The fast development of brain-inspired neuromorphic computing systems has stimulated urgent requirements for artificial synapses with low-power consumption. In this work, a photonic synaptic device based on (Al,Ga)N nanowire/graphene heterojunction has been proposed and demonstrated successfully. In the artificial synaptic device, the incident light, the nanowire/graphene heterojunction, and the light-generated carriers play the roles of action potential, pre-synaptic/post-synaptic membrane, and neurotransmitter in a biological synapse, respectively. As a key synaptic function, the paired pulse facilitation index of the photonic synapse can reach 202%, which can be modulated by the interval time between two adjacent light pulses. It is found that the graphene defects, the surface band bending, and the Al vacancies on the surface of (Al,Ga)N nanowires can be the key reasons contributing to the synaptic characteristics of artificial photonic devices. Hence, the dynamic “learning–forgetting” performance of the artificial synaptic device can resemble the “learning–forgetting” behavior of the human brain. Furthermore, the hand-written digits are set up to mimic a typical characteristic of human perceptual learning. After only three training epochs, the simulated network can achieve a high recognition rate of over 90% based on the experimental conductance for long-term potentiation and long-term depression. In supervised learning processes, such few training times are beneficial to reduce energy consumption significantly. Therefore, in the area of neuromorphic computing technology and artificial intelligence systems requiring low-power consumption, this work paves a potential way to develop the optoelectronic synapse based on semiconductor nanowires.

As the Internet of Things (IoT) and artificial intelligence (AI) develop rapidly, the von Neumann bottleneck becomes increasingly pronounced during the processing of complex information that involves interactions with a dynamic environment and various uncertainties.1,2 In contrast, the human brain exhibits excellent efficiency in solving complex and unstructured problems.3 The human brain is an interconnected network that is composed of organic biological microfibers (neurons and synapses), which can be used for information transport.4 Drawing inspiration from the neural architecture of the human brain, an artificial neural network (ANN) that emulates the functions of biological neural networks has garnered significant attention from the scientific community in recent years.3,5,6 Synaptic devices capable of synchronously storing and processing information are the essential components of an artificial neural network.7 Hence, the development of optoelectronic devices that can mimic synaptic behavior is quite crucial and necessary.3 

Due to the advantage of fast and broad-band information transmission with low-power consumption, photonic synaptic devices are considered the ideal device to simulate the biological synapse, which could confer significant advantages for neuromorphic computing.8–11 In general, light-stimulated synaptic devices contain active materials that have the ability to stimulate history-dependently, such as persistent photoconductivity (PPC).9,12,13 The varied conductivity or current can be utilized to emulate the functions of biological synapses under stimuli.9,13,14 Furthermore, (Al,Ga)N-based materials exhibit a stable PPC effect at room temperature, providing a promising route for the development of light-stimulated synaptic devices.15,16 In addition to their tunable bandgap and high stability, (Al,Ga)N-based materials are one of the most promising candidate materials for synapses devices.15 Moreover, combined with the excellent characteristics of a large surface-to-volume ratio and efficient charge carrier transport, these nanowires (NWs) can improve the memory duration of the photocurrent.17–22 

On the other hand, the heterojunction-based synapses device shows exceptionally sensitive photoconductance properties, and the photo-response spectral range can be easily controlled by using the absorber with an appropriate bandgap.23 Graphene has good conductivity and ultrahigh transmissivity in the UV range, which makes it suitable to be used in connecting NWs and conducting current.24,25 In our previous studies, we fabricated the GaN-based NWs and utilized them in photodetectors successfully.26–28 However, to date, very few works have been reported about utilizing (Al,Ga)N NW/graphene heterojunction in light-stimulated artificial synapses, not to mention studying the performance of power consumption and mechanisms.

In this work, we have demonstrated a light-stimulated synaptic device based on (Al,Ga)N NW/graphene heterojunction. This artificial synaptic device is used to emulate multiple functionalities of biological synapses, including spike-timing-dependence plasticity (STDP), spike-number-dependent plasticity (SNDP), learning behavior, and the transition from short-term memory (STM) to long-term memory (LTM). Apart from the experiments, we have also investigated the underlying mechanisms of this synaptic device. To evaluate the learning capabilities of the synaptic devices, an artificial neural network (ANN) for mimicking bio-vision has also been established and demonstrated.

By molecular beam epitaxy (MBE, Vecco G20), the (Al,Ga)N NWs [Fig. 1(a)] were grown on the n-type Si(111) substrates. The N atoms were supplied by the N plasma cell, while the Al and Ga atoms were generated by the Al and Ga effusion cells. To eliminate native oxides in the growth chamber before epitaxial growth, the Si substrates were heated to 900 °C for 15 min. First, a thin AlN buffer segment deposited on the Si surface forms islands by the Stranski–Krastanow growth mode, which acts as a material collector and a seed for the NW growth.29,30 After that, a GaN section was grown with a nominal Ga flux of 2.0 × 10−8 Torr for 100 min. Then a (Al,Ga)N segment was grown with a nominal Al/Ga flux ratio of 1.5 for 30 min. To characterize the growth rate of (Al,Ga)N NWs, two AlN segments were inserted into the (Al,Ga)N section with a nominal Al flux of 1.5 × 10−8 Torr for 4 min. During the epitaxial process, the nitrogen flow rate and plasma power were kept at 4.8 sccm and 450 W, respectively.

FIG. 1.

(a) MBE growth of the (Al,Ga)N NWs. (b) Deposit and selectively etch the SiO2 dielectric layer. (c) Fabricate the Ti/Al backside electrode and transfer the graphene. (d) Fabricate the Ti/Pt/Au front electrode and measure the device. (e) Schematic diagram of a biological synapse connecting two adjacent neurons. (f) Schematic illustration of the biological synapse. (g) Top-view and side-view SEM images of the (Al,Ga)N NWs. The white scale bars represent 500 nm. (h) Side-view AC-STEM image of an (Al,Ga)N NW. The white scale bars represent 50 nm. (i) AC-STEM image of the (Al,Ga)N NWs. The white scale bars represent 2 nm.

FIG. 1.

(a) MBE growth of the (Al,Ga)N NWs. (b) Deposit and selectively etch the SiO2 dielectric layer. (c) Fabricate the Ti/Al backside electrode and transfer the graphene. (d) Fabricate the Ti/Pt/Au front electrode and measure the device. (e) Schematic diagram of a biological synapse connecting two adjacent neurons. (f) Schematic illustration of the biological synapse. (g) Top-view and side-view SEM images of the (Al,Ga)N NWs. The white scale bars represent 500 nm. (h) Side-view AC-STEM image of an (Al,Ga)N NW. The white scale bars represent 50 nm. (i) AC-STEM image of the (Al,Ga)N NWs. The white scale bars represent 2 nm.

Close modal

In order to fabricate artificial synaptic devices, a thick layer of SiO2 was deposited on the surface of (Al,Ga)N NWs, followed by the selective removal of some portions of the SiO2 film to expose a certain square area of (Al,Ga)N NWs [Fig. 1(b)]. Subsequently, a backside electrode of Ti/Al (50/200 nm) was deposited onto the Si substrate by electron beam evaporation [Fig. 1(c)]. Then the sample was annealed in an N2 atmosphere at 400 °C for 30 min to form an ohmic contact. After that, graphene was transferred onto the NW surface by wet transfer [Fig. 1(c)]. After that, the artificial synaptic device was fabricated successfully by depositing another electrode consisting of Ti/Pt/Au (25/55/300 nm) onto the graphene surface within a certain area [Fig. 1(d)]. To study the mechanism and stability, another synaptic device (Device-HNO3) was prepared for comparison, whose NWs were treated with dilute nitric acid (HNO3) for 4 h.

Based on the experimental results of device parameters, such as nonlinear values, a multilayer perceptron (MLP) neural network model consisting of 784 input neurons, 100 hidden neurons, and 10 output neurons was established. In the network, each synaptic device was treated as a neuron. 28 × 28 image pixel information corresponds to the 784 input neurons. The Modified National Institute of Standards and Technology (MNIST) dataset was used for input. The sigmoid function was used as the activation function in this model. Furthermore, the output of ten classes of digits (0–9) was transformed into one pot vector for recognition, such as 2 => [0, 0, 1, 0, 0, 0, 0]. To optimize the weight value, a random gradient descent method was employed for continuous iteration. Moreover, the back-propagation algorithm was used to continually adjust the threshold values and weight of the network, updating weight values and minimizing the total square error in real time. Finally, the MNIST handwritten dataset was able to achieve accurate number recognition through the accurate evaluation and prediction of these numerical values.

The morphology and element distribution of (Al,Ga)N NWs were characterized in this study through scanning electron microscopy (SEM), spherical aberration corrected scanning transmission electron microscopy (AC-STEM), and high-resolution energy-dispersive X-ray (EDX) mapping. X-ray photoelectron spectroscopy (XPS) was used to characterize surface band bending (BB) and chemical states. All XPS spectra were analyzed by MultiPak software. Both core level and valence band maximum (VBM) were calibrated against C1s (284.8 eV). The measurement of current–time (I–T) characteristics was conducted using an Agilent B1505A semiconductor parameter analyzer. To determine the quality of the grown graphene layers, Raman spectroscopy was carried out using a laser excitation of 532 nm. To characterize the optical properties, a photoluminescence (PL) system with a 213 nm laser was utilized. In the response measurements, a function generator (FY6900-20M) was employed to control the 310 nm LED, which served as the light source.

In the biological nervous system, the synapse is an important structure to complete the signal transmission between two adjacent neurons and is composed of the pre-synapse, post-synapse, and synaptic cleft.31 The schematic images of biological neurons and synapses are illustrated in Figs. 1(e) and 1(f). When the biological pre-synapse is stimulated, the pre-synapse membrane will trigger the rapid release of neurotransmitters. Then the receptor on the post-synapse membrane opens the Ca2+ channel to realize signal transmission.32 In this work, the structure of synaptic devices based on (Al,Ga)N NWs is designed to emulate the functionalities of biological synapses [Figs. 1(d)1(f)]. The generation and transport of carriers in the artificial synaptic device under light stimulation are used to emulate the transport of neurotransmitters in the biological synapse.

In addition, as shown in Fig. 1(g), the as-grown (Al,Ga)N NWs show good uniformity and verticality. The diameter ranges from 30 to 80 nm, and the height of the NWs is around 580 nm. From Figs. 1(h) and 1(i), the structure of (Al,Ga)N NWs is labeled as I/II/I/II/III, corresponding to the (Al,Ga)N/AlN/(Al,Ga)N/AlN/GaN heterojunction, which agrees with the epitaxial design. Meanwhile, the lattice spacing between the two adjacent planes of GaN and (Al,Ga)N is around 2.7 and 3.1 Å, respectively. In the AC-STEM image, such a clear lattice fringe is observed, which is testimony to good crystallinity. Meanwhile, Fig. S1(a) (supplementary material) displays the PL spectrum of (Al,Ga)N NWs. The composition of Al elements within (Al,Ga)N NWs can be calculated by the following equation:
(1)
Eg(AlN) and Eg(GaN) are the bandgaps of AlN (6.2 eV) and GaN (3.4 eV), respectively.33,34 The PL peak of (Al,Ga)N NWs is centered at a wavelength of 322 nm, corresponding to an Al composition of ∼22.3%. Figure S1(b) shows a Raman spectrum describing the characteristics of the graphene. The D peak (∼1346 cm−1) is utilized to signify the presence of defects, and the G peak (∼1580 cm−1) is caused by the in-plane vibration of sp2 hybrid carbon atoms.35 In addition, the Raman intensity ratio of the 2D peak (∼2691 cm−1) and the G peak is 0.53, indicating that graphene has a multi-layer structure.36 As clearly illustrated in Fig. S1(b), the presence of both the D and G peaks serves as evidence for the presence of defects within the graphene, which can have natural carrier-trapping characteristics.37 
The collected current in the (Al,Ga)N NWs synaptic device is regarded as excitatory postsynaptic current (EPSC) under illumination, depending on the synaptic weight.38 The synaptic weight can be changed by adjusting the activity of pre-synaptic and post-synaptic neurons, which is commonly called synaptic plasticity.39 It lays the foundation for many functions in the human brain, such as learning, forgetting, and image recognition.40,41 Moreover, paired-pulse facilitation (PPF) is a typical phenomenon of short-range plasticity, which is essential for decoding provisional information in the auditory and visual systems.42 The PPF manifestation is that the EPSC responding to the second stimulus is always greater than the first stimulus when the pre-synaptic neurons receive two short stimuli continuously. From Fig. 2(a), EPSC is induced by two continuous light pulses with an interval of 0.05 s. A1 and A2 are the peak currents under the first and second stimulation spikes, respectively.43 The second peak (A2) is obviously larger than the first peak (A1), which corresponds to the typical PPF phenomenon. The PPF behavior is quantified by the PPF index (QPPF), which is defined as a ratio of A2/A1. As shown in Fig. 2(b), the PPF index is dependent on the light-pulse interval time (toff), which can be fitted well by the following equation:44,
(2)
C1 and C2 represent the initial facilitation magnitudes. τ1 and τ2 represent the characteristic relaxation times of the fast and slow decay processes, respectively. With the toff increasing from 0.02 to 1.98 s, the PPF index value decreases from 202% to 113%. By Eq. (2), the values of τ1 and τ2 are calculated to be 0.02 and 8.50 s, corresponding to rapid and slow decay, respectively. According to C1 (156) and C2 (138), the fast decay is facilitated more than the slow decay.45 The PPF difference between the slow and rapid decays can be regarded as a biological synaptic connection enhancement.46 Moreover, the maximum PPF index value is larger than some other NW structure synaptic devices.45,47,48 Hence, the artificial synaptic device based on the (Al, Ga)N NWs is demonstrated to have the ability to simulate biological synapses.
FIG. 2.

(a) EPSC of the (Al, Ga)N NW synaptic device triggered by two consecutive light pulses with a power density of 6.4 µW cm−2 (b) PPF index at different toff. The inset shows the top-view optical image of the synaptic device. ton is the light-pulse width.

FIG. 2.

(a) EPSC of the (Al, Ga)N NW synaptic device triggered by two consecutive light pulses with a power density of 6.4 µW cm−2 (b) PPF index at different toff. The inset shows the top-view optical image of the synaptic device. ton is the light-pulse width.

Close modal

In the biological nervous system, a brief and slight increase in synaptic weight will lead to the formation of STM, which is located in the hippocampus for only a few minutes at most. If repeated stimulation is applied to the synapse, the weight of the synapse will continue to increase. After that, STM will transfer from the hippocampus to the cerebral cortex and become LTM, where it can be long-lasting or even permanent.49 Inspired by the biological synapses, the STDP and SNDP of the (Al,Ga)N NW synaptic device are explored and shown in Fig. 3. As clearly shown in Fig. 3(a), a small EPSC triggered by a short ton (0.05 s) decays rapidly, while a larger EPSC triggered by a longer ton (0.5 s) requires more time to reach the level of the black curve, indicating a transition from STM to LTM. To reach the same baseline level (orange dashed line) after undergoing five light pulses [Fig. 3(b)], the synaptic device exhibits a higher EPSC when the toff is shorter (0.1 vs 0.9 s). Similar phenomena can be observed under different light power densities [Figs. S2(a)–S2(d)]. The spike numbers in Figs. 3(c) and 3(d) are one and five, respectively, leading to different EPSC results. When ton and toff are both 0.1 s, the EPSC results of the device triggered by five light pulses [Fig. 3(d)] are higher than those triggered by only one light pulse [Fig. 3(c)]. Therefore, the synaptic device exhibits SNDP performance. Under simulation with the different light power densities, the SNDP characteristics can also be observed in Figs. S2(e) and S2(f). For the synaptic device, the ability to transfer STM to LTM depends on the light-pulse time of duration, frequency, pulse number, and light intensity, reflecting the synaptic plasticity of the device.

FIG. 3.

(a) Dependence of the EPSC triggered by a light pulse (6.4 µW cm−2) with different light-pulse widths. (b) Dependence of the EPSC triggered by five light pulses (6.4 µW cm−2) with different toff. The ton is 0.1 s. (c) EPSC triggered by a light pulse with different ton and light power densities (from 6.4 to 26 µW cm−2). The ton and toff keep the same. (d) EPSC triggered by five light pulses with different toff and light power densities (from 6.4 to 26 µW cm−2). The ton is 0.1 s.

FIG. 3.

(a) Dependence of the EPSC triggered by a light pulse (6.4 µW cm−2) with different light-pulse widths. (b) Dependence of the EPSC triggered by five light pulses (6.4 µW cm−2) with different toff. The ton is 0.1 s. (c) EPSC triggered by a light pulse with different ton and light power densities (from 6.4 to 26 µW cm−2). The ton and toff keep the same. (d) EPSC triggered by five light pulses with different toff and light power densities (from 6.4 to 26 µW cm−2). The ton is 0.1 s.

Close modal
In order to investigate the effects of surface band bending, synaptic devices with and without HNO3 etching have been prepared, which are regarded as Device-HNO3 and Device (Figs. S3 and 4), respectively. Compared with those of Device under both 5 and 20 light-pulse stimulations [Figs. 4(a) and 4(b)], the EPSC of Device-HNO3 is smaller, while the decay time of Device-HNO3 is longer. To further study the underlying mechanism, the energy band diagrams are plotted in Figs. 4(c)4(f).50,51 Figure S3(a) illustrates an energy diagram used for calculating BB, which can be calculated by the following equation:52,
(3)
The bandgap of (Al,Ga)N NWs (Eg) is 3.85 eV, which is calculated by the PL peak [Fig. S1(a)]. (CLGa3dEV)bulk is the energy difference between Ga 3d core level and VBM within (Al,Ga)N NWs, which is 17.75 eV [Fig. S3(b)]. CLGa3dEFsurf represents the core level binding energy on (Al,Ga)N NW surface. Φn is the gap between the conduction band minimum and Fermi level, which is a constant in the work.52 According to Figs. S3(c) and S3(d), CLGa3dEFsurf of Device (20.11 eV) is higher than that of Device-HNO3 (19.72 eV). Hence, the BB value of Device-HNO3 is higher than that of Device (20.11 > 19.72), indicating that Device-HNO3 has a larger surface band bending [Figs. 4(e) and 4(f)].53 Furthermore, the Al/N ratio of Device-HNO3 is lower than that of Device (Fig. S4), indicating the increase of Al vacancy after HNO3 etching [Figs. 4(e) and 4(f)].52 In addition, the natural defects of graphene can form several carrier capture centers at the interfaces and surfaces [Figs. 4(c)4(f)], referring to the middle energy band (MGB).54 MGB can form a unique bandgap between the conduction and valence bands.55 Due to the obvious D peak in the Raman spectrum [Fig. S1(b)], the graphene is demonstrated to have defects, acting as the carrier reservoirs.
FIG. 4.

Dependence of the EPSC triggered by (a) 5 and (b) 20 light pulses. The light-pulse width and interval time are both 0.2 s. The light power density is 6.4 µW cm−2. Schematic energy band diagrams of the synaptic device (c) under light stimulation and (d) without light stimulation. Schematic energy band diagrams of the Device-HNO3 (e) under light stimulation and (f) without light stimulation.

FIG. 4.

Dependence of the EPSC triggered by (a) 5 and (b) 20 light pulses. The light-pulse width and interval time are both 0.2 s. The light power density is 6.4 µW cm−2. Schematic energy band diagrams of the synaptic device (c) under light stimulation and (d) without light stimulation. Schematic energy band diagrams of the Device-HNO3 (e) under light stimulation and (f) without light stimulation.

Close modal

In the synaptic device, (Al,Ga)N NWs are used as light absorbers. The heterojunction between graphene and (Al,Ga)N NWs can generate a built-in electric field, which facilitates the separation of photogenerated carriers. When the device is stimulated by light, photogenerated carriers will be generated and separated under the built-in electric field [Fig. 4(c)]. Subsequently, some of the photogenerated carriers are trapped by capture centers (e.g., Al vacancies and graphene defects), while others transport to the electrodes, contributing to the current signal. When the light is off, these photogenerated carriers captured could slow down the speed of current recovery [Fig. 4(d)]. The larger BB in Device-HNO3 can make the transport of carriers more difficult [Fig. 4(e)], leading to a lower EPSC [Figs. 4(a) and 4(b)]. Furthermore, due to the increased Al vacancies in Device-HNO3 (Fig. S4), more carriers will be captured. When the light stimulus is removed, the carriers in the trapped states need more time to transport [Fig. 4(f)], resulting in a longer decay time [Fig. 4(b)]. Therefore, graphene defects, surface band bending, and vacancies can be the key reasons contributing to the synaptic characteristics. In such a synaptic device, the incident light, the light-generated carriers, and the electrodes are used to play the roles of the action potential, neurotransmitters, and pre-synaptic/post-synaptic membranes in the biological synapse, respectively.

The energy consumption of a single synaptic event is calculated by the following equation:56 
(4)
t0 and t1 represent the time of light on and light off, respectively. V is the operating voltage, and I is the response current of the device. The light power density is 6.4 µW cm−2, which can be calibrated by the method shown in Ref. 27. For a single synaptic event with a pulse width (t1t0) of 10 ms and a voltage of 3 V, the energy consumption is calculated to be 2.6 µJ [Fig. 5(a)]. This energy consumption could be further reduced by decreasing the light pulse width and operating voltage.
FIG. 5.

(a) The energy consumption for a single synaptic event. (b) The learning-experience behavior of the human brain is simulated by continuous light pulses.

FIG. 5.

(a) The energy consumption for a single synaptic event. (b) The learning-experience behavior of the human brain is simulated by continuous light pulses.

Close modal
In addition, the process of learning in the human brain is complex and dynamic. When we experience something new, the patterns of electrical and chemical activities in the human brain are modified, leading to changes in synaptic weight. The strengthening and weakening of synaptic weight are thought to underlie our ability to learn and remember new information.43 From a biological perspective, the transition of synaptic plasticity in the nervous system is highly related to the processes of learning and forgetting.41 To investigate the synaptic device’s capability in simulating the learning behavior of the human brain, the synaptic weight (Δw) is utilized as a metric to represent the memory level of the device, which is defined as the following equation:57 
(5)
I0 and In are the initial current and the current after light stimulation (n is the number of light pulses), respectively. As shown in Fig. 5(b), the synaptic weight increases with increasing pulse numbers, corresponding to the learning process of the human brain. During the initial learning process, the synaptic weight increases from 52.9% to 90.1% after stimulating by 17 light pulses. After a period of forgetting, the relearning process begins, and it is observed that the previous maximum Δw value can be reacquired after stimulation by only ten light pulses. In other words, during the second learning process, a similar level of cognition can be achieved with a fewer number of light pulses.
To demonstrate the potential applications, the synaptic device based on (Al,Ga)N NWs has been employed to simulate ANNs used for pattern recognition by the instructions in NeuroSim V3.0.58 The ANN adopts a three-layer perception neural model [Fig. 6(a)], which utilizes supervised learning with back-propagation and random gradient descent algorithms.59 From the MNIST database, the hand-written digits can be achieved, which are divided into 28 × 28 pixels, corresponding to 784 input neurons.60 The hidden layer consists of 100 neurons, while the output layer comprises ten neurons corresponding to the digits from 0 to 9.42 By the calculation processes of algorithms, the synaptic weights will eventually be saved, which are extracted from the conductance of different potentiating and depressing states [Fig. 6(b)] by the following equations:47,
(6)
(7)
(8)
(9)
GLTP and GLTD are the conductance values associated with long-term potentiation (LTP) and long-term depression (LTD), respectively.47, Gmax, Gmin, and Pmax are extracted from the experimental data and represent the maximum conductance, minimum conductance, and corresponding pulse number, respectively.47 The parameter A controls the nonlinear behavior of weight updates. Gnorm is the conductance after normalization. Gn is the conductance value of the current state.47 In this work, Pmax is 28. From Fig. 6(b), A is fitted to be a positive value (0.1) for the green curve and a negative value (−0.12) for the red curve.47  B is a function of A by fitting Gmax, Gmin, and Pmax, which are also positive (1.0) for the green curve and negative (−2.4 × 10−4) for the red curve. Furthermore, the nonlinear data of LTP and LTD are fitted to be 8.26 and −7.82, respectively.58 By simulating ANNs on the MNIST database, 20 samples are randomly selected in Fig. 6(c), while 10 000 samples are identified in Fig. 6(d). Most numbers can be identified accurately [Fig. 6(c)]. As depicted in Fig. 6(d), the simulated network is capable of achieving a high recognition rate of up to 92% after 30 training epochs [Fig. 6(d)]. Due to the requirement of low-power consumption for the synaptic devices, reducing the number of training epochs can decrease the energy consumption significantly in the supervised learning process. After only three training epochs, the recognition rate can increase rapidly to over 90%.
FIG. 6.

(a) Schematic diagram of ANN simulation using 784 × 100 × 10 synaptic weights. (b) The experimental data and fitted curves of LTP/LTD characteristics triggered by light pulses. (c) Recognition results of the randomly selected numbers from the MNIST database. (d) Recognition accuracy vs training epoch in the simulation.

FIG. 6.

(a) Schematic diagram of ANN simulation using 784 × 100 × 10 synaptic weights. (b) The experimental data and fitted curves of LTP/LTD characteristics triggered by light pulses. (c) Recognition results of the randomly selected numbers from the MNIST database. (d) Recognition accuracy vs training epoch in the simulation.

Close modal

In this work, an artificial synaptic device based on (Al,Ga)N NW/graphene heterostructure has been successfully fabricated. Such a photonic synapse has been demonstrated to have the characteristics of STDP, SNDP, and light-intensity dependence. The energy consumption of the artificial synapse can be only 2.6 µJ for a single synaptic event. Furthermore, the PPF index of the photonic synapse can reach 202%. Thanks to the dynamic “learning–forgetting” performance, human learning behavior can also be emulated in light-stimulated synaptic devices. It is found that the reservoir effect of the graphene, the surface band bending, and the Al vacancies on the nanowire surfaces can contribute to the synaptic characteristics. Based on the experimental conductance for long-term potentiation and long-term depression, the recognition accuracy of simulating the neural network can reach 92%. Hence, such artificial synapses based on NW/graphene heterostructures could provide a pathway for applications requiring low-power consumption, including bio-realistic artificial intelligence, neuromorphic computing systems, etc.

See the supplementary material for additional details on the data analysis.

The authors are grateful for the Key Research Program of Frontier Sciences, CAS (Grant No. ZDBS-LY-JSC034), the Research Program of Scientific Instrument and Equipment of CAS (Grant No. YJKYYQ20200073), and the National Natural Science Foundation of China (Grant No. 62174172). The authors are thankful for the technical support from Vacuum Interconnected Nanotech Workstation (Nano-X, Grant No. F2309), Platform for Characterization and Test of SINANO, CAS.

The authors have no conflicts to disclose.

M.Z. completed all experiments in device fabrication, device measurements, and the corresponding data collections and analyses. Y.Z. conceived the idea and guided the work. Y.Z. and X.G. completed the SEM measurements. Q.Z. and M.Z. completed the numerical simulations. M.Z., Y.Z., and J.Z. completed the mechanism study. M.Z., Y.Z., and S.L. wrote the original draft of this work. S.L. and Y.Z. carried out the funding acquisition and project administration. Y.Z. carried out all MBE experiments. M.Z., Y.Z., J.Z., and M.J. carried out the methodology and visualization of this work. M.Z., J.Z., and M.J. performed the investigation. All authors reviewed this manuscript.

Min Zhou: Conceptualization (equal); Data curation (lead); Formal analysis (equal); Investigation (equal); Methodology (lead); Writing – original draft (equal); Writing – review & editing (equal). Yukun Zhao: Conceptualization (equal); Funding acquisition (equal); Project administration (equal); Supervision (lead); Writing – original draft (equal); Writing – review & editing (equal). Xiushuo Gu: Formal analysis (equal); Investigation (equal); Methodology (equal); Writing – review & editing (equal). Qianyi Zhang: Investigation (equal); Software (lead); Writing – review & editing (equal). Jianya Zhang: Formal analysis (equal); Validation (equal); Writing – review & editing (equal). Min Jiang: Methodology (supporting); Validation (equal); Writing – review & editing (equal). Shulong Lu: Funding acquisition (equal); Supervision (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

1.
P.
Monalisha
,
A. P. S.
Kumar
,
X. R.
Wang
, and
S. N.
Piramanayagam
, “
Emulation of synaptic plasticity on a cobalt-based synaptic transistor for neuromorphic computing
,”
ACS Appl. Mater. Interfaces
14
,
11864
(
2022
).
2.
D. G.
Roe
,
D. H.
Ho
,
Y. Y.
Choi
,
Y. J.
Choi
,
S.
Kim
,
S. B.
Jo
,
M. S.
Kang
,
J.-H.
Ahn
, and
J. H.
Cho
, “
Humanlike spontaneous motion coordination of robotic fingers through spatial multi-input spike signal multiplexing
,”
Nat. Commun.
14
,
5
(
2023
).
3.
P.
Monalisha
,
S.
Li
,
T.
Jin
,
P. S. A.
Kumar
, and
S. N.
Piramanayagam
, “
Synaptic plasticity investigation in permalloy based channel material for neuromorphic computing
,”
J. Phys. D: Appl. Phys.
56
,
015302
(
2023
).
4.
H.
Shi
,
B.
Wang
,
X.
Wang
,
L.
Qiu
, and
L.
Zheng
, “
Optically readable electrochromic-based microfiber synaptic device for photonic neuromorphic systems
,”
ACS Appl. Mater. Interfaces
15
,
9705
(
2023
).
5.
K.-T.
Chen
,
L.-C.
Shih
,
S.-C.
Mao
, and
J.-S.
Chen
, “
Mimicking pain-perceptual sensitization and pattern recognition based on capacitance- and conductance-regulated neuroplasticity in neural network
,”
ACS Appl. Mater. Interfaces
15
,
9593
(
2023
).
6.
A.
Saleem
,
D.
Kumar
,
F.
Wu
,
L. B.
Keong
, and
T.-Y.
Tseng
, “
An opto-electronic HfOx-based transparent memristive synapse for neuromorphic computing system
,”
IEEE Trans. Electron Devices
70
,
1351
(
2023
).
7.
M. M.
Mburu
,
K. T.
Lu
,
N. L.
Prine
,
A. N.
Au‐Duong
,
W. H.
Chiang
,
X.
Gu
, and
Y. C.
Chiu
, “
Conjugated polymer-wrapped single-wall carbon nanotubes for high-mobility photonic/electrical fully modulated synaptic transistor
,”
Adv. Mater. Technol.
7
,
2101506
(
2022
).
8.
Y.
Nie
,
P.
Xie
,
X.
Chen
,
C.
Jin
,
W.
Liu
,
X.
Shi
,
Y.
Xu
,
Y.
Peng
,
J. C.
Ho
,
J.
Sun
, and
J.
Yang
, “
Hybrid C8-BTBT/InGaAs nanowire heterojunction for artificial photosynaptic transistors
,”
J. Semicond.
43
,
112201
(
2022
).
9.
Y.
Mo
,
B.
Luo
,
H.
Dong
, and
B.
Hou
, “
Light-stimulated artificial synapses based on Si-doped GaN thin films
,”
J. Mater. Chem. C
10
,
13099
(
2022
).
10.
S.
Shrivastava
,
L. B.
Keong
,
S.
Pratik
,
A. S.
Lin
, and
T. Y.
Tseng
, “
Fully photon controlled synaptic memristor for neuro-inspired computing
,”
Adv. Electron. Mater.
9
,
2201093
(
2023
).
11.
Y.
Wang
,
K.
Wang
,
X.
Hu
,
Y.
Wang
,
W.
Gao
,
Y.
Zhang
,
Z.
Liu
,
Y.
Zheng
,
K.
Xu
,
D.
Yang
, and
X.
Pi
, “
Optogenetics-inspired fluorescent synaptic devices with nonvolatility
,”
ACS Nano
17
,
3696
(
2023
).
12.
H.
Seung
,
C.
Choi
,
D. C.
Kim
,
J. S.
Kim
,
J. H.
Kim
,
J.
Kim
,
S. I.
Park
,
J. A.
Lim
,
J.
Yang
,
M. K.
Choi
,
T.
Hyeon
, and
D. H.
Kim
, “
Integration of synaptic phototransistors and quantum dot light-emitting diodes for visualization and recognition of UV patterns
,”
Sci. Adv.
8
,
eabq3101
(
2022
).
13.
M.
Lee
,
W.
Lee
,
S.
Choi
,
J.-W.
Jo
,
J.
Kim
,
S. K.
Park
, and
Y.-H.
Kim
, “
Brain-inspired photonic neuromorphic devices using photodynamic amorphous oxide semiconductors and their persistent photoconductivity
,”
Adv. Mater.
29
,
1700951
(
2017
).
14.
J.
Zhou
,
J.
Dai
, and
S.
Weng
, “
Effect of adjacent lateral inhibition on light and electric-stimulated synaptic transistors
,”
IEEE Electron Device Lett.
43
,
573
(
2022
).
15.
T.-T.
Wang
,
S.
Dong
,
C.
Li
,
W.-C.
Yue
,
Y.-Y.
Lyu
,
C.-G.
Wang
,
C.-K.
Zeng
,
Z.
Yuan
,
W.
Zhu
,
Z.-L.
Xiao
,
X.
Lu
,
B.
Liu
,
H.
Lu
,
H.-B.
Wang
,
P.
Wu
,
W.-K.
Kwok
, and
Y.-L.
Wang
, “
In situ tunable giant electrical anisotropy in a grating gated AlGaN/GaN two-dimensional electron gas
,”
Appl. Phys. Lett.
121
,
092101
(
2022
).
16.
M.
Wohlfahrt
,
M. J.
Uren
,
F.
Kaess
,
O.
Laboutin
,
H.
Hirshy
, and
M.
Kuball
, “
UV-induced change in channel conductivity in AlGaN/GaN high electron mobility transistors to measure doping
,”
Appl. Phys. Lett.
118
,
163501
(
2021
).
17.
L.
Zhang
,
X.
Xiu
,
Y.
Li
,
Y.
Zhu
,
X.
Hua
,
Z.
Xie
,
T.
Tao
,
B.
Liu
,
P.
Chen
,
R.
Zhang
, and
Y.
Zheng
, “
Solar-blind ultraviolet photodetector based on vertically aligned single-crystalline beta-Ga2O3 nanowire arrays
,”
Nanophotonics
9
,
4497
(
2020
).
18.
S.
Xin
,
Y.
Chang
,
R.
Zhou
,
H.
Cong
,
L.
Zheng
,
Y.
Wang
,
Y.
Qin
,
P.
Xu
,
X.
Liu
, and
F.
Wang
, “
Ultraviolet-driven metal oxide semiconductor synapses with improved long-term potentiation
,”
J. Mater. Chem. C
11
,
722
(
2023
).
19.
S.
Poddar
,
Z.
Chen
,
Z.
Ma
,
Y.
Zhang
,
C. L. J.
Chan
,
B.
Ren
,
Q.
Zhang
,
D.
Zhang
,
G.
Shen
,
H.
Zeng
, and
Z.
Fan
, “
Robust lead-free perovskite nanowire array-based artificial synapses exemplifying Gestalt principle of closure via a letter recognition scheme
,”
Adv. Intell. Syst.
4
,
2200065
(
2022
).
20.
G.
Milano
,
G.
Pedretti
,
K.
Montano
,
S.
Ricci
,
S.
Hashemkhani
,
L.
Boarino
,
D.
Ielmini
, and
C.
Ricciardi
, “
In materia reservoir computing with a fully memristive architecture based on self-organizing nanowire networks
,”
Nat. Mater.
21
,
195
(
2022
).
21.
C.
Ebenhoch
and
L.
Schmidt‐Mende
, “
TiO2 nanowire array memristive devices emulating functionalities of biological synapses
,”
Adv. Electron. Mater.
7
,
2000950
(
2021
).
22.
M. H. R.
Ansari
,
U. M.
Kannan
, and
S.
Cho
, “
Core-shell dual-gate nanowire charge-trap memory for synaptic operations for neuromorphic applications
,”
Nanomaterials
11
,
1773
(
2021
).
23.
G.
Moon
,
S. Y.
Min
,
C.
Han
,
S. H.
Lee
,
H.
Ahn
,
S. Y.
Seo
,
F.
Ding
,
S.
Kim
, and
M. H.
Jo
, “
Atomically thin synapse networks on van der Waals photo-memtransistors
,”
Adv. Mater.
35
,
2203481
(
2022
).
24.
W.-Y.
Kong
,
G.-A.
Wu
,
K.-Y.
Wang
,
T.-F.
Zhang
,
Y.-F.
Zou
,
D.-D.
Wang
, and
L.-B.
Luo
, “
Graphene-β-Ga2O3 heterojunction for highly sensitive deep UV photodetector application
,”
Adv. Mater.
28
,
10725
(
2016
).
25.
Q.
Hu
,
Z.
Wang
,
Y.
Qiu
,
J.
Lin
,
X.
Lin
,
X.
Wei
,
D.
Ye
, and
W.
Zheng
, “
Ternary compound MgTiO3 combined with graphene for solar-blind deep ultraviolet photodetection
,”
J. Alloys Compd.
911
,
165031
(
2022
).
26.
J.
Zhang
,
B.
Jiao
,
J.
Dai
,
D.
Wu
,
Z.
Wu
,
L.
Bian
,
Y.
Zhao
,
W.
Yang
,
M.
Jiang
, and
S.
Lu
, “
Enhance the responsivity and response speed of self-powered ultraviolet photodetector by GaN/CsPbBr3 core-shell nanowire heterojunction and hydrogel
,”
Nano Energy
100
,
107437
(
2022
).
27.
M.
Jiang
,
Y.
Zhao
,
L.
Bian
,
W.
Yang
,
J.
Zhang
,
Y.
Wu
,
M.
Zhou
,
S.
Lu
, and
H.
Qin
, “
Self-powered photoelectrochemical (Al,Ga)N photodetector with an ultrahigh ultraviolet/visible reject ratio and a quasi-invisible functionality for 360 degrees omnidirectional detection
,”
ACS Photonics
8
,
3282
(
2021
).
28.
M.
Zhou
,
Y.
Zhao
,
W.
Yang
,
J.
Zhang
,
M.
Jiang
,
Y.
Wu
,
Z.
Xu
, and
S.
Lu
, “
Detached vertical (Al,Ga)N nanowires to realize the flexible ultraviolet photodetector with high ultraviolet/visible reject ratio and detectivity
,”
Energy Technol.
10
,
2200885
(
2022
).
29.
R.
Songmuang
,
O.
Landré
, and
B.
Daudin
, “
From nucleation to growth of catalyst-free GaN nanowires on thin AlN buffer layer
,”
Appl. Phys. Lett.
91
,
251902
(
2007
).
30.
T.
Auzelle
,
B.
Haas
,
A.
Minj
,
C.
Bougerol
,
J.-L.
Rouvière
,
A.
Cros
,
J.
Colchero
, and
B.
Daudin
, “
The influence of AlN buffer over the polarity and the nucleation of self-organized GaN nanowires
,”
J. Appl. Phys.
117
,
245303
(
2015
).
31.
U.
Das
,
P.
Sarkar
,
B.
Paul
, and
A.
Roy
, “
Halide perovskite two-terminal analog memristor capable of photo-activated synaptic weight modulation for neuromorphic computing
,”
Appl. Phys. Lett.
118
,
182103
(
2021
).
32.
G.
Hu
,
H.
An
,
J.
Xi
,
J.
Lu
,
Q.
Hua
, and
Z.
Peng
, “
A ZnO micro/nanowire-based photonic synapse with piezo-phototronic modulation
,”
Nano Energy
89
,
106282
(
2021
).
33.
N. A. A.
Zulkifli
,
K.
Park
,
J. W.
Min
,
B. S.
Ooi
,
R.
Zakaria
,
J.
Kim
, and
C. L.
Tan
, “
A highly sensitive, large area, and self-powered UV photodetector based on coalesced gallium nitride nanorods/graphene/silicon (111) heterostructure
,”
Appl. Phys. Lett.
117
,
191103
(
2020
).
34.
D.
Li
,
K.
Jiang
,
X.
Sun
, and
C.
Guo
, “
AlGaN photonics: Recent advances in materials and ultraviolet devices
,”
Adv. Opt. Photonics
10
,
43
(
2018
).
35.
C.
Ramesh
,
P.
Tyagi
,
V.
Aggarwal
,
S.
Pal
,
S.
Dhua
,
M.
Singh
,
P.
Pal
,
S. C.
Roy
,
S. P.
Singh
,
S. K.
Muthusamy
, and
S. S.
Kushvaha
, “
Hybrid reduced graphene oxide/GaN nanocolumns on flexible niobium foils for efficient photoelectrochemical water splitting
,”
ACS Appl. Nano Mater.
6
,
1898
(
2023
).
36.
A. C.
Ferrari
,
J. C.
Meyer
,
V.
Scardaci
,
C.
Casiraghi
,
M.
Lazzeri
,
F.
Mauri
,
S.
Piscanec
,
D.
Jiang
,
K. S.
Novoselov
,
S.
Roth
, and
A. K.
Geim
, “
Raman spectrum of graphene and graphene layers
,”
Phys. Rev. Lett.
97
,
187401
(
2006
).
37.
B.
Sahoo
,
J.
Joseph
,
A.
Sharma
, and
J.
Paul
, “
Surface modification of aluminium by graphene impregnation
,”
Mater. Des.
116
,
51
(
2017
).
38.
J.
Jiang
,
W.
Hu
,
D.
Xie
,
J.
Yang
,
J.
He
,
Y.
Gao
, and
Q.
Wan
, “
2D electric-double-layer phototransistor for photoelectronic and spatiotemporal hybrid neuromorphic integration
,”
Nanoscale
11
,
1360
(
2019
).
39.
A.
Destexhe
and
E.
Marder
, “
Plasticity in single neuron and circuit computations
,”
Nature
431
,
789
(
2004
).
40.
R. S.
Zucker
and
W. G.
Regehr
, “
Short-term synaptic plasticity
,”
Annu. Rev. Physiol.
64
,
355
(
2002
).
41.
R.
Li
,
Y.
Dong
,
F.
Qian
,
Y.
Xie
,
X.
Chen
,
Q.
Zhang
,
Z.
Yue
, and
M.
Gu
, “
CsPbBr3/graphene nanowall artificial optoelectronic synapses for controllable perceptual learning
,”
PhotoniX
4
,
4
(
2023
).
42.
C.
Shen
,
X.
Gao
,
C.
Chen
,
S.
Ren
,
J.-L.
Xu
,
Y.-D.
Xia
, and
S.-D.
Wang
, “
ZnO nanowire optoelectronic synapse for neuromorphic computing
,”
Nanotechnology
33
,
065205
(
2022
).
43.
J.
Zhao
,
S.
Zheng
,
L.
Zhou
,
W.
Mi
,
Y.
Ding
, and
M.
Wang
, “
An artificial optoelectronic synapse based on MoOx film
,”
Nanotechnology
34
,
145201
(
2023
).
44.
C.
Mahata
,
J.
Park
,
M.
Ismail
, and
S.
Kim
, “
Demonstration of electronic and optical synaptic properties modulation of reactively sputtered zinc-oxide-based artificial synapses
,”
J. Alloys Compd.
938
,
168539
(
2023
).
45.
T. S.
Rao
,
S.
Kundu
,
B.
Bannur
,
S. J.
George
, and
G. U.
Kulkarni
, “
Emulating Ebbinghaus forgetting behavior in a neuromorphic device based on 1D supramolecular nanofibres
,”
Nanoscale
15
,
7450
(
2023
).
46.
S. M.
Kwon
,
S. W.
Cho
,
M.
Kim
,
J. S.
Heo
,
Y. H.
Kim
, and
S. K.
Park
, “
Environment-adaptable artificial visual perception behaviors using a light-adjustable optoelectronic neuromorphic device array
,”
Adv. Mater.
31
,
1906433
(
2019
).
47.
P.
Xie
,
Y.
Huang
,
W.
Wang
,
Y.
Meng
,
Z.
Lai
,
F.
Wang
,
S.
Yip
,
X.
Bu
,
W.
Wang
,
D.
Li
,
J.
Sun
, and
J. C.
Ho
, “
Ferroelectric P(VDF-TrFE) wrapped InGaAs nanowires for ultralow-power artificial synapses
,”
Nano Energy
91
,
106654
(
2022
).
48.
D.
Kumar
,
H.
Li
,
U. K.
Das
,
A. M.
Syed
, and
N.
El-Atab
, “
Flexible solution-processable black-phosphorus-based optoelectronic memristive synapses for neuromorphic computing and artificial visual perception applications
,”
Adv. Mater.
(published online)
(
2023
).
49.
R.
Lamprecht
and
J.
LeDoux
, “
Structural plasticity and memory
,”
Nat. Rev. Neurosci.
5
,
45
(
2004
).
50.
H.
Yue
,
A.
Hu
,
Q.
Liu
,
H.
Tian
,
C.
Hu
,
X.
Ren
,
N.
Chen
,
C.
Ge
,
K.
Jin
, and
X.
Guo
, “
Graphene/SrTiO3 interface-based UV photodetectors with high responsivity
,”
Chin. Phys. B
30
,
038502
(
2021
).
51.
J.
Sun
,
S.
Zhang
,
T.
Zhan
,
Z.
Liu
,
J.
Wang
,
X.
Yi
,
J.
Li
,
P. M.
Sarro
, and
G.
Zhang
, “
A high responsivity and controllable recovery ultraviolet detector based on a WO3 gate AlGaN/GaN heterostructure with an integrated micro-heater
,”
J. Mater. Chem. C
8
,
5409
(
2020
).
52.
R.
Huang
,
T.
Liu
,
Y.
Zhao
,
Y.
Zhu
,
Z.
Huang
,
F.
Li
,
J.
Liu
,
L.
Zhang
,
S.
Zhang
,
A.
Dingsun
, and
H.
Yang
, “
Angular dependent XPS study of surface band bending on Ga-polar n-GaN
,”
Appl. Surf. Sci.
440
,
637
(
2018
).
53.
K. A.
Hatch
,
D. C.
Messina
,
H.
Fu
,
K.
Fu
,
Y.
Zhao
, and
R. J.
Nemanich
, “
External charge compensation in etched gallium nitride measured by x-ray photoelectron spectroscopy
,”
J. Appl. Phys.
131
,
185301
(
2022
).
54.
X.
Xu
,
C.
Liu
,
Z.
Sun
,
T.
Cao
,
Z.
Zhang
,
E.
Wang
,
Z.
Liu
, and
K.
Liu
, “
Interfacial engineering in graphene bandgap
,”
Chem. Soc. Rev.
47
,
3059
(
2018
).
55.
B. Y.
Zhang
,
T.
Liu
,
B.
Meng
,
X.
Li
,
G.
Liang
,
X.
Hu
, and
Q. J.
Wang
, “
Broadband high photoresponse from pure monolayer graphene photodetector
,”
Nat. Commun.
4
,
1811
(
2013
).
56.
S. M.
Kwon
,
S.-H.
Kang
,
S. S.
Cho
,
Y.-W.
Jang
,
S.-J.
Nam
,
J.
Kim
,
U.-C.
Moon
,
J.-W.
Shin
,
Y.-H.
Kim
, and
S. K.
Park
, “
Bidirectionally modulated synaptic plasticity with optically tunable ionic transistors
,”
ACS Appl. Electron. Mater.
4
,
2629
(
2022
).
57.
K.
He
,
Y.
Liu
,
J.
Yu
,
X.
Guo
,
M.
Wang
,
L.
Zhang
,
C.
Wan
,
T.
Wang
,
C.
Zhou
, and
X.
Chen
, “
Artificial neural pathway based on a memristor synapse for optically mediated motion learning
,”
ACS Nano
16
,
9691
(
2022
).
58.
P. Y.
Chen
,
X.
Peng
, and
S.
Yu
, “
NeuroSim plus: An integrated device-to-algorithm framework for benchmarking synaptic devices and array architectures
,” in
63rd IEEE Annual International Electron Devices Meeting (IEDM)
(
IEEE
,
2017
).
59.
C.
Han
,
X.
Han
,
J.
Han
,
M.
He
,
S.
Peng
,
C.
Zhang
,
X.
Liu
,
J.
Gou
, and
J.
Wang
, “
Light-stimulated synaptic transistor with high PPF feature for artificial visual perception system application
,”
Adv. Funct. Mater.
32
,
2113053
(
2022
).
60.
T.
Ahmed
,
M.
Tahir
,
M. X.
Low
,
Y.
Ren
,
S. A.
Tawfik
,
E. L. H.
Mayes
,
S.
Kuriakose
,
S.
Nawaz
,
M. J. S.
Spencer
,
H.
Chen
,
M.
Bhaskaran
,
S.
Sriram
, and
S.
Walia
, “
Fully light-controlled memory and neuromorphic computation in layered black phosphorus
,”
Adv. Mater.
33
,
2004207
(
2021
).

Supplementary Material