Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Enhancement in the photonic response of ZnO nanorod–gated AlGaN/GaN HEMTs with N2O plasma treatment

Open Access Open Access

Abstract

We demonstrate an improvement in the photoresponse characteristics of ultraviolet (UV) photodetectors (PDs) using the N2O plasma-treated ZnO nanorod (NR) gated AlGaN/GaN high electron mobility transistor (HEMT) structure. The PDs fabricated with ZnO NRs plasma-treated for 6 min show superior performance in terms of responsivity (∼1.54×10 5 A/W), specific detectivity (∼ 4.7×1013 cm·Hz−1/2/W), and on/off current ratio (∼40). These improved performance parameters are the best among those from HEMT-based PDs reported to date. Photoluminescence analysis shows a significant enhancement in near band edge emission due to the effective suppression of native defects near the surface of ZnO NRs after plasma treatment. As our X-ray photoelectron spectroscopy reveals a very high O/Zn ratio of ∼0.96 from the NR samples plasma-treated for 6 min, the N2O plasma radicals also show a clear impact on ZnO stoichiometry. From our X-ray diffraction analysis, the plasma-treated ZnO NRs show much greater improvement in (002) peak intensity and degree of (002) orientation (∼0.996) than those of as-grown NRs. This significant enhancement in (002) degree of orientation and stoichiometry in ZnO nano-crystals contribute to the enhancement in photoresponse characteristics of the PDs.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Wide band gap (WBG) materials such as ZnO (∼3.37 eV) and AlxGa1-xN (∼3.4 at x = 0 to ∼6.2 eV at x = 1) have been highlighted in recent decades for their high responsivity and solar blind ultraviolet (UV) light photodetector (PD) applications [13]. Moreover, due to their excellent chemical and thermal stability, they could be one of potential choices for the future applications, such as early detection of flame [4] and UV fluorescence tissue diagnostics in-vivo for human tissues [5]. Utilization of the AlGaN/GaN heterojunction with two dimensional electron gas (2-DEG) at the hetero-interface has especially leaded to high-performance UV PDs with cut-off wavelength down to 360 nm using a gateless high electron mobility transistor (HEMT) structure [6,7]. However, these HEMT-based PD structures could not exhibit desirable on/off current ratio (photocurrent/dark-current) and sensitivity despite of their impressive device performance of high responsivity. For example, UV PDs with ZnO nanorod (NR) gated AlGaN/GaN HEMT structure showed an extremely high responsivity of ∼105 A/W, but they revealed fairly low on/off current ratio of ∼4 when the conventional mesa isolation scheme was used [8].

It has been shown that the UV photoresponse characteristics of ZnO NR-gated HEMTs can be greatly influenced by the dimensional characteristics, such as gate length, gate width, gate capacitance, and channel geometry [8,9]. In this type of device, the photonic response especially depends on the chemisorption kinetics of the oxygen ion (O2) at the surface of UV-absorbing ZnO nanostructures [10]. Chemisorption physics of O2 ions at the ZnO surface was also extensively examined by many earlier studies [11,12], and it was known to be affected by a variety of factors, such as free electron concentration and the defect density near the surface of ZnO with native n-type doping [1315]. Not many research efforts have been made to date on investigating the effect of ZnO NR surface modification on the photoresponse characteristics of AlGaN/GaN HEMT-based PDs to promote the structural quality of ZnO surface. For instance, surface modification methods for the ZnO thin films or nanostructures grown on the planar devices based on the diode or homojunction structures have shown to be effective in terms of various aspects of device performance, such as responsivity, response speed, and sensitivity when plasma treated by O2 or NH3 and post-annealed in modified ambient conditions [2,1620].

Solution-based hydrothermal method [21] used in this work to grow ZnO NR crystals is known to deliver various kinds of extrinsic and intrinsic defects [22]. In the recent decades, much effort has been made on improving the crystalline quality of the ZnO NRs by plasma treatments using conventional gases, such as H2, N2, NH3, and O2, for various ZnO passive-type PDs [19,2328]. N2O plasma treatment was also reported in thin film transistors (TFTs) of the bottom-gate configuration to suppress the intrinsic defects, such as oxygen vacancies of ZnO [29]. Rapid thermal annealing and subsequent N2O plasma treatment increased the carrier mobility and on/off current ratio by two order of the magnitude with an improvement in crystalline quality of the thin film channel.

In this work, the influence of plasma treatment on nanocrystalline structure, photo-emission properties, intrinsic defects, and stoichiometry in ZnO crystals were examined using a variety of surface characterizations. We also investigated the effect of N2O plasma treatment on ZnO NRs grown atop the gate area (2×100 µm2) of AlGaN/GaN HEMT PDs. The performance parameters of the PDs fabricated with NRs plasma-treated under different conditions were discussed with our material characterizations.

2. Experiments

Figure 1(a) shows the schematic illustrations of the HEMT-based PD structures fabricated in this work. For this, epitaxial structure of the GaN buffer/3000-4000 nm GaN channel/20 nm Al0.25Ga0.75N barrier/1.25 nm GaN cap layer from the bottom were grown on a (111) Si substrate by using the metal organic chemical vapor deposition. Active mesa region was defined by an optical lithography pattern of a 100×40 µm2 dimension. The mesa region was then isolated by etching a depth of 100 nm from the AlGaN/GaN epitaxial layer in a reactive ion chamber (RIE) using BCl3 and Cl2 gases. The ohmic contacts were achieved by the subsequent deposition of Ti/Al/Ni/Au layers (30/180/40/150 nm) followed by a rapid thermal annealing at 900 °C for 35 s in nitrogen ambient. 100 nm thick silicon nitride (Si3N4) passivation layer was then deposited by using a plasma-enhanced chemical vapor deposition (PECVD). After the completion of gateless HEMT structure to this point, ZnO NRs were grown on the gate region. For this patterned growth, the passivation layer was first selectively etched from the gate area (2×100 µm2) using CF4 plasma in our RIE system. For the ZnO NRs growth, a seed layer (SL) solution was prepared by 0.2 M zinc acetate dehydrate in 30 ml 1-propanol. The SL solution was then spun onto the gate area at 3000 rpm, and this spin coating step was repeated 15 times with the subsequent prebake (120 °C, 60 s) for each iteration to ensure achieving a final SL thickness of 20 nm. After the deposition of SLs, all the samples were annealed at 300 °C for 1 hr on a hotplate in air ambient [30].

 figure: Fig. 1.

Fig. 1. (a) Schematic illustrations (untreated and plasma-treated), (b) top-view optical image, and (c) top-view SEM micrograph of the fabricated ZnO NR-gated AlGaN/GaN HEMT. (d) Top-view (left), expanded top-view (top-right), and cross-sectional SEM view (bottom right) of as-grown ZnO NRs.

Download Full Size | PDF

In the next step, the NRs were grown in an equimolar (25 mM) aqueous solution of zinc nitrate hexahydrate (Zn(NO3)2·6H2O, 99%) and hexamethylenetetramine (HMTA, C6H12N4, 99.5%). After the aqueous solution was prepared in sealable glass beaker, the samples were immersed upside down into the surface of the growth solution. The beaker was then sealed with the aluminum foil and placed on a hotplate at 90 °C for 6 hr. The samples were finally washed by de-ionized water and dried at room temperature. We synthesized the NRs prepared under four different conditions including a control sample (CS) and N2O plasma-treated samples for three different time intervals of 1, 3, and 6 min in a reactive ion plasma chamber at 50 mTorr base pressure, N2O gas flow rate of 100 sccm, and a RF power of 100 W.

Top-view morphologies of the ZnO NRs and fabricated PDs were characterized by optical microscopy and field emission scanning electron microscopy (FE-SEM, S-4800, Hitachi). The crystalline structures and optical properties of the NR samples were respectively characterized by X-ray diffraction (XRD, D8Advance spectrometer of Bruker AXS with Cu Kα 1.540 Å radiation) and room temperature (RT) photoluminescence (PL, RPM 2000 Accent). The oxygen stoichiometry in the chemical bonding for each NR sample was investigated by X-ray photoelectron spectroscopy (XPS, Thermo-fisher scientific, NEXSA) spectrometer with an X-ray source gun of 400 µm spot size. I-V characteristics of the devices were measured by a Keithely SMU under both dark and UV illumination conditions. The spectral response of the HEMT-based PDs were measured by using a monochromatic light source with a wavelength range of 290-700 nm. A broad band Xenon lamp (200-1100 nm, 300 W) light was filtered and chopped into monochromatic light using a computer program controlled diffraction grating. Uniform monochromatic light was passed through a frequency chopper wheel at 30 Hz and illuminated on to the sample. The frequency of monochromatic light was fed to a lock-in amplifier, and the photocurrents from the devices were recorded at each wavelength (290-700 nm) and analyzed.

3. Results and discussion

As shown in Figs. 1(b) and 1(c), ZnO NRs were grown on 2×100 µm2 gate area of the gateless AlGaN/GaN HEMT structure fabricated with source and drain electrodes partially covered by Si3N4 passivation. Top inset of Fig. 1(c) demonstrates the growth profile of NRs in an expanded view with a defined gate length of ∼2 µm. Shown in Fig. 1(d) are the vertically aligned growths of NRs with diameters ranging from 70 to 80 nm (see top right inset) and an average length of ∼1.4 µm (see bottom right inset).

UV photo-detection of ZnO based PDs is associated with the desorption of oxygen (O2) molecules which are chemisorbed at the surface of ZnO under dark condition [9,15]. As illustrated in Fig. 2(a), O2 molecules are adsorbed near the surface of the ZnO NRs by scavenging the free electrons from the surface volume of NRs and thereby forming O2 ions under dark condition. Due to the adsorption of O2 ions, the surface depletion region is expanded, and it produces a negative potential in the gate region of the AlGaN/GaN HEMT. The development of this negative potential at the surface of ZnO NRs (UV light-adsorbing structure) pushes the electron in 2-DEG deep into the GaN channel layer and reduces the free electron concentration. This process decreases the channel conductance and the drain current of the HEMT-based PD structure under dark condition as shown in Fig. 2(a). Upon the irradiation of UV light, electron-hole pairs are generated near the surface and in the bulk of ZnO NRs. While the holes recombine with the adsorbed O2 ions, desorbed O2 molecules migrate from the NR surface into the atmosphere. Consequently, the negative gate potential at the surface of UV absorbing structure is reduced, and it increases the carrier concentration in 2-DEG channel as shown in Fig. 2(b). This reaction will eventually increase the drain current under the UV illumination.

 figure: Fig. 2.

Fig. 2. Illustration of O2 (a) adsorption in the dark condition (b) and desorption under UV illumination taking place on the ZnO NR surface. Free electron concentration in two-dimensional electron gas (2-DEG) of HEMT is influenced by this surface adsorption-desorption process.

Download Full Size | PDF

Because of this sensitive dependence of channel conductance on the ZnO surface state, to control the defect states of NR surface will be one of the most crucial parameters affecting the performance of ZnO NRs gated HEMT-based PDs. Broadband (400-750 nm) visible PL emissions shown in Fig. 3 are generally associated with the intrinsic defect states such as, oxygen vacancies (VO), zinc vacancy (VZn), oxygen interstitial (Oi), zinc interstitial (Zni), oxygen anti-site (OZn), zinc anti-site (ZnO), and native defect clusters such as VOZni [22]. Experimental observation and its theoretical interpretation performed by Deng et al. [31] revealed that ZnO nanostructures have the highest concentration of VO at the surface, and these vacancies act as the trap centers for the n-type carriers in the conduction band (free electron). Since the atmospheric O2 molecules are highly electronegative, they can take the free electrons easily from conduction band near the surface of ZnO NRs thereby forming the negatively charged O2 ions [32]. Therefore, suppressing the defect density can give rise to the increases in free electron concentration and chemisorbed O2 ions at the surface by extending the depletion region even further. A. R. Jason [33] studied the roles of absorbed O2 ions and VO on the green emission of the ZnO NRs and suggested that the green emission is associated with the trap-assisted recombination of the electrons taken by the adsorbed O2 ions [14,33]. Extensive spectral analyses performed by earlier studies suggested that the green (∼510 nm) and yellow/orange emissions can arise from optical transitions in the single ionized VO (∼2.3 eV) and the single negatively charged Oi (∼2.0 eV) [34] inside the ZnO crystals. The formation of these defects are typically associated with either oxygen deficiency or excess oxygen in the ZnO crystals. On the other hand, unlike other emissions, yellow emissions are not significantly influenced by the surface state. It was also found that these various visible emissions from the ZnO crystals are strongly dependent on the surface reactivity and chemical composition at the surface. All these research efforts indicate the same direction that any slight modification in the ZnO surface state can have significant effect on non-radiative transitions at various deep level defects and luminescence emission properties.

 figure: Fig. 3.

Fig. 3. PL spectra of (a) CS and N2O plasma-treated ZnO NRs for (b) 1, (c) 3, and (d) 6 min with deconvoluted visible emission spectra. (e) Comparison of PL emissions from CS and N2O plasma-treated samples for 1, 3, and 6 min. (f) Energy band diagrams for the ZnO NR crystals in the cases of (left) CS and (right) plasma-treated samples.

Download Full Size | PDF

In our PL analysis, the most significant reduction was observed in green emission when the samples were exposed under N2O plasma as shown in Fig. 3. On the contrary, UV emission was significantly increased by the plasma treatment, which suggest that the defect concentration near the ZnO surface were reduced consequently.

Underlying effect of plasma treatment can be explained by the chemistry of N2O plasma. N2O molecules are very stable at RT; however, N-O bonds of N­2O in cold plasma, which are relatively weaker than N-N bonds, can be broken easily by producing various plasma radicals such as NO*, N*, and O* [35]. These plasma radicals are highly reactive, and they can have enough energy to initiate reactions with the active sites (VO, VZn, Oi) present in ZnO NRs at low temperature. The origin of these defects can be associated to the growth mechanism of ZnO NRs. Chemical reactions associated with the formation of ZnO crystallites are given by the followings.

$$Zn{(N{O_3})_2} \to Z{n^{2 + }} + 2N{O_3}^ -$$
$$N{H_4}OH \; \to \; N{H_3} + {H_2}O$$
$${({C{H_2}} )_6}{N_4} + 6{H_2}O \; \to \; 6HCH + 4N{H_3}$$
$$Z{n^{2 + }} + 4N{H_3} \to Zn{[{{{(N{H_3})}_4}} ]^{2 + }}$$
$$2{H_2}O\mathop \to \limits {H_3}{O^ + } + O{H^ - }$$
$$Z{n^{2 + }} + 2O{H^ - }\mathop \to \limits Zn{({OH} )_2}$$
$$Zn{({OH} )_2}\mathop \to \limits^{\; heat} ZnO + {H_2}O$$

In general, ZnO NRs grown by hydrothermal methods tend to have many carbon functional group complexes near the surface, such as hydrocarbons (COOH groups) and carbon oxides (COOR). These impurities have strong electron scavenging properties, therefore they can lead to serious degradation in the near band edge emission (NBE) and increase the trap-assisted emission in each respective wavelength range [22,33]. Since the N2O plasma radicals can provide enough enthalpy for the reaction to reduce the COOH and COOR groups, the following reaction can take place during the N2O plasma treatment process.

$$COO{H^ - } + N{O^\ast } \to {N_2}O + C{O_2} + {H_2}O$$
$$COO{H^ - } + {O^\ast }\mathop \to \limits CO + C{O_2} + {H_2}O$$
$$Zn{({OH} )_2}\mathop \to \limits^{\; {N_2}{O^\ast }} ZnO + {H_2}O$$

Furthermore, it can be assumed that many O* radicals in the plasma can reduce the concentration of VO defects. The activation energy of these defects near the surface is ∼2.3 eV [34] which can easily be overcome by the high-energy plasma radicals. As demonstrated in Figs. 3(a)–3(d), the suppression of visible emission defects is evident as revealed in PL spectra with the reduction of defects causing the radiative recombination with oxygen adsorbates after the plasma treatment. On the other hand, the NBE was significantly increased with the plasma treatment. This UV emission is associated with band-to-band excitonic recombination in ZnO, and its intensity is closely related to the crystalline quality. The reason behind this evolutionary change in this NBE is that shallow-level defects and non-radiative defects near band edge were passivated by the plasma treatment.

Shown in Fig. 4 are the 2θ scan XRD spectra measured from the samples prepared under different plasma treatment conditions. The most intense peak intensities were obtained from (002) reflections (2θ = 34.5°) in every samples, while the negligible or no reflection from (100), (101), and (102) planes were shown. This indicates that the ZnO nano-crystallites have a strong (002) preferred orientation aligned to the vertical direction to the substrate surface. One interesting observation is a prominent reflection from (100) β-phase Zn(OH)2 at 2θ = 32.8° according to JCPDS files (card No. 24-1444) as shown in logarithmic-scale plot of the CS. In solution-based growth process, a pivotal reaction of $Z{n^{2 + }} + \textrm{ }2O{H^ - } \leftrightarrow \textrm{ }Zn{({OH} )_2} \leftrightarrow ZnO(s )\textrm{ } + \textrm{ }{H_2}O$ can take place from the principal aqueous species of $Z{n^{2 + }}({aq} )$ during the formation of ZnO NRs [21]. Once theses hydroxides are formed, they can be either decomposed into the solution or consumed for the ZnO formation; but, a small amount of colloidal clusters Zn(OH)2 can remain as precipitates inside the crystals. When the samples were exposed to plasma treatment, the peak intensity associated with hydroxides started to decrease with the increase of (002) intensity of ZnO. This suggests that the residual Zn (OH)2 starts to dissociate into ZnO crystals in the presence of plasma radicals. The degree of preferred orientation F(hkl) of (hkl) plane can be defined by $F({hkl} )= [{P({hkl} )- {P_0}({hkl} )} ]/[{1 - {P_0}({hkl} )} ]$, where $P({hkl} )= I({hkl} )/\Sigma I({hkl} )$, ${P_0}({hkl} )= {I_0}({hkl} )/\Sigma {I_0}({hkl} )$, $I({hkl} )$ is the intensity of peak measured from (hkl) plane, and ${I_0}({hkl} )$ is the reference peak intensity of (hkl) plane given by JCPDS card number 36-1451 [3639]. As shown in Table 1, the highest F(002) of ∼0.996 was obtained from the samples plasma-treated for 6 min. This remarkable improvement in F(002) suggests that our plasma treatment promoted the most stable growth mode of ZnO crystal along (002) direction by effectively suppressing the impurities and intrinsic defects present in significant amount caused by the solution-based synthesis process.

 figure: Fig. 4.

Fig. 4. XRD 2-theta scan patterns in (a) logarithmic scale and (b) linear scale in intensity (theta = θ).

Download Full Size | PDF

Tables Icon

Table 1. Parameters extracted from XRD analysis (PT: plasma treatment).

Shown in Fig. 5(a) is the XPS spectra in a wide binding energy (BE) range of 0-1200 eV measured from each sample. In these spectra, the photoemissions from each core level, spin orbital splittings, and Auger-electron peaks from key elements of Zn and O were observed, while traces of carbon were also found in each sample due to the carbon rich solution-based growth of ZnO NRs. Expanded views of the emission spectra in a BE range of 280-295 eV were shown in Fig. 5(b). Two peaks from C1s were shown at ∼284.5 (major) and ∼288.5 eV (minor), where each peak corresponds to the BE of C-C bonds and oxygen containing functional groups such as COC epoxy, COOH hydroxyl, and COOR groups [40]. A clear trend in decreasing peak intensity with the increase of plasma time was shown from C-1s spectra; therefore, it can be assumed that carbon-related impurities in various forms present near the surface were effectively suppressed by the plasma treatment. As discussed earlier with the chemical reactions in Eqs. (8) and (9), the reduction in carbon-related impurities can be due to surface reaction of N2O plasma radicals with these impurities. As shown in Table 2, the atomic percentage of C was initially ∼5% in the CS, however it was greatly reduced to ∼2% after the N2O plasma treatment for 6 min. Figure 5(c) presents the BE spectra of the Zn core electrons from the CS as well as plasma-treated samples. As observed in the spectra, spin-orbital splittings of Zn2p1/2 (∼1044.28 eV) and Zn2p3/2 (∼1021.18 eV) were shown with a doublet splitting of 23.1 eV, which represent the different chemical state of Zn as Zn2+. More careful observation on the positions of each peak listed in Table 2 reveals that both peaks were shifted toward higher BE by ∼0.2 eV after the plasma treatment for 6 min. These chemical shifts in BE indicate that the interaction of Zn interstitials and stoichiometric Zn with O* radicals produced by N2O plasma can take place by producing the higher charged states of Zn atoms as reported in the previous studies [20].

 figure: Fig. 5.

Fig. 5. XPS spectra of each sample in BE ranges of (a) 0-1200, (b) 280-295, (c) 1020-1050, (d) 534-527 eV (inset shows a magnified view in a BE range of 533.5-529.5 eV). Deconvoluted O1s peaks into three satellites of Oa, Ob, and Oc extracted from (e) CS and N2O plasma-treated samples for (f) 1, (g) 3, and (h) 6 min. (i) Ratios of ∫Ob/∫Oa and ∫Oc/∫Oa in deconvoluted satellite spectra of O1s versus plasma treatment time.

Download Full Size | PDF

Tables Icon

Table 2. Binding energies (BE) for Zn2p doublet, atomic percentage contributions of the key elements, and Oa/Zn ratio in ZnO NRs (PT: plasma treatment).

Figures 5(d)–5(h) illustrate the core levels of O1s in an expanded spectra including three Gaussian deconvoluted peaks of Oa, Ob, and Oc. The main peak Oa (529.88 eV) originates from the binding state of O2+ with metal ions such as Zn2+ in the ZnO crystals [41]. Because the amount of Oa emission is a good measure of stoichiometric oxygen present in the crystal, we estimated the O stoichiometry by calculating the atomic percentage contribution of Oa in O­1s (%Oa = %O1s× [∫Oa/(∫Oa + ∫Ob + ∫Oc)]) from each sample as summarized in Table 2, where %O1s is the O1s curve integration divided by total curve integration, and ∫Oi (i = a, b, c) is the curve integration of each satellite peak. It was shown that the stoichiometric oxygen percentage was increased to 40.1% in the case of N2O plasma-treated sample for 6 min compared to 35.7% in the CS. At the same time, the atomic percentage of Zn was also increased from 40.7% of the CS to 41.8% with much improved stoichiometric ratio of Oa to Zn (∼0.96) in the case of the sample plasma-treated for 6 min. This improvement can be attributed to the reduction of non-stoichiometric oxygen including the VO defects in plasma treated samples caused by the interaction with plasma radicals.

Table 2 also summarized the percentage contribution from each satellite peak of Ob and Oc. Ob is centered at ∼531 eV and known to arise from O2 ions in the oxygen-deficient regions (where oxygen vacancies are present) of the ZnO matrix [42]. %Ob also exhibited a significant reduction from ∼9.5% of the CS to 4.0% in the case of the sample plasma-treated for 6 min. The third component of O1s peak at higher BE (∼532 eV) expressed by Oc was reported to be associated with either chemisorbed oxygen entity or the species of OH during the fabrication process in organic solutions [43]. This phenomenon was not fully understood, however it is thought that the increase of %Oc under plasma exposure for 1-3 min increased the chemisorbed oxygen on the ZnO surface under an environment ambient in O-radicals, which can eventually give rise to more band bending and increased surface potential at the surface of ZnO. However, further plasma exposure can activate the surface and suppress the hydroxyl groups present near the surface with the reaction with radicals. A quantitative analysis is also shown in Fig. 5(i) to present the change of our deconvoluted O1s peaks with respect to plasma treatment time. As shown in the plot, the area ratio ∫Ob/∫Oa was significantly decreased with the increase of plasma treatment time from ∼0.28 (in the case of CS) to ∼0.12 (in the sample plasma-treated for 6 min). This trend showed a good agreement with our PL observation, where the green emission was reduced with the plasma treatment, while the band edge emission was enhanced on the contrary. The suppression of green emission can be most likely caused by the reduction of VO defects on the NR surface [34] due to the irradiation of high energy O* radicals in the plasma. The area ratio of ∫Oc/∫Oa showed an opposite trend with plasma exposure time, which can be due to the enhanced adsorption of oxygen related species as discussed earlier.

Shown in Fig. 6 are the measured photoresponse characteristics of the HEMT-based PDs fabricated with ZnO NRs prepared under different treatment conditions. Figure 6(a) presents drain current (Ids) versus drain voltage (Vds) in the dark (Ids_DARK) and UV exposure (Ids_PHOTO) conditions. Measured values of the Ids_DARK, Ids_PHOTO, and on/off ratio (Ids_PHOTO/Ids_DARK) at a Vds of 5 V were summarized in Table 3. As discussed in the sensing mechanism, the Ids depends upon the negative surface potential built in the gate area by the adsorption of O2 ions at the ZnO surface. As discussed in our material characterizations, evolutionary change in surface state and crystalline quality of ZnO NRs was made after plasma treatment with further promoted adsorption of O2 ions at the surface in the dark condition. For this reason, measured Ids_DARK showed the lowest value of ∼0.59 mA/mm after plasma treatment for 6 min, whereas much higher Ids_DARK of ∼1.91 mA/mm was measured from the CS. The correlation of Ids­_DARK with plasma exposure time can be attributed to the improved surface crystalline state of ZnO NRs, which promotes the adsorption of O2 ions with the increase in negative surface potential on the NRs. The negative surface potential pushes 2-DEG electrons (present at the interface of GaN/AlGaN) deep into the GaN channel, thereby reducing the concentration of 2-DEG electrons in the channel and resulting in lower Ids_DARK after the plasma treatment. Similarly, Ids_PHOTO was also increased by almost 4 times from the PDs with NRs plasma-treated for 6 min compared to that in the case of untreated NRs, and the improvement is closely related to the improved crystalline quality near the surface of ZnO NRs as revealed in our PL analysis. This supports our assumption that the plasma-treated NRs will have smaller numbers of surface traps contributing to the negative potential barrier in the gate region near the surface [14]. However, as shown in Fig. 6, the effect of plasma treatment for 1 min did not make clear effect on both Ids_DARK and Ids_PHOTO values due to insufficient plasma exposure for the change in NR crystal surface quality. In consequence, lower trap density leads to more reduced negative surface potential with plasma treatment even after the desorption of O2 ions in the presence of UV light and gives rise to higher Ids_PHOTO as shown in Table  3. On/off ratios measured from the PDs at a Vds of 5 V were summarized in Table  3. The PDs with NRs plasma-treated for 6 min showed the highest on/off ratio of ∼39, while the PD with untreated NRs exhibited the lowest value of ∼5.

 figure: Fig. 6.

Fig. 6. (a) Measured Ids versus Vds, (b) Ids_PHOTO versus Pi measured at Vds of 5 V, (c) spectral responsivity, (d) specific detectivity as functions of radiant light wavelength measured from the PDs, (e) time resolved photoresponse characteristics at Vds = 8 V, and (f) magnified views with τr (rise time) and τf (fall time) of transient characteristics upon UV on and off.

Download Full Size | PDF

Tables Icon

Table 3. Summary of Ids_PHOTO, Ids_DARK, on/off ratio and gm,photo measured at Vds of 5 V from HEMT-based PDs fabricated with untreated and plasma-treated ZnO NRs (1, 3, 6 min) (PT: plasma treatment).

Spectral responsivity R and specific detectivity D* are the vital figures of merit for the PDs as defined in the following equations [1,8,9]

$$\boldsymbol{R} = \frac{{{\boldsymbol{I}_{\boldsymbol{ds}\_\boldsymbol{PHOTO}}} - {\boldsymbol{I}_{\boldsymbol{ds}\_\boldsymbol{DARK}}}}}{{{\boldsymbol{P}_{\boldsymbol{i}}}}}$$
$${\boldsymbol{D}^{\ast}} = \frac{{\sqrt{\boldsymbol{A}} \boldsymbol{R}}}{{\sqrt {2\boldsymbol{q}{\boldsymbol{I}_{\boldsymbol{ds}\_\boldsymbol{DARK}}}}}}$$
, where A is the active area of device, q is the electron charge, and Pi is the radiant light power incident on the active area. Responsivities of the HEMT-based PDs were measured using a monochromatic light source with an intensity of 16 µW/cm2 and an effective area of 2.4×2.4 mm2. As shown in Fig. 6(c), The PDs with NRs plasma-treated for 6 min showed the highest R of ∼1.5×105 A/W at a wavelength of 350 nm and a Vds of 1 V. Plasma treatment for 6 min brought about an increase of R by ∼3 times, and this improvement can be explained by a higher photonic transconductance (gm,photo) characteristic of our HEMT-based PDs [8]. As expressed in Eq. (13), the gm,photo represents the ratio of change in Ids_PHOTO given by the change in Pi.
$${\boldsymbol{g}_{\boldsymbol{m},\boldsymbol{photo}}} = \frac{{\mathrm{\Delta}{\boldsymbol{I}_{\boldsymbol{ds}\_\boldsymbol{PHOTO}}}}}{{\mathrm{\Delta}{\boldsymbol{P}_{\boldsymbol{i}}}}}\;\left( {\frac{\boldsymbol{A}}{\boldsymbol{W}}} \right)$$
This gm,photo parameter can be used to evaluate the gain characteristics of our gateless HEMTs because the change in incident optical power is directly linked to the virtual gate voltage induced by the change in numbers of adsorbed O2 on the ZnO surface upon UV irradiation. These parameters were extracted from the slopes in the measured Ids_PHOTO­ versus Pi shown in Fig. 6(b) and summarized in Table  3. The PDs with NRs plasma-treated for 6 min showed the highest gm,photo of ∼7.54 A/W. This enhancement in R indicates that, when the NRs are plasma treated, a better gate controllability is achieved in the gateless HEMT devices due to improved surface crystalline quality of ZnO promoting the adsorption and desorption of O2 ions. D* for a PD is used to characterize performance equal to the reciprocal of noise-equivalent power normalized per square root of the detector area and frequency bandwidth. The maximum D* of ∼4.7×1013 Jones (cmHz−1/2W−1) was obtained at 370 nm from the PDs with NRs plasma-treated for 6 min. This detectivity is comparable with that (∼8.5×1013 Jones) of the PDs comprising ZnO thin films deposited by atomic layer deposition with metal-semiconductor-metal (MSM) structure [44]; however, the reported responsivity of this MSM PDs was much lower (∼27 A/W) than that of out HEMT-based PDs.

The real time transient characteristics of the PDs were measured at a Vds of 8 V by using the UV light source of a peak wavelength at 350 nm and a programmable light shutter, as shown in Figs. 6(e) and 6(f). The photocurrents measured from every device examined in this study showed very fast rise-ups upon UV exposure and relatively slower fall-offs with UV turn-off. The photocurrent oscillations showed a fairly good repeatability over 100 cycles from each device. We obtained the rise time (the time for photocurrent to rise up to 90% of maximum saturation value under UV turn-on) of ∼0.7 s and the fall time (the time for photocurrent to fall off by 90% from maximum value under UV turn-off) of ∼3.3 s from the reference device (no plasma-treated), as shown in Fig. 6(f). After 6 min plasma treatment, the rise time was slightly reduced to 0.5 s, however the fall time was increased to 6.2 s. These measured response speeds of our HEMT-based PDs were far faster than those (few tens of seconds) of the ZnO NR-based passive detectors reported to date [45,46].

4. Conclusions

Visible-blind UV PDs were fabricated using the gateless AlGaN/GaN HEMTs with ZnO NRs of which surface structures were modified by N2O plasma treatment. The best photoresponse characteristics in terms of photocurrent (22.6 mA/mm), dark current (0.59 mA/mm), on/off current ratio (∼38.6), responsivity (∼1.5×105 A/W), and detectivity (D*, ∼ 4.7×1013 Jones) were achieved when the NRs were plasma-treated for 6 min. The PL analysis showed the highest intensity in near edge emission and the greatest suppression in defect-related visible emissions upon plasma treatment. The stoichiometry in ZnO crystals also showed a significant improvement after plasma treatment as an oxygen to zinc atomic ratio of ∼0.96 was measured by XPS analysis. The highest (002) degree of orientation (∼0.996) were demonstrated by the XRD investigation from the NRs plasma-treated for 6 min.

Funding

National Research Foundation of Korea (2019R1A2C2086747, K200304003).

Acknowledgments

This work was supported by the Mid-career Researcher Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (No. 2019R1A2C2086747). Authors would like to thank Mr. Jihyun Kim in University of Michigan for his sincere analysis of the experimental data. Material characterization was supported by Nano-material Technology Development Program through the NRF funded by the Ministry of Science, ICT and Future Planning (No. K200304003).

Disclosures

The authors declare no conflicts of interest.

References

1. Z. Bai, X. Yan, X. Chen, K. Zhao, P. Lin, and Y. Zhang, “High sensitivity, fast speed and self-powered ultraviolet photodetectors based on ZnO micro/nanowire networks,” Prog. Nat. Sci. 24(1), 1–5 (2014). [CrossRef]  

2. K. Liu, M. Sakurai, and M. Aono, “ZnO-Based Ultraviolet Photodetectors,” Sensors 10(9), 8604–8634 (2010). [CrossRef]  

3. D. Walker, X. Zhang, P. Kung, A. Saxler, S. Javadpour, J. Xu, and M. Razeghi, “AlGaN ultraviolet photoconductors grown on sapphire,” Appl. Phys. Lett. 68(15), 2100–2101 (1996). [CrossRef]  

4. A. Hirano, C. Pernot, M. Iwaya, T. Detchprohm, H. Amano, and I. Akasaki, “Demonstration of Flame Detection in Room Light Background by Solar-Blind AlGaN PIN Photodiode,” phys. stat. sol. (a) 188(1), 293–296 (2001). [CrossRef]  

5. I. J. Bigio and J. R. Mourant, “Ultraviolet and visible spectroscopies for tissue diagnostics: fluorescence spectroscopy and elastic-scattering spectroscopy,” Phys. Med. Biol. 42(5), 803–814 (1997). [CrossRef]  

6. Y.-F. Wu, D. Kapolnek, J. P. Ibbetson, P. Parikh, B. P. Keller, and U. K. Mishra, “Very-high power density AlGaN/GaN HEMTs,” IEEE Trans. Electron Devices 48(3), 586–590 (2001). [CrossRef]  

7. Z. H. Zaidi and P. A. Houston, “Highly Sensitive UV Detection Mechanism in AlGaN/GaN HEMTs,” IEEE Trans. Electron Devices 60(9), 2776–2781 (2013). [CrossRef]  

8. F. Khan, W. Khan, and S.-D. Kim, “High-Performance Ultraviolet Light Detection Using Nano-Scale-Fin Isolation AlGaN/GaN Heterostructures with ZnO Nanorods,” Nanomaterials 9(3), 440 (2019). [CrossRef]  

9. F. Khan, W. Khan, J. H. Kim, N. ul Huda, H. M. Salman Ajmal, and S.-D. Kim, “Oxygen desorption kinetics of ZnO nanorod–gated AlGaN/GaN HEMT-based UV photodetectors,” AIP Adv. 8(7), 075225 (2018). [CrossRef]  

10. D. A. Melnick, “Zinc Oxide Photoconduction, an Oxygen Adsorption Process,” J. Chem. Phys. 26(5), 1136–1146 (1957). [CrossRef]  

11. Y. Liu, X. Zhang, J. Su, H. Li, Q. Zhang, and Y. Gao, “Ag nanoparticles@ZnO nanowire composite arrays: an absorption enhanced UV photodetector,” Opt. Express 22(24), 30148 (2014). [CrossRef]  

12. X. Zhang, X. Han, J. Su, Q. Zhang, and Y. Gao, “Well vertically aligned ZnO nanowire arrays with an ultra-fast recovery time for UV photodetector,” Appl. Phys. A 107(2), 255–260 (2012). [CrossRef]  

13. C. Aharoni and F. C. Tompkins, “Kinetics of Adsorption and Desorption and the Elovich Equation,” in Advances in Catalysis, D. D. Eley, H. Pines, and P. B. Weisz, eds., 21st ed. (Elsevier, 1970), pp. 1–49.

14. J. Moore and C. Thompson, “A Phenomenological Model for the Photocurrent Transient Relaxation Observed in ZnO-Based Photodetector Devices,” Sensors 13(8), 9921–9940 (2013). [CrossRef]  

15. Q. H. Li, T. Gao, Y. G. Wang, and T. H. Wang, “Adsorption and desorption of oxygen probed from ZnO nanowire films by photocurrent measurements,” Appl. Phys. Lett. 86(12), 123117 (2005). [CrossRef]  

16. N. H. Al-Hardan, M. A. Abdul Hamid, N. M. Ahmed, A. Jalar, R. Shamsudin, N. K. Othman, L. Kar Keng, and S. M. Mohammed, “A Study on the UV Photoresponse of Hydrothermally Grown Zinc Oxide Nanorods With Different Aspect Ratios,” IEEE Sens. J. 15(12), 6811–6818 (2015). [CrossRef]  

17. S. Dogar, W. Khan, F. Khan, and S.-D. Kim, “Effect of NH3 plasma treatment on the transient characteristics of ZnO nanorod-gated AlGaN/GaN high electron mobility transistor-based UV sensors,” Thin Solid Films 642, 69–75 (2017). [CrossRef]  

18. W. Khan, H. Ajmal, F. Khan, N. Huda, and S.-D. Kim, “Induced Photonic Response of ZnO Nanorods Grown on Oxygen Plasma-Treated Seed Crystallites,” Nanomaterials 8(6), 371 (2018). [CrossRef]  

19. C. Lin, H.-P. Chen, H. Liao, and S. Chen, “Enhanced luminescent and electrical properties of hydrogen-plasma ZnO nanorods grown on wafer-scale flexible substrates,” Appl. Phys. Lett. 86(18), 183103 (2005). [CrossRef]  

20. W. Khan, F. Khan, H. Ajmal, N. Huda, J. Kim, and S.-D. Kim, “Evolution of Structural and Optical Properties of ZnO Nanorods Grown on Vacuum Annealed Seed Crystallites,” Nanomaterials 8(2), 68 (2018). [CrossRef]  

21. Y. Sun, D. J. Riley, and M. N. R. Ashfold, “Mechanism of ZnO Nanotube Growth by Hydrothermal Methods on ZnO Film-Coated Si Substrates,” J. Phys. Chem. B 110(31), 15186–15192 (2006). [CrossRef]  

22. M. Willander, O. Nur, J. R. Sadaf, M. I. Qadir, S. Zaman, A. Zainelabdin, N. Bano, and I. Hussain, “Luminescence from Zinc Oxide Nanostructures and Polymers and their Hybrid Devices,” Materials 3(4), 2643–2667 (2010). [CrossRef]  

23. H. Faber, J. Hirschmann, M. Klaumünzer, B. Braunschweig, W. Peukert, and M. Halik, “Impact of Oxygen Plasma Treatment on the Device Performance of Zinc Oxide Nanoparticle-Based Thin-Film Transistors,” ACS Appl. Mater. Interfaces 4(3), 1693–1696 (2012). [CrossRef]  

24. A. Talukder, “Plasma Treatment of Zinc Oxide Thin Film and Temperature Sensing Using the Zinc Oxide Thin Film,” South Dakota State University (2016).

25. N. K. Reddy, M. Devika, and Y.-B. Hahn, “Excellent enhancement in the device performance of nitrogen plasma treated ZnO nanorods based diodes,” Nano Convergence 1(1), 26 (2014). [CrossRef]  

26. S.-J. Cho, S. Doğan, S. Sabuktagin, M. A. Reshchikov, D. K. Johnstone, and H. Morkoç, “Surface band bending in as-grown and plasma-treated n-type GaN films using surface potential electric force microscopy,” Appl. Phys. Lett. 84(16), 3070–3072 (2004). [CrossRef]  

27. J. Y. Lee, C. M. Shin, J. H. Heo, C. R. Kim, J. H. Park, T. M. Lee, H. Ryu, C. S. Son, B. C. Shin, and W. J. Lee, “Effects of O2 plasma pre-treatment on ZnO thin films grown on polyethersulfone substrates at various deposition temperatures by atomic layer deposition,” Curr. Appl. Phys. 10(2), S290–S293 (2010). [CrossRef]  

28. S. Jiang, Z. Ren, S. Gong, S. Yin, Y. Yu, X. Li, G. Xu, G. Shen, and G. Han, “Tunable photoluminescence properties of well-aligned ZnO nanorod array by oxygen plasma post-treatment,” Appl. Surf. Sci. 289, 252–256 (2014). [CrossRef]  

29. K. Remashan, D. K. Hwang, S. D. Park, J. W. Bae, G. Y. Yeom, S. J. Park, and J. H. Jang, “Effect of N2O Plasma Treatment on the Performance of ZnO TFTs,” Electrochem. Solid-State Lett. 11(3), H55 (2008). [CrossRef]  

30. Y.-C. Yoon, K.-S. Park, and S.-D. Kim, “Effects of low preheating temperature for ZnO seed layer deposited by sol–gel spin coating on the structural properties of hydrothermal ZnO nanorods,” Thin Solid Films 597, 125–130 (2015). [CrossRef]  

31. B. Deng, A. Luisa da Rosa, T. Frauenheim, J. P. Xiao, X. Q. Shi, R. Q. Zhang, and M. A. Van Hove, “Oxygen vacancy diffusion in bare ZnO nanowires,” Nanoscale 6(20), 11882–11886 (2014). [CrossRef]  

32. X. Wang, B. Lu, L. Li, and H. Qiu, “Exploring the Interactions of Oxygen with Defective ZnO,” ChemistryOpen 7(7), 491–494 (2018). [CrossRef]  

33. J. A. Röhr, J. Sá, and S. J. Konezny, “The role of adsorbates in the green emission and conductivity of zinc oxide,” Commun. Chem. 2(1), 52 (2019). [CrossRef]  

34. A. Janotti and C. G. Van de Walle, “Oxygen vacancies in ZnO,” Appl. Phys. Lett. 87(12), 122102 (2005). [CrossRef]  

35. X. Fan, S. Kang, J. Li, and T. Zhu, “Conversion of dilute nitrous oxide (N 2 O) in N 2 and N 2 –O 2 mixtures by plasma and plasma-catalytic processes,” RSC Adv. 8(47), 26998–27007 (2018). [CrossRef]  

36. R. A. Talib, M. J. Abdullah, H. S. Al-Salman, S. M. Mohammad, N. M. Ahmed, and M. Bououdina, “Effect of growth time on structure, optical and photoresponse characteristics of ZnO nanorods deposited onto various substrates,” J. Ovonic Res. 12, 171–184 (2016).

37. R. Shabannia, “Synthesis and characterization of Cu-doped ZnO nanorods chemically grown on flexible substrate,” J. Mol. Struct. 1118, 157–160 (2016). [CrossRef]  

38. A. Peić, T. Dimopoulos, R. Resel, S. Abermann, M. Postl, E. J. W. List, and H. Brückl, “Effect of AZO Substrates on Self-Seeded Electrochemical Growth of Vertically Aligned ZnO Nanorod Arrays and Their Optical Properties,” J. Nanomater. 2012, 1–14 (2012). [CrossRef]  

39. R. A. Talib, M. J. Abdullah, H. S. Al-Salman, S. M. Mohammad, and N. K. Allam, “ZnO nanorods/polyaniline heterojunctions for low-power flexible light sensors,” Mater. Chem. Phys. 181, 7–11 (2016). [CrossRef]  

40. R. Blume, D. Rosenthal, J.-P. Tessonnier, H. Li, A. Knop-Gericke, and R. Schlögl, “Characterizing Graphitic Carbon with X-ray Photoelectron Spectroscopy: A Step-by-Step Approach,” ChemCatChem 7(18), 2871–2881 (2015). [CrossRef]  

41. P. S. Shewale and Y. S. Yu, “UV photodetection properties of pulsed laser deposited Cu-doped ZnO thin film,” Ceram. Int. 43(5), 4175–4182 (2017). [CrossRef]  

42. J. Lv, Q. Zhu, Z. Zeng, M. Zhang, J. Yang, M. Zhao, W. Wang, Y. Cheng, G. He, and Z. Sun, “Enhanced photocurrent and photocatalytic properties of porous ZnO thin film by Ag nanoparticles,” J. Phys. Chem. Solids 111, 104–109 (2017). [CrossRef]  

43. N. J. Nicholas, G. V. Franks, and W. A. Ducker, “The mechanism for hydrothermal growth of zinc oxide,” CrystEngComm 14(4), 1232–1240 (2012). [CrossRef]  

44. A. B. Djurišić, Y. H. Leung, K. H. Tam, L. Ding, W. K. Ge, H. Y. Chen, and S. Gwo, “Green, yellow, and orange defect emission from ZnO nanostructures: Influence of excitation wavelength,” Appl. Phys. Lett. 88(10), 103107 (2006). [CrossRef]  

45. M. R. Alenezi, S. J. Henley, and S. R. P. Silva, “On-chip Fabrication of High Performance Nanostructured ZnO UV Detectors,” Sci. Rep. 5(1), 8516 (2015). [CrossRef]  

46. Q. Humayun, M. Kashif, U. Hashim, and A. Qurashi, “Selective growth of ZnO nanorods on microgap electrodes and their applications in UV sensors,” Nanoscale Res. Lett. 9(1), 29 (2014). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. (a) Schematic illustrations (untreated and plasma-treated), (b) top-view optical image, and (c) top-view SEM micrograph of the fabricated ZnO NR-gated AlGaN/GaN HEMT. (d) Top-view (left), expanded top-view (top-right), and cross-sectional SEM view (bottom right) of as-grown ZnO NRs.
Fig. 2.
Fig. 2. Illustration of O2 (a) adsorption in the dark condition (b) and desorption under UV illumination taking place on the ZnO NR surface. Free electron concentration in two-dimensional electron gas (2-DEG) of HEMT is influenced by this surface adsorption-desorption process.
Fig. 3.
Fig. 3. PL spectra of (a) CS and N2O plasma-treated ZnO NRs for (b) 1, (c) 3, and (d) 6 min with deconvoluted visible emission spectra. (e) Comparison of PL emissions from CS and N2O plasma-treated samples for 1, 3, and 6 min. (f) Energy band diagrams for the ZnO NR crystals in the cases of (left) CS and (right) plasma-treated samples.
Fig. 4.
Fig. 4. XRD 2-theta scan patterns in (a) logarithmic scale and (b) linear scale in intensity (theta = θ).
Fig. 5.
Fig. 5. XPS spectra of each sample in BE ranges of (a) 0-1200, (b) 280-295, (c) 1020-1050, (d) 534-527 eV (inset shows a magnified view in a BE range of 533.5-529.5 eV). Deconvoluted O1s peaks into three satellites of Oa, Ob, and Oc extracted from (e) CS and N2O plasma-treated samples for (f) 1, (g) 3, and (h) 6 min. (i) Ratios of ∫Ob/∫Oa and ∫Oc/∫Oa in deconvoluted satellite spectra of O1s versus plasma treatment time.
Fig. 6.
Fig. 6. (a) Measured Ids versus Vds, (b) Ids_PHOTO versus Pi measured at Vds of 5 V, (c) spectral responsivity, (d) specific detectivity as functions of radiant light wavelength measured from the PDs, (e) time resolved photoresponse characteristics at Vds = 8 V, and (f) magnified views with τr (rise time) and τf (fall time) of transient characteristics upon UV on and off.

Tables (3)

Tables Icon

Table 1. Parameters extracted from XRD analysis (PT: plasma treatment).

Tables Icon

Table 2. Binding energies (BE) for Zn2p doublet, atomic percentage contributions of the key elements, and Oa/Zn ratio in ZnO NRs (PT: plasma treatment).

Tables Icon

Table 3. Summary of Ids_PHOTO, Ids_DARK, on/off ratio and gm,photo measured at Vds of 5 V from HEMT-based PDs fabricated with untreated and plasma-treated ZnO NRs (1, 3, 6 min) (PT: plasma treatment).

Equations (13)

Equations on this page are rendered with MathJax. Learn more.

Z n ( N O 3 ) 2 Z n 2 + + 2 N O 3
N H 4 O H N H 3 + H 2 O
( C H 2 ) 6 N 4 + 6 H 2 O 6 H C H + 4 N H 3
Z n 2 + + 4 N H 3 Z n [ ( N H 3 ) 4 ] 2 +
2 H 2 O H 3 O + + O H
Z n 2 + + 2 O H Z n ( O H ) 2
Z n ( O H ) 2 h e a t Z n O + H 2 O
C O O H + N O N 2 O + C O 2 + H 2 O
C O O H + O C O + C O 2 + H 2 O
Z n ( O H ) 2 N 2 O Z n O + H 2 O
R = I d s _ P H O T O I d s _ D A R K P i
D = A R 2 q I d s _ D A R K
g m , p h o t o = Δ I d s _ P H O T O Δ P i ( A W )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.