www.mdpi.com/2079-4991/10/12/2326/s1.

Author Contributions

S.R. and M.F.K. conceived and designed the experiments; M.F.K., M.K.R., H.P. and S.A.K. performed the experiments and analyzed the data; S.R., M.F.K. and H.K. choose the material and fabricated the devices; S.R., M.F.K., D.-k.K. and M.H.K. helped to write and review the manuscript. All authors have read and agree to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This research was supported by National Research Foundation of Korea (NRF), ICT (2020R1G1A1012022). This work is also supported by non-tenure track new faculty research fund (2020–2021) by Sejong University. This research was also supported by the MOTIE (Ministry of Trade, Industry & Energy (10080581, 20010574) and KSRC (Korea Semiconductor Research Consortium) support program for the development of the future semiconductor device.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Indiveri, G.; Douglas, R. Neuromorphic vision sensors. Science 2000, 288, 1189–1190. [Google Scholar] [CrossRef]
  2. Chen, S.; Lou, Z.; Chen, D.; Shen, G. An artificial flexible visual memory system based on an UV-motivated Memristor. Adv. Mater. 2018, 30, 1705400. [Google Scholar] [CrossRef]
  3. Lee, G.J.; Choi, C.; Kim, D.H.; Song, Y.M. Bioinspired artificial eyes: Optic components, digital cameras, and visual prostheses. Adv. Funct. Mater. 2018, 28, 1705202. [Google Scholar] [CrossRef]
  4. Sun, J.; Oh, S.; Choi, Y.; Seo, S.; Oh, M.J.; Lee, M.; Lee, W.B.; Yoo, P.J.; Cho, J.H.; Park, J.H. Optoelectronic Synapse Based on IGZO-Alkylated Graphene Oxide Hybrid Structure. Adv. Funct. Mater. 2018, 28, 1804397. [Google Scholar] [CrossRef]
  5. Taur, Y.; Buchanan, D.A.; Chen, W.; Frank, D.J.; Ismail, K.E.; Lo, S.-H.; Sai-Halasz, G.A.; Viswanathan, R.G.; Wann, H.-J.; Wind, S.J. CMOS scaling into the nanometer regime. Proc. IEEE 1997, 85, 486–504. [Google Scholar] [CrossRef]
  6. Brooks, R.A. Intelligence without representation. Artif. Intell. 1991, 47, 139–159. [Google Scholar] [CrossRef]
  7. Chicca, E.; Badoni, D.; Dante, V.; D’Andreagiovanni, M.; Salina, G.; Carota, L.; Fusi, S.; Del Giudice, P. A VLSI recurrent network of integrate-and-fire neurons connected by plastic synapses with long-term memory. IEEE Trans. Neural Netw. 2003, 14, 1297–1307. [Google Scholar] [CrossRef]
  8. Indiveri, G.; Liu, S.-C. Memory and information processing in neuromorphic systems. Proc. IEEE 2015, 103, 1379–1397. [Google Scholar] [CrossRef] [Green Version]
  9. Silver, D.; Huang, A.; Maddison, C.J.; Guez, A.; Sifre, L.; Van Den Driessche, G.; Schrittwieser, J.; Antonoglou, I.; Panneershelvam, V.; Lanctot, M. Mastering the game of Go with deep neural networks and tree search. Nature 2016, 529, 484–489. [Google Scholar] [CrossRef]
  10. Striedter, G.F. Neurobiology: A Functional Approach; Oxford University Press: Oxford, UK, 2016. [Google Scholar]
  11. Gerstner, W.; Kistler, W.M.; Naud, R.; Paninski, L. Neuronal Dynamics: From Single Neurons to Networks and Models of Cognition; Cambridge University Press: Cambridge, UK, 2014. [Google Scholar]
  12. Nestler, E.J. Molecular basis of long-term plasticity underlying addiction. Nat. Rev. Neurosci. 2001, 2, 119–128. [Google Scholar] [CrossRef]
  13. Zucker, R.S.; Regehr, W.G. Short-term synaptic plasticity. Annu. Rev. Physiol. 2002, 64, 355–405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Bhattacharjee, S.; Wigchering, R.; Manning, H.G.; Boland, J.J.; Hurley, P.K. Emulating synaptic response in n-and p-channel MoS 2 transistors by utilizing charge trapping dynamics. Sci. Rep. 2020, 10, 1–8. [Google Scholar] [CrossRef] [PubMed]
  15. Xiao, Z.; Huang, J. Energy-efficient hybrid perovskite memristors and synaptic devices. Adv. Electron. Mater. 2016, 2, 1600100. [Google Scholar] [CrossRef]
  16. Rehman, S.; Kim, H.; Khan, M.F.; Hur, J.-H.; Eom, J.; Kim, D.-K. Tunable resistive switching of vertical ReSe2/graphene hetero-structure enabled by Schottky barrier height and DUV light. J. Alloys Compd. 2020, 855, 157310. [Google Scholar] [CrossRef]
  17. Khan, M.F.; Ahmed, F.; Rehman, S.; Akhtar, I.; Rehman, M.A.; Shinde, P.A.; Khan, K.; Kim, D.-K.; Eom, J.; Lipsanen, H. High performance complementary WS2 devices with hybrid Gr/Ni contacts. Nanoscale 2020. [Google Scholar] [CrossRef] [PubMed]
  18. Khan, M.F.; Rehman, S.; Rehman, M.A.; Sagar, R.U.R.; Kim, D.-K.; Khalil, H.W.; Shinde, P.A.; ul Hassan, N.; Sharma, P.R.; Eom, J. Multi-heterostructured spin-valve junction of vertical FLG/MoSe. APL Mater. 2020, 8, 071104. [Google Scholar] [CrossRef]
  19. Khan, M.F.; Rehman, S.; Rehman, M.A.; Basit, M.A.; Kim, D.-K.; Ahmed, F.; Khalil, H.W.; Akhtar, I.; Jun, S.C. Modulation of Magnetoresistance Polarity in BLG/SL-MoSe 2 Heterostacks. Nanoscale Res. Lett. 2020, 15, 1–8. [Google Scholar] [CrossRef]
  20. Lee, C.; Lee, J.; Kim, M.; Woo, J.; Koo, S.-M.; Oh, J.-M.; Lee, D. Two-terminal structured synaptic device using ionic electrochemical reaction mechanism for neuromorphic system. IEEE Electron Device Lett. 2019, 40, 546–549. [Google Scholar] [CrossRef]
  21. 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]
  22. Rahmani, M.K.; Kim, M.-H.; Hussain, F.; Abbas, Y.; Ismail, M.; Hong, K.; Mahata, C.; Choi, C.; Park, B.-G.; Kim, S. Memristive and Synaptic Characteristics of Nitride-Based Heterostructures on Si Substrate. Nanomaterials 2020, 10, 994. [Google Scholar] [CrossRef]
  23. Du, C.; Ma, W.; Chang, T.; Sheridan, P.; Lu, W.D. Biorealistic implementation of synaptic functions with oxide memristors through internal ionic dynamics. Adv. Funct. Mater. 2015, 25, 4290–4299. [Google Scholar] [CrossRef]
  24. Kim, K.; Chen, C.L.; Truong, Q.; Shen, A.M.; Chen, Y. A carbon nanotube synapse with dynamic logic and learning. Adv. Mater. 2013, 25, 1693–1698. [Google Scholar] [CrossRef] [PubMed]
  25. Stephan, M.T.; Stephan, S.B.; Bak, P.; Chen, J.; Irvine, D.J. Synapse-directed delivery of immunomodulators using T-cell-conjugated nanoparticles. Biomaterials 2012, 33, 5776–5787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Oh, S.; Kim, T.; Kwak, M.; Song, J.; Woo, J.; Jeon, S.; Yoo, I.K.; Hwang, H. HfZrO x-based ferroelectric synapse device with 32 levels of conductance states for neuromorphic applications. IEEE Electron Device Lett. 2017, 38, 732–735. [Google Scholar] [CrossRef]
  27. Li, S.; Li, B.; Feng, X.; Chen, L.; Li, Y.; Huang, L.; Gong, X.; Fong, X.; Ang, K.-W. Gradual Resistive Switching in Electron Beam Irradiated ReS 2 Transistor and its Application as Electronic Synapse. In Proceedings of the 2020 International Symposium on VLSI Technology, Systems and Applications (VLSI-TSA), Hsinchu, Taiwan, 10–13 August 2020; pp. 157–158. [Google Scholar]
  28. Arnold, A.J.; Razavieh, A.; Nasr, J.R.; Schulman, D.S.; Eichfeld, C.M.; Das, S. Mimicking neurotransmitter release in chemical synapses via hysteresis engineering in MoS2 transistors. ACS Nano 2017, 11, 3110–3118. [Google Scholar] [CrossRef]
  29. Aftab, S.; Khan, M.F.; Gautam, P.; Noh, H.; Eom, J. MoTe2 van der Waals homojunction p-n diode with low resistive metal contacts. Nanoscale 2019, 11, 9518. [Google Scholar] [CrossRef]
  30. Koo, J.B.; Ku, C.H.; Lim, S.C.; Kim, S.H.; Lee, J.H. Hysteresis and threshold voltage shift of pentacene thin-film transistors and inverters with Al2O3 gate dielectric. Appl. Phys. Lett. 2007, 90, 133503. [Google Scholar] [CrossRef]
  31. Khan, M.F.; Nazir, G.; Lermolenko, V.M.; Eom, J. Electrical and photo-electrical properties of MoS2 nanosheets with and without an Al2O3 capping layer under various environmental conditions. Sci. Technol. Adv. Mater. 2016, 17, 166–176. [Google Scholar] [CrossRef] [Green Version]
  32. Illarionov, Y.Y.; Waltl, M.; Rzepa, G.; Knobloch, T.; Kim, J.-S.; Akinwande, D.; Grasser, T. Highly-stable black phosphorus field-effect transistors with low density of oxide traps. npj 2D Mater. Appl. 2017, 1, 1–7. [Google Scholar] [CrossRef]
  33. Kullmann, D.M.; Lamsa, K.P. Long-term synaptic plasticity in hippocampal interneurons. Nat. Rev. Neurosci. 2007, 8, 687–699. [Google Scholar] [CrossRef]
  34. Wan, X.; Yang, Y.; Feng, P.; Shi, Y.; Wan, Q. Short-term plasticity and synaptic filtering emulated in electrolyte-gated IGZO transistors. IEEE Electron Device Lett. 2016, 37, 299–302. [Google Scholar] [CrossRef]
  35. Dan, Y.; Poo, M.-M. Spike timing-dependent plasticity: From synapse to perception. Physiol. Rev. 2006, 86, 1033–1048. [Google Scholar] [CrossRef] [PubMed]
  36. Du, J.Y.; Ge, C.; Riahi, H.; Guo, E.J.; He, M.; Wang, C.; Yang, G.Z.; Jin, K.J. Dual-Gated MoS2 Transistors for Synaptic and Programmable Logic Functions. Adv. Electron. Mater. 2020, 1901408. [Google Scholar] [CrossRef]
  37. Zhou, L.; Mao, J.Y.; Ren, Y.; Yang, J.Q.; Zhang, S.R.; Zhou, Y.; Liao, Q.; Zeng, Y.J.; Shan, H.; Xu, Z. Biological spiking synapse constructed from solution processed bimetal core-shell nanoparticle based composites. Small 2018, 14, 1800288. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) The schematic illustration of MoTe2 three terminal field effect transistors (FET) where charge trapping and de-trapping dynamics at atomically thin channel 2D material/SiO2 interface. (b) The final optical microscope image of MoTe2 FET with Cr/Au electrodes. (c) Raman spectroscopy of few layer MoTe2. (d) Presentation of a biological synapse where action potentials control the release and absorption of neurotransmitters.
Figure 1. (a) The schematic illustration of MoTe2 three terminal field effect transistors (FET) where charge trapping and de-trapping dynamics at atomically thin channel 2D material/SiO2 interface. (b) The final optical microscope image of MoTe2 FET with Cr/Au electrodes. (c) Raman spectroscopy of few layer MoTe2. (d) Presentation of a biological synapse where action potentials control the release and absorption of neurotransmitters.
Nanomaterials 10 02326 g001
Figure 2. (a) Transfer characteristics of the MoTe2 FET having hysteresis loops at different delay times. (b) Mobility as a function of the delay time of MoTe2 FET. (c) Delay time dependent effective trap states density of MoTe2/SiO2 FET. (d) The transfer curves before and after UV treatment (10 s) of MoTe2/SiO2 FET.
Figure 2. (a) Transfer characteristics of the MoTe2 FET having hysteresis loops at different delay times. (b) Mobility as a function of the delay time of MoTe2 FET. (c) Delay time dependent effective trap states density of MoTe2/SiO2 FET. (d) The transfer curves before and after UV treatment (10 s) of MoTe2/SiO2 FET.
Nanomaterials 10 02326 g002
Figure 3. Schematic band diagram of MoTe2 FET with trapped states at the MoTe2/SiO2 interface. (a) When negative gate is applied. (b) When positive gate is applied.
Figure 3. Schematic band diagram of MoTe2 FET with trapped states at the MoTe2/SiO2 interface. (a) When negative gate is applied. (b) When positive gate is applied.
Nanomaterials 10 02326 g003
Figure 4. (a) Change in the threshold voltage of MoTe2 FET after 50 consecutive pulses to the gate terminal. (b) The excitatory post-synaptic current (EPSC) changes in MoTe2 synaptic transistor at negative gate pulse. (c) The inhibitory post-synaptic current (IPSC) changes in the MoTe2 synaptic transistor at positive gate pulse. (d) The application of negative and positive voltage pulses (V = −4 V and +4 V) demonstrating exponential trend in PSC, showing long-term depression (LTD) and long-term potentiation (LTP) processes, respectively.
Figure 4. (a) Change in the threshold voltage of MoTe2 FET after 50 consecutive pulses to the gate terminal. (b) The excitatory post-synaptic current (EPSC) changes in MoTe2 synaptic transistor at negative gate pulse. (c) The inhibitory post-synaptic current (IPSC) changes in the MoTe2 synaptic transistor at positive gate pulse. (d) The application of negative and positive voltage pulses (V = −4 V and +4 V) demonstrating exponential trend in PSC, showing long-term depression (LTD) and long-term potentiation (LTP) processes, respectively.
Nanomaterials 10 02326 g004
Figure 5. (a) Post-synaptic current vs. pulse number at different gate voltage pulse height, mimicking spiking amplitude-dependent plasticity. (b) The change in synaptic weight as a function of the time interval (Δt) between paired gate pulses of +4 and −4 V. (c) The change in EPSC after the application of two closely spaced pulses are applied to the gate terminal. (d) Quantification of STP using pulse paired facilitation measurements. Paired pulse facilitation (PPF) index designates the % change in the PSC due to the second pulse.
Figure 5. (a) Post-synaptic current vs. pulse number at different gate voltage pulse height, mimicking spiking amplitude-dependent plasticity. (b) The change in synaptic weight as a function of the time interval (Δt) between paired gate pulses of +4 and −4 V. (c) The change in EPSC after the application of two closely spaced pulses are applied to the gate terminal. (d) Quantification of STP using pulse paired facilitation measurements. Paired pulse facilitation (PPF) index designates the % change in the PSC due to the second pulse.
Nanomaterials 10 02326 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

Rehman, S.; Khan, M.F.; Rahmani, M.K.; Kim, H.; Patil, H.; Khan, S.A.; Kang, M.H.; Kim, D.-k. Neuro-Transistor Based on UV-Treated Charge Trapping in MoTe2 for Artificial Synaptic Features. Nanomaterials 2020, 10, 2326. https://doi.org/10.3390/nano10122326

AMA Style

Rehman S, Khan MF, Rahmani MK, Kim H, Patil H, Khan SA, Kang MH, Kim D-k. Neuro-Transistor Based on UV-Treated Charge Trapping in MoTe2 for Artificial Synaptic Features. Nanomaterials. 2020; 10(12):2326. https://doi.org/10.3390/nano10122326

Chicago/Turabian Style

Rehman, Shania, Muhammad Farooq Khan, Mehr Khalid Rahmani, Honggyun Kim, Harshada Patil, Sobia Ali Khan, Moon Hee Kang, and Deok-kee Kim. 2020. "Neuro-Transistor Based on UV-Treated Charge Trapping in MoTe2 for Artificial Synaptic Features" Nanomaterials 10, no. 12: 2326. https://doi.org/10.3390/nano10122326

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