Startseite Solar-blind ultraviolet detection based on TiO2 nanoparticles decorated graphene field-effect transistors
Artikel Open Access

Solar-blind ultraviolet detection based on TiO2 nanoparticles decorated graphene field-effect transistors

  • Shasha Li ORCID logo , Tao Deng EMAIL logo , Yang Zhang , Yuning Li , Weijie Yin , Qi Chen und Zewen Liu
Veröffentlicht/Copyright: 26. April 2019
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

Sensitive solar-blind ultraviolet (UV) photodetectors are important to various military and civilian applications, such as flame sensors, missile interception, biological analysis, and UV radiation monitoring below the ozone hole. In this paper, a solar-blind UV photodetector based on a buried-gate graphene field-effect transistor (GFET) decorated with titanium dioxide (TiO2) nanoparticles (NPs) was demonstrated. Under the illumination of a 325-nm laser (spot size ~2 μm) with a total power of 0.35 μW, a photoresponsivity as high as 118.3 A/W was obtained, at the conditions of zero gate bias and a source-drain bias voltage of 0.2 V. This photoresponsivity is over 600 times higher than that of a recently reported solar-blind UV photodetector based on graphene/vertical Ga2O3 nanowire array heterojunction (0.185 A/W). Experiments showed that the photoresponsivity of the TiO2 NPs decorated GFET photodetectors can be further enhanced by increasing the source-drain bias voltage or properly tuning the gate bias voltage. Furthermore, the photoresponse time of the TiO2 NPs decorated GFET photodetectors can also be tuned by the source-drain bias and gate bias. This study paves a simple and feasible way to fabricate highly sensitive, cost-efficient, and integrable solar-blind UV photodetectors.

1 Introduction

Ultraviolet (UV) photodetectors are important for many applications such as space communication, flame detection, military surveillance, industrial quality control, and environmental monitoring [1], [2], [3], [4], [5], [6], [7], [8], [9], [10]. Due to the excellent electrical and optical properties, such as ultrafast carrier mobility [11], [12], high conductivity, ultra-wide spectral range (from UV to terahertz) [13], [14], [15] and tunable optical properties via electrostatic doping [16], [17], graphene has attracted enormous attention as a promising candidate for high-performance photodetectors. Especially, the pronounced photon absorption near the saddle-point singularity makes graphene well suited for UV photodetectors [18]. However, the photoresponsivity of pure graphene-based photodetectors is still limited within 1–2×10−2 A W−1 [19], [20], [21], owing to the intrinsic low optical absorption and short minority carrier lifetime of graphene [22].

A feasible way to increase the photoresponsivity of graphene-based photodetectors is to construct Schottky-diode-like graphene-semiconductor heterojunctions [23], [24], [25], [26], [27]. Using a graphene-silicon heterojunction device, a photoresponsivity about 0.1 A/W was obtained in the UV region [23]. By spin-coating a layer of titanium dioxide (TiO2) nanoparticles (NPs, size, 3–5 nm) [24] or Ag NPs [25] on such devices for UV light absorption, the photoresponsivity could be further increased by up to ~100%. To further increase the photoresponsivity of graphene-based UV photodetectors, graphene membranes were integrated with various semiconducting light harvesters, such as transition-metal dichalcogenides [28], [29], fullerene (C60) [30], organic molecules [31], and quantum-dots (QDs) [32], [33], based on the photogating effect. For example by synthesizing a layer of silicon QDs doped with boron on a back-gated graphene field-effect transistor (GFET), a photoresponsivity up to ~108 A/W and a detectivity of ~1012 Jones were obtained at 375 nm [33]. However, as most of the existing light harvesters have modest bandgaps, the corresponding graphene UV photodetectors are not solar-blind, limiting their applications to some military and civilian fields such as missile interception, biological analysis, and UV radiation monitoring below the ozone hole [26]. Furthermore, the synthesis techniques of some light harvesters and the processes of integrating them with graphene are complex and require special equipment, leading to high cost and unavailability. Therefore, sensitive, solar-blind, and cost-effective UV photodetectors are still highly desirable.

In this paper, a simple and feasible technique was demonstrated to realize highly sensitive and solar-blind UV photodetectors, by decorating a buried-gated GFET with solution-synthesized TiO2 NPs. The buried-gate structure allows tunable optical properties via a small gate voltage, while the solution-synthesized TiO2 NPs, which are solar-blind, nontoxic, low-cost and long-term stable [24], [34], [35], significantly increases the photoresponsivity. Under the illumination of a 325-nm laser with an intensity of 11 W/cm2, a photoresponsivity of 118.3 A/W and a detectivity of 1.75×1011 Jones were obtained, at the conditions of zero gate bias and a source-drain bias voltage of 0.2 V. Furthermore, the photoresponsivity and photoresponse time of the devices can be easily modulated by applying a small (≤1 V) gate bias or/and changing the source-drain bias voltages, to meet the needs in different application situations.

2 Experiments

The buried-gate GFETs were fabricated using a similar method shown in our previous work [36]. The buried chromium (Cr)/gold (Au) gate electrodes with thicknesses of 10 and 30 nm were deposited on a substrate consisted of 200-nm-thick silicon nitride (Si3N4), 50-nm-thick aluminum (Al), and 500-μm-thick silicon (Si), using magnetron sputtering. Next, a layer of silicon dioxide (SiO2) with the thickness of 30 nm was deposited by plasma-enhanced chemical vapor deposition, to form the dielectric layer. Then, monolayer graphene (ACS Material, LLC) was transferred onto the SiO2 layer and patterned using oxygen plasma etching. After this step, Cr/Au (10 nm/50 nm) source and drain electrodes were deposited on the graphene layer by electron beam evaporation and the buried-gate GFETs were obtained (Figure 1A).

Figure 1: Titanium dioxide (TiO2) nanoparticles (NPs) decorated buried-gate graphene field-effect transistors (GFETs).(A) and (B) Buried-gate GFETs before and after the decoration of TiO2 NPs on the surface of graphene. (C) Scanning electron microscopy micro-image of a 2×5 array of TiO2 NPs decorated GFETs. (D) Zoomed-in details of the graphene/TiO2 NPs conductive channel. (E) Atomic force microscopy micro-image of agglomerated TiO2 NPs (marked in red box in Figure 1D). (F) Typical Raman spectra of the GFETs with (red line) and without (blue line) TiO2 NPs.
Figure 1:

Titanium dioxide (TiO2) nanoparticles (NPs) decorated buried-gate graphene field-effect transistors (GFETs).

(A) and (B) Buried-gate GFETs before and after the decoration of TiO2 NPs on the surface of graphene. (C) Scanning electron microscopy micro-image of a 2×5 array of TiO2 NPs decorated GFETs. (D) Zoomed-in details of the graphene/TiO2 NPs conductive channel. (E) Atomic force microscopy micro-image of agglomerated TiO2 NPs (marked in red box in Figure 1D). (F) Typical Raman spectra of the GFETs with (red line) and without (blue line) TiO2 NPs.

The TiO2 NPs were synthesized using the sol-gel method [24]. Firstly, a mixed solution of TiCl4 (1 ml), ethanol (5 ml), and benzyl alcohol (35 ml) with a volume ratio of 2.4:12.2:85.4 was heated at 80°C for 6 h. The TiCl4, ethanol and benzyl alcohol were purchased from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. Then, the product was washed three times with diethyl ether (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) and followed by centrifugation at 12000 rpm (9200 g). for 10 min, to obtain the white TiO2 precipitate. Finally, the TiO2 precipitate was dispersed in ethanol to obtain an ethanol suspension of TiO2 NPs with a mass concentration of 4 mg/ml. The average diameter of the TiO2 NPs synthesized using this method is about 3–5 nm [24]. By dropping the suspension of TiO2 NPs (4 mg/ml) onto the surface of graphene of the buried-gate GFETs, and drying them in the air, TiO2 NPs decorated GFETs were obtained (Figure 1B).

Scanning electron microscopy (SEM) and atomic force microscopy (AFM) measurements were carried out to characterize the morphological and structural properties of the buried-gate GFETs before and after the TiO2 NPs decoration process. Raman spectroscopy and energy dispersive X-ray spectroscopy (EDS) were also performed to analyze the components of the devices. The electrical properties of the buried-gate GFETs with and without TiO2 NPs were investigated using a semiconductor parameter analyzer (B1500A, Keysight) and a probe station (Summit 12000, Cascade) at room temperature.

In order to investigate the photoelectrical properties of the TiO2 NPs decorated GFETs, the wafer was diced into small (5 mm×5 mm) chips, and bonded onto printed circuit boards with Au wires. In the photoelectrical experiments, a 325-nm laser beam with a spot size of ~2 μm, which was generated by a laser confocal Raman spectrometer (LabRam HR-800, Horiba JobinYvon), was illuminated to the conductive channel of the TiO2 NPs decorated GFETs. A precision source/measure unit (B2911A, Agilent) was used to supply the source-drain bias voltage and monitor the source-drain current of the GFETs, while a source-meter (2400, Keithley) was used to supply the gate bias. All photoelectrical measurements were performed at room temperature and under ambient conditions.

3 Results and discussion

The schematics of buried-gate GFETs before and after the decoration of TiO2 NPs are shown in Figure 1A,B, respectively. The dimension (length×width) of the graphene conductive channel is 30 μm×200 μm, while the lengths of the source, drain, and gate electrodes and the intervals between them are 10 μm. Figure 1C shows a SEM micro-image of a 2×5 array of TiO2 NPs decorated GFETs, indicating the potential for large-scale integration of the devices, which is important in sensing and imaging areas [13], [30], [37]. From the zoomed-in SEM micro-image of a TiO2 NPs decorated GFET (Figure 1D), the source, drain, and gate electrodes are well constructed, and the conductive channel is relatively smooth, except for several agglomerates. An EDS measurement was performed on one of the agglomerates in the center of the conductive channel (labeled by red box in Figure 1D). Significant Ti and O peaks were observed in the EDS spectroscopy (Figure S1C), indicating the agglomerates consisted of TiO2 NPs. More EDS measurements were carried out at different positions in the conductive channel, and demonstrated that the whole channel was coated with a thin layer of TiO2 NPs (Figure S1C–S1F).

Atomic force microscopy measurements were also performed to characterize the morphology of the TiO2 NPs decorated graphene conductive channel, as shown in Figure 1E. It can be seen that the bottom diameter of the TiO2 NPs agglomerate is about 2 μm and the height is about 1 μm. The average thickness of the TiO2 NP layer far away from the agglomerate is less than 500 nm. According to the work reported by Zhu et al. [24], the optimal thickness of the TiO2 NP layer for UV light harvesting is between 250 and 500 nm, and the agglomerates of TiO2 NPs will decrease the UV light harvesting efficiency. Therefore, the TiO2 NPs decoration process should be optimized carefully. We further characterized the buried-gate GFETs with and without TiO2 NPs using the Raman spectroscopy and confirmed the presence of both graphene and TiO2 in the TiO2 NPs decorated GFETs (Figure 1F). For the buried-gate GFETs without TiO2 NPs, the characteristic G and 2D peaks of graphene locate at 1594 and 2686 cm−1, respectively. After the decoration of TiO2 NPs, both G and 2D peaks of graphene are slightly up-shifted, 1 cm−1 for the G peak and 5 cm−1 for the 2D peak. Simultaneously, the characteristic TiO2 peaks are observed at 153, 400, 508, and 640 cm−1 [38], [39], in the TiO2 NPs decorated GFETs.

Figure 2A shows the transfer curves of the buried-gate GFETs before and after the TiO2 NPs decoration. A typical ambipolar characteristic can be clearly observed in the GFET without TiO2 NPs, where the Dirac point, which is defined as the gate voltage (Vgs) at minimum conductance, locates at 9.8 V, indicating the graphene is p-type doped by H2O, O2, or other contaminations [40], [41]. After the TiO2 NPs decoration, no Dirac point is observed within the gate voltage range from −2 to 20 V, and the source-drain current decreases monotonously with the increase of gate voltage. It is worth noting that the slope of the transfer curve of the GFET with TiO2 NPs is smaller than that of the GFET without TiO2 NPs in the p-type linear region. The source-drain current (Ids) of the former is also smaller than that of the latter, under the same gate voltages, indicating the conductivity of the graphene degrades in the TiO2 NPs decoration process. These phenomena were opposite with the observation reported by Zheng et al. [35], and need to be carefully investigated in the future. From the enlarged transfer characteristic curve of the TiO2 NPs decorated GFET (Figure 2B), the gate modulation effect still exists, although it is weaker than that of the GFET without TiO2 NPs. In order to investigate the effect of TiO2 NPs on the carrier mobility (μ) and contact resistance (Rcontact) of the device, a model proposed by Kim et al. [42] was used to analyze the resistive behavior of the devices. Detailed information about the analysis can be found in Figure S2. For the GFET without TiO2 NPs, μ=764 cm2/(V s), and Rcontact=147 Ω. After coating TiO2 NPs, the carrier mobility of the GFET decreases to μ=690 cm2/(V s), and the contact resistance increases to Rcontact=405 Ω. These phenomena can be attributed to increased disorder caused by perturbations to the graphene surface during the coating process [43], [44]. Figure 2C,D display the output characteristics of the buried-gate GFETs without and with TiO2 NPs, respectively. Under a certain gate voltage, the source-drain current of the GFET without TiO2 NPs increases linearly with the increase of the gate voltage (Figure 2C). Similar phenomenon is also observed in the GFET with TiO2 NPs (Figure 2D). The source-drain current of the GFETs with and without TiO2 NPs also changes with the gate voltage, indicating reliable gate modulation capability.

Figure 2: Electrical properties of the GFETs before and after the TiO2 NPs decoration process.(A) Comparison of the transfer characteristics for GFETs with (red curve) and without (blue curve) TiO2 NPs. (B) The transfer characteristic of the TiO2 NPs decorated GFET. (C) and (D) Output characteristics of the GFETs without and with TiO2 NPs, respectively.
Figure 2:

Electrical properties of the GFETs before and after the TiO2 NPs decoration process.

(A) Comparison of the transfer characteristics for GFETs with (red curve) and without (blue curve) TiO2 NPs. (B) The transfer characteristic of the TiO2 NPs decorated GFET. (C) and (D) Output characteristics of the GFETs without and with TiO2 NPs, respectively.

The photoelectrical properties of the buried-gate GFETs with and without TiO2 NPs decoration were characterized using an experimental setup shown in Figure 3A. Figure 3B shows the on/off source-drain current (Ids) of the TiO2 NPs decorated GFET under the illumination of a 325-nm laser with a power of 3.47 mW, at the conditions of zero gate bias (Vgs) and a source-drain bias voltage (Vds) of 0.1 V. As mentioned above, the Dirac point of the TiO2 NPs decorated GFET is over than 20 V (Figure 2B), the carrier transport in the graphene channel of the TiO2 NPs decorated GFET under zero gate bias is hole-dominant. When the TiO2 NPs decorated GFET is exposed to UV light, the ground-state electrons in TiO2 NPs are activated into the excited state, producing a large amount of free electrons which are injected into the graphene layer (see the inset of Figure 3C). The photogenerated electrons from TiO2 NPs and photogating-induced electrons recombine with the dominant holes in the graphene channel, leading to a decreased conductivity. Thus, the Ids decreases immediately and significantly when the laser is turned on. Once the laser is turned off, the photogenerated electrons and holes in the TiO2 NPs recombine with each other, resulting in that the density of the dominant holes in the graphene channel recovers to its original level, so does the Ids. After several on-off cycles, the photocurrent, which is defined as the difference between the source-drain current under UV light illumination and the dark source-drain current, is well retained, demonstrating good reliability and reversibility of the devices. It is worth noting that there was no significant photocurrent observed under the 514-nm laser illumination (Figure S3), at the same conditions. These results indicate that the TiO2 NPs decorated GFET photodetector is solar-blind.

Figure 3: Photoresponse of the TiO2 NPs decorated GFETs.(A) Schematic of the experimental setup. (B) Temporal photocurrent response of the TiO2 NPs decorated GFETs, under zero gate bias (Vgs) and a source-drain voltage (Vds) of 0.1 V. (C) A single modulation cycle of temporal photocurrent response for the TiO2 NPs decorated GFET. The inset illustrates the operation mechanism of the device. (D) The laser position dependence of the photocurrent, under the conditions of Vgs=0 V and Vds=0.1 V. Inset: SEM micro-image of the device with 7 laser positions. A 325-nm laser was used in the experiments, the laser powers for (B) and (C) were 3.47 mW, while that for (D) was 1.74 mW.
Figure 3:

Photoresponse of the TiO2 NPs decorated GFETs.

(A) Schematic of the experimental setup. (B) Temporal photocurrent response of the TiO2 NPs decorated GFETs, under zero gate bias (Vgs) and a source-drain voltage (Vds) of 0.1 V. (C) A single modulation cycle of temporal photocurrent response for the TiO2 NPs decorated GFET. The inset illustrates the operation mechanism of the device. (D) The laser position dependence of the photocurrent, under the conditions of Vgs=0 V and Vds=0.1 V. Inset: SEM micro-image of the device with 7 laser positions. A 325-nm laser was used in the experiments, the laser powers for (B) and (C) were 3.47 mW, while that for (D) was 1.74 mW.

To investigate the photoresponse speed of the TiO2 NPs decorated GFET photodetectors, the rise/decay time is defined as the photocurrent increases/decreases from 10/90% to 90/10% of the stable photocurrent. From a zoomed-in single modulation cycle (Figure 3C), the rise time and the decay time are extracted to be 2.0 and 158.7 s, respectively. The photoresponse speed of the TiO2 NPs decorated GFET photodetector is relatively slow, which can be attributed to the long-lived charge trapping achieved by the spatial separation of photoexcited charges across the graphene/TiO2 interface [35], [39]. The lifetime of the trapped charges in the semiconductor light harvesters is influenced by many factors, such as defect types, trap-state density, and defect trap depths [45]. By varying the above factors, the charge trap lifetime can be significantly reduced [33]. Furthermore, the photoresponse speed of the TiO2 NPs decorated GFET photodetector can also be tuned by the source-drain bias and gate bias, which will be discussed below.

Thanks to the small spot size of the laser (~2 μm), scanning photocurrent experiments were performed by mechanically moving the TiO2 NPs decorated GFET photodetector step by step to change the laser beam position between the source and drain electrodes, as illustrated in the inset of Figure 3D. The experimental result (Figure 3D) demonstrates that photocurrent generation clearly arises from the graphene/TiO2 NPs hybrid channel rather than electrode/channel junctions or other regions, indicating the photogating effect dominates the photoresponse [46]. More experiments found that the laser position on the TiO2/graphene channel of the device affects not only the photocurrent but also the photoresponse speed, as shown in Figure S4. In the following experiments, the laser position was fixed in the middle of the channel.

Figure 4A shows temporal photoresponses of the TiO2 NPs decorated GFET under different source-drain bias voltages (Vds), at the conditions of an incident laser power of 1.74 mW (325 nm) and zero gate bias (Vgs=0 V). With the increase of the Vds, the photocurrent increases, as well as the dark current. Furthermore, the rise times (τon) of the photoresponses under different source-drain bias voltages of 0.1, 0.2, and 0.4 V are 3.23, 3.75, and 3.91 s, respectively. This result indicates that the photoresponse speed becomes slower with the increase of the source-drain bias voltage. The photoresponsivity (Rph) is defined as Rph=Iph/P, where P is total power of the incident laser. More experiments showed that the Rph increases almost linearly with the Vds, as shown in Figure 4B. When the Vds increases from 0.1 to 1 V, the Rph increases 7.2 times. Liu and Kar [47] and Liu et al. [48] also observed similar phenomena of the linear bias voltage dependence of photoresponsivity. The photoconductive gain can be calculated by using τ/ttr, where τ is the lifetime of the trapped charges, and ttr is the time for the transfer of carriers through the device channel [49]. As ttr=L2/(μVds), where L is the length of the device channel and μ is the carrier mobility, the photoconductive gain is proportional to Vds. The tunable photoresponsivity of photodetectors is important in some practical applications such as imaging for variable light intensity.

Figure 4: Photoelectrical properties of the TiO2 NPs decorated GFET photodetectors.(A) Photoresponses at different source-drain bias voltages (Vds), under the conditions of zero gate bias (Vgs=0 V) and an incident laser power of 1.74 mW. (B) The normalized photoresponsivity [Rph/Rph,0, where Rph,0 is the photoresponsivity (Rph) obtained at Vds=0.1 V] as a function of the Vds, under the conditions of an incident laser power of 1.74 mW and Vgs=0 V. (C) Photoresponses under different incident laser powers. (D) The photoresponsivity as a function of incident laser power, at Vds=0.1 and Vgs=0 V.
Figure 4:

Photoelectrical properties of the TiO2 NPs decorated GFET photodetectors.

(A) Photoresponses at different source-drain bias voltages (Vds), under the conditions of zero gate bias (Vgs=0 V) and an incident laser power of 1.74 mW. (B) The normalized photoresponsivity [Rph/Rph,0, where Rph,0 is the photoresponsivity (Rph) obtained at Vds=0.1 V] as a function of the Vds, under the conditions of an incident laser power of 1.74 mW and Vgs=0 V. (C) Photoresponses under different incident laser powers. (D) The photoresponsivity as a function of incident laser power, at Vds=0.1 and Vgs=0 V.

The photoresponses of the TiO2 NPs decorated GFET under different incident laser powers are plotted in Figure 4C. The source-drain bias voltage is 0.1 V, while the gate bias voltage is 0 V. It can be seen that the photocurrent rises with the increase of the incident laser power. Furthermore, the rise times (τon) of the photoresponses under different laser powers of 3.47 mW, 0.35 mW, and 34.7 μW are 1.1, 12.7, and 91.4 s, respectively, as shown in Table 1, indicating that higher laser power results in faster photoresponse speed. This phenomenon was also observed by Tan et al. in a black phosphorus carbide phototransistor based on the photogating effect [50].

Table 1:

The photoresponse times of the TiO2 nanoparticles (NPs) decorated graphene field-effect transistor (GFET) photodetector under 325-nm laser illuminations with different powers (Vgs=0 V, Vds=0.1 V).

Laser power (μW)3470174087035034.7
Photoresponse time (s)1.11.64.112.791.4

Figure 4D shows the photoresponsivity of the TiO2 NPs decorated GFET photodetector as a function of the incident laser power, under the conditions of zero gate bias and a source-drain bias of 0.1 V. As with most graphene-based phototransistors based on the photogating effect [47], [48], the photoresponsivity decreases at high incident photon densities due to the saturated absorption and the weakened built-in field [32]. A photoresponsivity as high as 60.3 A/W was obtained under the laser power of 0.35 μW (11.1 W/cm2). It is worth noting that the laser power mentioned here is the total incident laser power, rather than the laser power on the graphene conductive channel. By simply increasing the source-drain bias voltage from 0.1 to 0.2 V, the photoresponsivity of the TiO2 NPs decorated GFET can be increased to 118.3 A/W (Figure S5), which is much higher than that of the GFET before coating TiO2 NPs (0.78 μA/W, as shown in Figure S6). Furthermore, this photoresponsivity (118.3 A/W) is over 600 times higher than that of a recently reported solar-blind UV photodetector based on graphene/vertical Ga2O3 nanowire array heterojunction (0.185 A/W) [26]. We would like to point that the photoresponse time of our device now is in the level of several seconds to hundreds of seconds, which is much slower than that of the graphene/Ga2O3 photodetector (~8 ms) [26]. A major reason is that the graphene/Ga2O3 photodetector is a photodiode, while our device is a phototransistor. The typical photoresponse time of graphene-based phototransistors is in the level of several seconds, such as the Si quantum dot/graphene phototransistors reported by Ni et al. [33] and the graphene/MoS2 phototransistors reported by Roy et al. [51]. Other reasons for the slow photoresponse of our device are related to the photosensitive materials and their morphologies. For example, there are a lot of trap states in TiO2 NPs and the graphene/TiO2 interface, which trap photogenerated electrons, resulting in a slow photoresponse [24]. In order to improve the photoresponse speed of the TiO2 NPs decorated GFET photodetectors, gate pulses [52] may be employed to sweep out the trapped carriers of TiO2 NPs within an appropriate time frame in the future. Another important parameter of photodetector is the specific detectivity (D*), which determines the capability of a photodetector to detect a weak light signal, can be given as D=RA1/22eIdark [53], where R is photoresponsivity, A is the effective photoactive area of detector, e is the electron charge, and Idark is the dark current. Using the photoresponsivity of 60.3 A/W, a D*=1.24×1011 jones (1 jones=1 cm Hz1/2 W−1) is obtained, which is comparable to the best performances of reported graphene-based UV photodetectors [39], [54]. Importantly, the photoresponsivity and detectivity can be further enhanced at lower incident laser powers [27], [29], [53].

Temporal photoresponses of the TiO2 NPs decorated GFET under different gate bias voltages are plotted in Figure 5A. The magnitudes of the photocurrents under different gate bias voltages are different, so are the photoresponse speeds. The photocurrent under Vgs=–0.5 V is 52.7% larger than that under Vgs=–1 V, while the rise time of the photoresponse under Vgs=–0.5 V is 65.7% slower than that under Vgs=–1 V. More experiments showed that the photoresposivity of the TiO2 NPs decorated GFET can be modulated by gate bias voltages, as shown in Figure 5B. With the increase of the gate voltage from 0 to 1 V, the photoresponsivity increases first and then decreases the threshold gate voltage which is about 0.5 V. Similar characteristic is also observed in the negative gate voltage range. By applying a –0.5 V voltage to the buried-gate electrode, the photoresponsivity can be increased by ~20%. These results indicate both the photoresponsivity and the photoresponse speed can be modulated by the gate bias. Table 2 summarizes the reported high-performance UV photodetectors based on graphene. Obviously, the device in this work has extremely high photoresponsivity under small bias voltages, but the photoresponse speed needs to improve in the future.

Figure 5: The gate-voltage-dependent photoresponse of the TiO2 NPs decorated GFET photodetectors.(A) Temporal photocurrent responses of the TiO2 NPs decorated GFET photodetector at different gate bias voltages (Vgs), under the conditions of a laser power of 0.347 mW and Vds=0.1 V. (B) The gate-voltage-dependent photoresponsivity (Rph) of the TiO2 NPs decorated GFET photodetectors, where Rph,0 is the Rph at Vds=0.1 V and Vgs=0 V. The incident laser power is 0.347 mW.
Figure 5:

The gate-voltage-dependent photoresponse of the TiO2 NPs decorated GFET photodetectors.

(A) Temporal photocurrent responses of the TiO2 NPs decorated GFET photodetector at different gate bias voltages (Vgs), under the conditions of a laser power of 0.347 mW and Vds=0.1 V. (B) The gate-voltage-dependent photoresponsivity (Rph) of the TiO2 NPs decorated GFET photodetectors, where Rph,0 is the Rph at Vds=0.1 V and Vgs=0 V. The incident laser power is 0.347 mW.

Table 2:

Photoelectrical characteristics of typical ultraviolet (UV) photodetectors based on graphene.

RefDescriptionWavelengthResponsivityDetectivity (Jones)Rise time
[26]Graphene/vertical Ga2O3 nanowire array heterojunction254 nm0.185 A/W@–5 V9 ms (254 nm)
258 nm
[39]Graphene/TiO2 nanotube heterojunction365 nm~20 A/W@5 V1.9×1012>15 s
[34]Graphene/anatase-nanosheets Schottky junction254, 302, 365 nm0.5 A/W@–5 V140 ms (254 nm)

220 ms (302 nm)
[54]Graphene-TiO2 nanotube arrays heterojunction365 nm0.126 A/W@–0.5 V3.3×1011
[55]3D tubular GFETsUV, visible, MIR, THz1 A/W@0 V(UV)265 ns
[56]Graphene/β-Ga2O3 heterojunction254 nm12.8 A/W@–6 V1.3×10131.5 ms
[25]Ag NPs enhanced graphene/GaAs heterojunction325–980 nm0.186 A/W@0 V

(325 nm)
2.63×1013

(325 nm)
[27]Graphene β-Ga2O3 heterojunction254 nm39.3 A/W@20 V5.92×101394.83 s@4 V
[57]Graphene/n-type SiC Schottky325 nm2.18 A/W@–5 V>0.39 s and <4.46 s
This workTiO2 NPs decorated GFETs325 nm118.3 A/W@0.2 V1.75×1011>1.1 s

4 Conclusions

A highly sensitive, solar-blind, and cost-effective UV photodetector based on a buried-gate GFET decorated with TiO2 NPs was demonstrated in this paper. The TiO2 NPs significantly enhance the UV light absorption of the device. Under the illumination of a 325-nm laser with an incident power of 0.35 μW, a photoresponsivity as high as 118.3 A/W and a detectivity of ~1011 Jones were obtained, at zero gate bias and a source-drain bias voltage of 0.2 V. Importantly, the photoresponsivity can be further enhanced by increasing the source-drain bias voltage or properly tuning the gate bias voltage. Furthermore, the photoresponse speed of the TiO2 NPs decorated GFET photodetectors can also be tuned by the source-drain bias and gate bias.

Supplementary material: See Supplementary material for energy X-ray spectrum of the buried-gate TiO2 NPs GFETs, the carrier mobility and graphene-metal contact resistance of the buried-gate GFET with and without TiO2 NPs, the photoresponse of the device under the 514 nm illumination, the influence of the incident laser position on photoresponse time, the photoresponsivity of the device before coating TiO2 NPs, and the temporal photoresponse of the buried-gate TiO2 NPs GFET at Vds=0.2 V.

Award Identifier / Grant number: 61604009

Award Identifier / Grant number: 4164095

Funding statement: This research was supported by the National Natural Science Foundation of China (61604009), the Fundamental Research Funds for the Central Universities (2018JBM002), and the Beijing Municipal Natural Science Foundation (4164095).

References

[1] Zhang H, Babichev AV, Jacopin G, et al. Characterization and modeling of a ZnO nanowire ultraviolet photodetector with graphene transparent contact. J Appl Phys 2013;114:234505.10.1063/1.4854455Suche in Google Scholar

[2] Chen CS, Zhou P, Wang N, Ma Y, San HS. UV-assisted photochemical synthesis of reduced graphene oxide/ZnO nanowires composite for photoresponse enhancement in UV photodetectors. Nanomaterials 2018;8:26.10.3390/nano8010026Suche in Google Scholar PubMed PubMed Central

[3] Hajimazdarani M, Naderi N, Yarmand B, Kolahi A, Sangpour P. Enhanced optical properties of ZnS-rGO nanocomposites for ultraviolet detection applications. Ceramics Int 2018;44:17878–84.10.1016/j.ceramint.2018.06.260Suche in Google Scholar

[4] Chava VSN, Omar SU, Brown G, et al. Evidence of minority carrier injection efficiency >90% in an epitaxial graphene/SiC Schottky emitter bipolar junction phototransistor for ultraviolet detection. Appl Phys Lett 2016;108:043502.10.1063/1.4940385Suche in Google Scholar

[5] Tsukazaki A, Ohtomo A, Onuma T, et al. Repeated temperature modulation epitaxy for p-type doping and light-emitting diode based on ZnO. Nat Mater 2005;4:42–6.10.1038/nmat1284Suche in Google Scholar

[6] Wan Q, Li QH, Chen YJ, et al. Fabrication and ethanol sensing characteristics of ZnO nanowire gas sensors. Appl Phys Lett 2004;84:3654.10.1063/1.1738932Suche in Google Scholar

[7] Law M, Greene LE, Johnson JC, Saykally R, Yang PD. Nanowire dye-sensitized solar cells. Nat Mater 2005;4:455–9.10.1038/nmat1387Suche in Google Scholar PubMed

[8] Wang ZL, Song JH. Piezoelectric nanogenerators based on zinc oxide nanowire arrays. Science 2006;312:242–6.10.1126/science.1124005Suche in Google Scholar PubMed

[9] Prakash N, Singh M, Kumar G, et al. Ultrasensitive self-powered large area planar GaN UV-photodetector using reduced graphene oxide electrodes. Appl Phys Lett 2016;109:242102.10.1063/1.4971982Suche in Google Scholar

[10] Arora K, Goel N, Kumar M, Kumar M. Ultrahigh performance of self-powered β-Ga2O3 thin film solar-blind photodetector grown on cost-effective Si substrate using high-temperature seed layer. ACS Photonics 2018;5:2391–401.10.1021/acsphotonics.8b00174Suche in Google Scholar

[11] Dawlaty JM, Shivaraman S, Chandrashekhar M, Rana F, Spencer MG. Measurement of ultrafast carrier dynamics in epitaxial graphene. Appl Phys Lett 2008;92:042116.10.1063/1.2837539Suche in Google Scholar

[12] Brida D, Tomadin A, Manzoni C, et al. Ultrafast collinear scattering and carrier multiplication in graphene. Nat Commun 2013;4:1987.10.1038/ncomms2987Suche in Google Scholar PubMed

[13] Koppens FHL, Mueller T, Avouris P, Ferrari AC, Vitiello MS, Polini M. Photodetectors based on graphene, other two-dimensional materials and hybrid systems. Nat Nanotechnol 2014;9:780–93.10.1038/nnano.2014.215Suche in Google Scholar PubMed

[14] Long MS, Wang P, Fang HH, Hu WD. Progress, challenges, and opportunities for 2D material based photodetectors. Adv Funct Mater 2018;1803807.10.1002/adfm.201803807Suche in Google Scholar

[15] Liu N, Tian H, Schwartz G, Tok JBH, Ren TL, Bao ZN. Large-area, transparent, and flexible infrared photodetector fabricated using P-N junctions formed by N-doping chemical vapor deposition grown graphene. Nano Lett 2014;14:3702–8.10.1021/nl500443jSuche in Google Scholar PubMed

[16] Li ZQ, Henriksen EA, Jiang Z, et al. Dirac charge dynamics in graphene by infrared spectroscopy. Nat Phys 2008;4:532–5.10.1038/nphys989Suche in Google Scholar

[17] Wang F, Zhang YB, Tian CS, et al. Gate-variable optical transitions in graphene. Science 2008;320:206–9.10.1126/science.1152793Suche in Google Scholar PubMed

[18] Mak KF, Shan J, Heinz TF. Seeing many-body effects in single-and few-layer graphene: observation of two-dimensional saddle-point excitons. Phys Rev Lett 2011;106:046401.10.1103/PhysRevLett.106.046401Suche in Google Scholar PubMed

[19] Echtermeyer TJ, Britnell L, Jasnos PK, et al. Strong plasmonic enhancement of photovoltage in graphene. Nat Commun 2011;2:458.10.1038/ncomms1464Suche in Google Scholar PubMed

[20] Engel M, Steiner M, Lombardo A, et al. Light-matter interaction in a microcavity-controlled graphene transistor. Nat Commun 2012;3:906.10.1038/ncomms1911Suche in Google Scholar PubMed PubMed Central

[21] Furchi M, Urich A, Pospischil A, et al. Microcavity-integrated graphene photodetector. Nano Lett 2012;12:2773–7.10.1021/nl204512xSuche in Google Scholar PubMed PubMed Central

[22] Breusing M, Ropers C, Elsaesser T. Ultrafast carrier dynamics in graphite. Phys Rev Lett 2009;102:086809.10.1103/PhysRevLett.102.086809Suche in Google Scholar PubMed

[23] An XH, Liu FZ, Jung YJ, Kar S. Tunable graphene-silicon heterojunctions for ultrasensitive photodetection. Nano Lett 2013;13:909–16.10.1021/nl303682jSuche in Google Scholar PubMed

[24] Zhu M, Zhang L, Li XM, et al. TiO2 enhanced ultraviolet detection based on a graphene/Si Schottky diode. J Mater Chem A 2015;3:8133–8.10.1039/C5TA00702JSuche in Google Scholar

[25] Lu YH, Feng SR, Wu ZQ, et al. Broadband surface plasmon resonance enhanced self-powered graphene/GaAs photodetector with ultrahigh detectivity. Nano Energy 2018;47:140–9.10.1016/j.nanoen.2018.02.056Suche in Google Scholar

[26] He T, Zhao YK, Zhang XD, et al. Solar-blind ultraviolet photodetector based on graphene/vertical Ga2O3 nanowire array heterojunction. Nanophotonics 2018;7:1557–62.10.1515/nanoph-2018-0061Suche in Google Scholar

[27] Kong WY, Wu GA, Wang KY, et al. Graphene-β-Ga2O3heterojunction for highly sensitive deep UV photodetector application. Adv Mater 2016;28:10725–31.10.1002/adma.201604049Suche in Google Scholar PubMed

[28] Mehew JD, Unal S, Alonso ET, et al. Fast and highly sensitive ionic-polymer-gated WS2-graphene photodetectors. Adv Mater 2017;29:1700222.10.1002/adma.201700222Suche in Google Scholar PubMed

[29] Iqbal MZ, Khan S, Siddique S. Ultraviolet-light-driven photoresponse of chemical vapor deposition grown molybdenum disulfide/graphene heterostructured FET. Appl Surf Sci 2018;459:853–9.10.1016/j.apsusc.2018.08.027Suche in Google Scholar

[30] Qin SC, Chen XQ, Du QQ, et al. Sensitive and robust ultraviolet photodetector array based on self-assembled graphene/C60 hybrid films. ACS Appl Mater Interfaces 2018;10:38326–33.10.1021/acsami.8b11596Suche in Google Scholar PubMed

[31] Liu XL, Luo XG, Nan HY, et al. Epitaxial ultrathin organic crystals on graphene for high efficiency phototransistors. Adv Mater 2016;28:5200–5.10.1002/adma.201600400Suche in Google Scholar PubMed

[32] Konstantatos G, Badioli M, Gaudreau L, et al. Hybrid graphene-quantum dot phototransistors with ultrahigh gain. Nat Nanotechnol 2012;7:363–8.10.1038/nnano.2012.60Suche in Google Scholar PubMed

[33] Ni ZY, Ma LL, Du SC, et al. Plasmonic silicon quantum dots enabled high-sensitivity ultra-broadband photodetection of graphene-based hybrid phototransistors. ACS Nano 2017;11:9854–62.10.1021/acsnano.7b03569Suche in Google Scholar PubMed

[34] Liu Y, Cai HL, Wang FF, et al. Graphene on {116} faceted monocrystalline anatase nanosheet array for ultraviolet detection. Nanoscale 2018;10:3606–12.10.1039/C7NR08037ASuche in Google Scholar PubMed

[35] Zheng KH, Meng FB, Jiang L, Yan QY, Hng HH, Chen XD. Visible photoresponse of single-layer graphene decorated with TiO2 nanoparticles. Small 2013;9:2076–80.10.1002/smll.201202885Suche in Google Scholar PubMed

[36] Su F, Zhang ZH, Li SS, Li PA, Deng T. Long-term stability of photodetectors based on graphene field-effect transistors encapsulated with Si3N4 layers. Appl Surf Sci 2018;459:164–70.10.1016/j.apsusc.2018.07.208Suche in Google Scholar

[37] Tian H, Yang Y, Xie D, et al. Wafer-scale integration of graphene-based electronic, optoelectronic and electroacoustic devices. Sci Rep 2014;4:3598.10.1038/srep03598Suche in Google Scholar PubMed PubMed Central

[38] Chen DJ, Zou LL, Li SX, Zheng FY. Nanospherical like reduced graphene oxide decorated TiO2 nanoparticles: an advanced catalyst for the hydrogen evolution reaction. Sci Rep 2016;6:20335.10.1038/srep20335Suche in Google Scholar PubMed PubMed Central

[39] Noothongkaew S, Thumthan O, An KS. Minimal layer graphene/TiO2 nanotube membranes used for enhancement of UV photodetectors. Mater Lett 2018;218:274–9.10.1016/j.matlet.2018.02.033Suche in Google Scholar

[40] Wang ZG, Chen YF, Li PJ, et al. Effects of methane flux on structural and transport properties of CVD-grown graphene films. Vacuum 2012;86:895–8.10.1016/j.vacuum.2011.05.011Suche in Google Scholar

[41] Srivastava A, Galande C, Ci L, et al. Novel liquid precursor-based facile synthesis of large-area continuous, single, and few-layer graphene films. Chem Mater 2010;22:3457–61.10.1021/cm101027cSuche in Google Scholar

[42] Kim S, Nah J, Jo I, et al. Realization of a high mobility dual-gated graphene field-effect transistor with Al2O3 dielectric. Appl Phys Lett 2009;94:062107.10.1063/1.3077021Suche in Google Scholar

[43] Farmer DB, Chiu HY, Lin YM, Jenkins KA, Xia FN, Avouris P. Utilization of a buffered dielectric to achieve high field-effect carrier mobility in graphene transistors. Nano Lett 2009;9:4474–8.10.1021/nl902788uSuche in Google Scholar PubMed

[44] Ishigami M, Chen JH, Cullen WG, Fuhrer MS, Williams ED. Atomic structure of graphene on SiO2. Nano Lett 2007;7:1643–8.10.1021/nl070613aSuche in Google Scholar PubMed

[45] Zhang YZ, Liu T, Meng B, et al. Broadband high photoresponse from pure monolayer graphene photodetector. Nat Commun 2013;4:1811.10.1038/ncomms2830Suche in Google Scholar PubMed

[46] Xia FN, Mueller T, Golizadeh-Mojarad R, et al. Photocurrent imaging and efficient photon detection in a graphene transistor. Nano Lett 2009;9:1039–44.10.1021/nl8033812Suche in Google Scholar PubMed

[47] Liu FZ, Kar S. Quantum carrier reinvestment-induced ultrahigh and broadband photocurrent responses in graphene-silicon junctions. ACS Nano 2014;8:10270–9.10.1021/nn503484sSuche in Google Scholar PubMed

[48] Liu YD, Wang FQ, Wang XM, et al. Planar carbon nanotube-graphene hybrid films for high-performance broadband photodetectors. Nat Commun 2015;6:8589.10.1038/ncomms9589Suche in Google Scholar PubMed PubMed Central

[49] Kufer D, Konstantatos G. Photo-FETs: phototransistors enabled by 2D and 0D nanomaterials. ACS Photonics 2016;3:2197–210.10.1021/acsphotonics.6b00391Suche in Google Scholar

[50] Tan WC, Huang L, Ng RJ, et al. A black phosphorus carbide infrared phototransistor. Adv Mater 2018;30:1705039.10.1002/adma.201705039Suche in Google Scholar PubMed

[51] Roy K, Padmanabhan M, Goswami S, et al. Graphene-MoS2 hybrid structures for multifunctional photoresponsive memory devices. Nat Nanotechnol 2013;8:826–30.10.1038/nnano.2013.206Suche in Google Scholar PubMed

[52] Lopez-Sanchez O, Lembke D, Kayci M, Radenovic A, Kis A. Ultrasensitive photodetectors based on monolayer MoS2. Nat Nanotechnol 2013;8:497–501.10.1038/nnano.2013.100Suche in Google Scholar PubMed

[53] Guo N, Gong F, Liu JK, et al. Hybrid WSe2−In2O3 phototransistor with ultrahigh detectivity by efficient suppression of dark currents. ACS Appl Mater Interfaces 2017;9:34489–96.10.1021/acsami.7b10698Suche in Google Scholar PubMed

[54] Zhang DY, Ge CW, Wang JZ, Zhang TF, Wu YC, Liang FX. Single-layer graphene-TiO2 nanotubes array heterojunction for ultraviolet photodetector application. Appl Surf Sci 2016;387:1162–8.10.1016/j.apsusc.2016.07.041Suche in Google Scholar

[55] Deng T, Zhang ZH, Liu YX, et al. Three-dimensional graphene field-effect transistors as high-performance photodetectors. Nano Lett 2019;19:1494–503.10.1021/acs.nanolett.8b04099Suche in Google Scholar PubMed

[56] Lin RC, Zheng W, Zhang D, et al. High-performance graphene/β-Ga2O3heterojunction deep-ultraviolet photodetector with hot-electron excited carrier multiplication. ACS Appl Mater Interfaces 2018;10:22419–26.10.1021/acsami.8b05336Suche in Google Scholar PubMed

[57] Yang JW, Gou LW, Gou YL, Hu WJ, Zhang ZS. Epitaxial graphene/SiC Schottky ultraviolet photodiode with orders of magnitude adjustability in responsivity and response speed. Appl Phys Lett 2018;112:103501.10.1063/1.5019435Suche in Google Scholar


Supplementary Material

The online version of this article offers supplementary material (https://doi.org/10.1515/nanoph-2019-0060).


Received: 2019-02-26
Revised: 2019-04-02
Accepted: 2019-04-06
Published Online: 2019-04-26

©2019 Tao Deng et al., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 Public License.

Artikel in diesem Heft

  1. Review Articles
  2. Non-Markovian features in semiconductor quantum optics: quantifying the role of phonons in experiment and theory
  3. Gradient index phononic crystals and metamaterials
  4. Information metamaterials – from effective media to real-time information processing systems
  5. Nonradiating photonics with resonant dielectric nanostructures
  6. Single-photon sources with quantum dots in III–V nanowires
  7. Plasmon-enhanced solar vapor generation
  8. Research Articles
  9. Polymorph-induced photosensitivity change in titanylphthalocyanine revealed by the charge transfer integral
  10. Topological plasmonic edge states in a planar array of metallic nanoparticles
  11. Laser-written colours on silver: optical effect of alumina coating
  12. Bismuth-based metamaterials: from narrowband reflective color filter to extremely broadband near perfect absorber
  13. Vortex index identification and unidirectional propagation in Kagome photonic crystals
  14. A compact structure for realizing Lorentzian, Fano, and electromagnetically induced transparency resonance lineshapes in a microring resonator
  15. Low- and high-order nonlinear optical properties of Ag2S quantum dot thin films
  16. Monolithic waveguide laser mode-locked by embedded Ag nanoparticles operating at 1 μm
  17. Fabrication of highly homogeneous and controllable nanogratings on silicon via chemical etching-assisted femtosecond laser modification
  18. Quantitative analysis and modeling of line edge roughness in near-field lithography: toward high pattern quality in nanofabrication
  19. WDM-compatible multimode optical switching system-on-chip
  20. Solar-blind ultraviolet detection based on TiO2 nanoparticles decorated graphene field-effect transistors
  21. Measuring circular phase-dichroism of chiral metasurface
  22. Highly efficient plasmonic nanofocusing on a metallized fiber tip with internal illumination of the radial vector mode using an acousto-optic coupling approach
  23. Tunable Brillouin and Raman microlasers using hybrid microbottle resonators
  24. Importance of higher-order multipole transitions on chiral nearfield interactions
Heruntergeladen am 4.11.2025 von https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2019-0060/html
Button zum nach oben scrollen