Next Article in Journal
Electromagnetically Induced Transparency-Like Effect by Dark-Dark Mode Coupling
Next Article in Special Issue
Electronic and Optical Properties of Atomic-Scale Heterostructure Based on MXene and MN (M = Al, Ga): A DFT Investigation
Previous Article in Journal
Modified Camellia oleifera Shell Carbon with Enhanced Performance for the Adsorption of Cooking Fumes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photosensing and Characterizing of the Pristine and In-, Sn-Doped Bi2Se3 Nanoplatelets Fabricated by Thermal V–S Process

1
Department of Materials Science and Engineering, National Chung Hsing University, Taichung 40227, Taiwan
2
Department of Chemical Engineering and Materials Science, Chinese Culture University, Taipei 11114, Taiwan
*
Authors to whom correspondence should be addressed.
Submission received: 6 April 2021 / Revised: 17 May 2021 / Accepted: 18 May 2021 / Published: 20 May 2021
(This article belongs to the Special Issue Optoelectronic Properties and Applications of Nanomaterials)

Abstract

:
Pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets synthesized on Al2O3(100) substrate by a vapor–solid mechanism in thermal CVD process via at 600 °C under 2 × 10−2 Torr. XRD and HRTEM reveal that In or Sn dopants had no effect on the crystal structure of the synthesized rhombohedral-Bi2Se3. FPA–FTIR reveals that the optical bandgap of doped Bi2Se3 was 26.3%, 34.1%, and 43.7% lower than pristine Bi2Se3. XRD, FESEM–EDS, Raman spectroscopy, and XPS confirm defects ( I n 3 + B i 3 + ), ( I n 3 + V 0 ), ( S n 4 + B i 3 + ), ( V 0 B i 3 + ), and ( S n 2 + B i 3 + ). Photocurrent that was generated in (In,Sn)-doped Bi2Se3 under UV(8 W) and red (5 W) light revealed stable photocurrents of 5.20 × 10−10 and 0.35 × 10−10 A and high Iphoto/Idark ratios of 30.7 and 52.2. The rise and fall times of the photocurrent under UV light were 4.1 × 10−2 and 6.6 × 10−2 s. Under UV light, (In,Sn)-dopedBi2Se3 had 15.3% longer photocurrent decay time and 22.6% shorter rise time than pristine Bi2Se3, indicating that (In,Sn)-doped Bi2Se3 exhibited good surface conduction and greater photosensitivity. These results suggest that In, Sn, or both dopants enhance photodetection of pristine Bi2Se3 under UV and red light. The findings also suggest that type of defect is a more important factor than optical bandgap in determining photo-detection sensitivity. (In,Sn)-doped Bi2Se3 has greater potential than undoped Bi2Se3 for use in UV and red-light photodetectors.

1. Introduction

Bi2Se3 is a well-known second-generation topological insulator (TI) with a narrow bandgap of 0.35 eV and a rhombohedral crystal structure [1]. Se vacancies ( v Se ), which act as electron donors, are the main defects in the Bi2Se3 structure, making it an n-type topological insulator [2]. The crystalline Bi2Se3 is composed of layered structures; each layer consists of five stacked monoatomic layers, as in Se-Bi-Se’-Bi-Se, and is thus known as a quintuple layer (QL) [3]. Covalent bonds dominate the QL [4], whereas Van der Waals’ forces dominate between QLs; hence, the dopants can be adequately intercalated among them [5]. A TI has an insulating bulk state and a topologically protected gapless surface state in three dimensions and an edge state in two dimensions, owing to spin-orbital coupling (SOC) and time-reversal symmetry (TRS) [6,7]. Both SOC and TRS suppress backscattering and reduce the sensitivity to surface impurities or defects when electrons are transported on the surface of a TI. These gapless states thus lead to a high electronic conductivity [8,9] and the following features: (1) photon-like electrons, (2) low power dissipation, (3) spin-polarized electrons, and (4) the quantum spin Hall effect [10,11,12,13]. Owing to TIs’ unique electronic properties, they have many potential applications, including photodetectors [14], lasers [15], gas sensors [16], spintronic devices [17], magnetoelectronic devices [18], quantum computers [19], and topological superconductors [20]. Several methods are commonly used to synthesize TIs; they include chemical vapor deposition [21], mechanical exfoliation [22], solvothermal synthesis [23], molecular beam epitaxy [24], atomic layer epitaxy [25], metal–organic chemical vapor deposition [26], pulsed laser deposition [27], magnetron sputtering [28], and the Bridgeman method [29].
Broadband photodetectors are made of potential materials, whose conductivity can be changed by incident light such as UV, visible, or IR light; they have photoabsorbance over a wide range of wavelengths, high photosensitivity, high carrier mobility, high conservation efficiency, operability at a low voltage, and long operational stability [30,31]. Photon–electron transfer is the primary detection mechanism of photodetectors [32]. Thus, photodetectors can be used as potential materials in optical information communication, imaging detection, and biodetectors [33,34]. As required for use in photodetectors, topological insulating Bi2Se3 is a potential material and has fascinating optoelectronic properties, such as a tunable surface bandgap, a polarization-sensitive photocurrent, and thickness-dependent optical absorption [35]. Bulk Bi2Se3 has a narrow bandgap of 0.35 eV and therefore a wide range of absorption wavelengths. However, its surface transportation can be suppressed by the free carriers that are presented in its bulk state. Shrinking to the nanoscale, as for nanoplates, nanowires, nanoribbons, and thin films, it suppresses the contribution of the bulk state to reduce the surface transportation, strengthening its surface transportation by increasing the surface-to-volume ratio [7,9]. Supplementary Table S1 lists Bi2Se3-based photodetectors.
The optical bandgap of Bi2Se3 depends on the synthesizing conditions. Pejova et al. reported that chemically deposited Bi2Se3 thin films have an optical bandgap energy of 2.3 eV [36]; Pramanik et al. fabricated Bi2Se3 with optical bandgaps of 1.03 and 1.15 eV by chemical deposition [37]. Garcia et al. reported that chemically prepared Bi2Se3 films had optical bandgaps of 1.41–1.7 eV and that postannealing (200 °C) Bi2Se3 films had bandgaps of 1.06–1.57 eV [38]. Manjulavalli et al. found that Bi2Se3 films that were prepared by thermal evaporation had an optical bandgap of 0.825 eV, and annealed films had a gap of 0.61 eV [39]. Augustine et al. reported that Bi2Se3 films that were fabricated by thermal evaporation had an optical bandgap of 0.67 eV [40]. Alemi et al. synthesized Bi2Se3 nanoplatelets, which had a bandgap of 2.95 eV, by a hydrothermal process [41]. These results reveal that the absorption range of Bi2Se3 can be narrowed, affecting its photo-detective efficiency. According to the relevant literature, doping can modify the optical bandgaps of Bi2Se3 nanostructures. Supplementary Table S2 presents the variations of bandgap energy of Bi2Se3 that is doped with various dopants.
Bismuth (Bi, melting point = 271.4 °C), selenium (Se, melting point = 220 °C), indium (In, melting point = 156.6 °C), and tin (Sn, melting point = 231.9 °C) have similar melting points. The covalent radii of Bi (~148 pm), In (~142 pm), and Sn (~139 pm) are similar. The thermal–CVD process was used to synthesize pristine Bi2Se3 nanoplatelets on Al2O3(100) substrates by the vapor–solid growth mechanism at 600 °C under 1.2 × 10−2 Torr. The photocurrents in In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets have not been studied. Dopants In and Sn are used herein to fabricate In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets. The types of defects and effects of In and Sn dopants on the Bi2Se3 crystal structure were investigated by XRD, HRTEM, XPS, and Raman spectroscopy. The obtained optical bandgaps were estimated using an FPA–FTIR spectrometer. The photosensitivities were systematically obtained by measuring the photocurrents under UV and red light.

2. Materials and Methods

2.1. Synthesis of Pristine and In-, Sn-Doped Bi2Se3 Nanoplatelets

Supplementary Figure S1 schematically depicts the synthetic system. Pristine Bi2Se3 nanoplatelets were synthesized on Al2O3(100) substrate (0.5 × 0.5 mm2) by the catalyst-free V–S mechanism using a thermally activated process in a quartz tube furnace. A 0.2 g mixture of precursor powders of high-purity 0.1 g Bi (Merck, 99%, 4.78 × 10−4 mole, Darmstadt, Germany) and 0.1 g Se (Alfa Aesar, 99%, 1.27 × 10−3 mole, Ward Hill, MA, USA) was placed on an alumina boat in the heating zone at the center of a quartz tube and heated at a rate of 25 °C/min under 1.2 × 10−2 Torr to 600 °C, which was maintained for 60 min. Al2O3(100) substrate was placed upstream in the quartz tube at about 150 °C, about 21 cm away from the alumina boat. The pristine Bi2Se3 nanoplatelets thus formed were then deposited on the Al2O3(100) substrate. After the 60 min deposition process, the deposition system was subsequently cooled to room temperature. The starting materials of In-doped Bi2Se3 were 0.1 g Bi, 0.1 g Se, and 25 mg of high-quality In as dopant (Alfa Aesar, 99.99%, 2.17 × 10−4 mole, USA); those of Sn-doped Bi2Se3 included 25 mg Sn as dopant (Alfa Aesar, 99.8%, 2.11 × 10−4 mole, USA); that of (In, Sn)-doped Bi2Se3 included 12.5 mg co-dopants In (1.09 × 10−4 mole) and 12.5 mg Sn (1.05 × 10−4 mole). The In-, Sn-, and (In, Sn)-dopedBi2Se3 nanoplatelets were synthesized in the same manner as the pristine Bi2Se3 nanoplatelets.

2.2. Characterization of Nanoplatelets

The phase and crystal structures of the undoped, as well as the In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets were determined using a mass absorption coefficient glancing incident X-ray diffractometer with an incidence angle of 0.5° (λ = 0.154 nm, 30 A, 40 kV, Bruker D2 PHASER) and a high-resolution transmission electron microscope (JEOL, JEM-3000 F, Tokyo, Japan). The chemical binding energies and vibration modes of the chemical bonds were obtained using an X-ray photoelectron spectroscope (XPS, Perkin-Elmer model PHI1600 system, Waltham, MA, USA) and a Raman spectroscope (3D Nanometer Scale Raman PL Microspectrometer, Tokyo Instruments, Inc., Tokyo, Japan) with a semiconductor laser at an excitation wavelength of 633 nm. The surface morphology and EDS spectra were obtained using FESEM (ZEISS ULTRA PLUS, Carl Zeiss Microscopy GmbH, Oberkochen, Germany). The optical absorbance values were recorded using a focal plane array–FTIR spectrometer (FPA–FTIR, Bruker Vertex 70V, Hyperion 3000, 64 × 64 MCT Focal Plane Array, Bruker Optik GmbH, Ettlingen, Germany).

2.3. Photocurrent Analysis

Photocurrents were measured using a semiconductor I–V property analyzer (Keysight B2901A Precision Source/Measure Unit 100 fA, Keysight Technologies, Santa Rosa, CA, USA) under irradiations of UV or the red light at atmospheric pressure and room temperature, while the bias voltage was kept at 0 V during the photocurrent measuring. The irradiation sources were 30 cm long UV (8 W, λ = 365 nm) and red (5 W, λ = 700–900 nm) LED lamps. The distance between each lamp and the sample was 20 cm. Supplementary Figure S2 schematically depicts the photocurrent measuring system. The sample was placed in a closed box/darkroom to eliminate any effect of ambient light. The silver paste was dropped and deposited onto the surface of each sample of the nanoplatelets and connected to the photocurrent analyzer using copper wires. The photocurrent of each sample was measured in five runs; in each run, the light was on for 10 s and off for 10 s.

3. Results

3.1. XRD Analysis

Figure 1a presents the XRD patterns of the pristine and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets. These nanoplatelets have the typical rhombohedral Bi2Se3 structure (JCPDS 89-2008). Table 1 presents the lattice constants a, b, c, and the c/a ratio. The lattice constants are calculated as the formula 1 d h k l 2 = 4 3 h 2 + k 2 + h k + l 2 a c 2 1 a 2 , where h, k, and l are the Miller indices, and a and c are the lattice constants. For the In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets, lattice constant a deviates by −1.312, −0.037, and −1.776%, respectively, and lattice constant c deviates by 0.315, −2.339, and 0.148%, respectively, from their values for the undoped nanoplatelets. The decreases in a are attributed to the substitution of Bi (covalent radius: 148 ± 4 pm) rather than Se (covalent radius: 120 ± 4 pm) by the In (covalent radius: 142 ± 5 pm) or Sn (covalent radius: 139 ± 4 pm) dopant. The lowering of c by doping with Sn is attributed to the substitution of Bi by Sn [42]. The bond length of Bi-Se is 2.86/3.05 Å, and the gap between each QL is 2.62 Å [43]; therefore, the increase in c upon doping with In or (In, Sn) is attributable to the intercalation of In atoms between the QLs [44]. The decrease and increase of the lattice constants a and c, respectively, imply that the In and Sn dopants affected the crystallinity of the Bi2Se3 nanoplatelets. Defects are formed by In at Bi lattice sites ( I n B i ), Sn at Bi lattice sites ( S n B i ), and In in vacancies ( I n V ).

3.2. Structural and Surface Morphology Analyses

Figure 2a–d and the respective thickness of Figure S3a–d show the cross-sectional and the plane-view images of pristine and In-, Sn-, and (In,Sn)-doped Bi2Se3 nanoplatelets, respectively. The overall thickness of the samples of pristine and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets are, respectively, of the 3.391, 0.859, 1.066, and 0.563 μm, as shown in Figure S3a–d. The hexagonal Bi2Se3 nanoplatelets (Figure 2a and Figure S3a) appear in similar morphologies after the doping.
The In- (Figure 2b and Figure S3b) and (In,Sn)-doped Bi2Se3 (Figure 2d and Figure S3d) reveal less well-defined hexagonal structures; however, the Sn-doped Bi2Se3 (Figure 2c and Figure S3c) exhibits a very well-defined hexagonal structure. On average, the Bi2Se3 nanoplatelets are unequivocally hexagonal-like in shape, typical of the rhombohedral structure. The average thickness (40 nanoplatelets) and average diameter (40 nanoplatelets) of the pristine and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets are, respectively, listed in Table 2.
Table 2 shows EDS results for Bi, Se, In, and Sn, which reveal that the ratio Bi:Se increases upon the addition of dopants. This result is attributable to the substitution of Bi by In and/or Sn dopants. Under the Se-rich condition (mole ratio, Bi/Se = 0.755) in this work, the formation energy of V S e defects increases from 1.14 to 2.16 eV, and that of V B i decreases from 4.13 to 2.60 eV [42]; Bi is thus determined to be substituted by In and/or Sn dopants, consistent with the XRD results. Figure 3 presents the HRTEM images and SAD patterns of pristine (Figure 3a), and In- (Figure 3b), Sn- (Figure 3c), and (In, Sn)-doped Bi2Se3 nanoplatelets (Figure 3d). Table 3 provides d-spacings and diffraction planes. These results are consistent with the rhombohedral Bi2Se3 structure and confirm that the dopants, such as In and Sn, have no effect on the crystal structure of Bi2Se3.

3.3. XPS Analysis

Figure 4 presents the XPS results for Bi 4f, Se 3d, In 3d, and Sn 3d. Figure 4a shows the binding energy of Bi 4f. Peaks at 157.9 and 163.2 eV are attributed to the Bi 4f7/2 and Bi 4f5/2 orbitals in the Bi2Se3 phase [45]. Binding energies 159.1 and 164.3 eV are associated with the Bi 4f7/2 and Bi 4f5/2 orbitals in the Bi2O3 phase [46,47]. Figure 4b presents the XPS spectra of Se 3d5/2 and Se 3d3/2 and the binding energies of 53.6 and 54.6 eV are associated with the Bi2Se3 phase [45]. The peak at 58.9 eV is attributed to the SeO2 phase [48,49]. The samples are stored in the ambient environment, causing the Bi2O3 and SeO2 phases to form on the surface of the nanoplatelets. These results confirm the formation of the Bi2Se3 phase by the thermal V–S mechanism. Figure 4c displays the binding energies of the In 3d5/2 and In 3d3/2 orbitals, 444.7 and 452.3 eV, associated with the In-Se bond [50]. This finding reveals that In3+ was doped into the Bi2Se3 structure. The typical Bi2Se3 structure consists of a stack of several quintuple layers (QLs). Each QL comprises Se-Bi-Se’-Bi-Se. In3+ has two possible positions as a dopant: (1) In3+ may substitute at the Bi3+ lattice sites, producing the neutral defect ( I n 3 + B i 3 + ) and (2) In3+ intercalates between pairs of QLs, ind icating the possible formation of potentially forming the donor defect ( I n 3 + V 0 ), where V is the vacancy in the Van der Waals gap. Therefore, In-Se bonds form inside the Bi2Se3 structure or between QLs. The peak at 441.5 eV is ascribed to the Bi3+ 4d5/2 orbital [51]. Figure 4d presents XPS spectra of Sn 3d. Peaks at 485.1 and 493.7 eV are associated with the Sn2+ 3d5/2 and Sn2+ 3d3/2 orbitals, respectively, of the SnSe phase [52]. Peaks at 486.6 and 495.1 eV are associated with the Sn4+ 3d5/2 and Sn4+ 3d3/2 orbitals of the SnSe2 phase [53]. Accordingly, the Sn dopants substitute at some of the Bi lattice sites within the Bi2Se3 crystal structure and bond with Se to form Sn-Se bonds. The integral area of the Sn4+ in the XPS spectrum exceeds that of Sn2+ (as shown in Figure 4d), implying that the Sn4+ contents are higher than the Sn2+ content. The XRD results reveal that the defect ( S n B i ) is formed in the Bi2Se3 structure during the thermal V–S process. The concentration of the donor defects ( S n 4 + B i 3 + ) should be higher than those of the acceptor defects ( S n 2 + B i 3 + ).

3.4. Raman Spectra

Figure 5a presents the typical Raman active mode at A1g1, Eg2, and A1g2 of the rhombohedral Bi2Se3 structure [6,54]. No other peaks besides Bi-Se vibrational modes are observed; hence, the dopant does not change the crystal structure of the nanoplatelets or form a second-phase compound, consistent with the XRD results. The formation of Bi2Se3 nanoplatelets is thus confirmed.
The typical structure of Bi2Se3 is layered; each layer comprises five monoatomic planes and is, therefore, a quintuple layer (QL). The QL is denoted as A1-B1-A1’-B1-A1, as shown in Figure 5b [55]. A1 and A1’ are the Se atoms; B1 is the Bi atom. Covalent bonds dominate the binding within each QL; Van der Waals’ forces dominate the bonds between QLs [4,5]. The inset in Figure 5a schematically depicts the Raman peaks of A1g1, Eg2, and A1g2 [56]. A1g is a symmetric out-of-plane stretching mode associated with the vibration of A1-B1 atoms in the same (A1g1 mode) or the opposite (A1g2 mode) direction. A1g2 has a shorter atomic displacement than A1g1. Therefore, the A1g2 mode has higher phonon energy than the A1g1 mode [56]. Eg2 is a symmetric in-plane bending mode and shearing the upper two layers of A1-B1 atoms that vibrate in the opposite direction increasing the atomic displacement to a value greater than that in the A1g2 mode but smaller than that in the A1g1 mode. Thus, the Eg2 mode has a phonon energy between those of the A1g2 and A1g1 modes [56].
Figure 5c shows the variations in characteristic Raman peaks at A1g1, Eg2, and A1g2 with the species of dopant. A comparison with pristine Bi2Se3 nanoplatelets reveals that both In and Sn dopants cause a redshift of the peaks of A1g1, Eg2, and A1g2, whereas (In, Sn) co-dopants do not shift the A1g2 or Eg2 peak but do cause a redshift in the A1g1 peak. Table 4 presents the Raman peaks of pristine and doped Bi2Se3 nanoplatelets, showing the redshifts with the various dopants. The redshifts of the Raman peaks are frequently suggested to involve the heavier atomic weight and/or high-electronegativity dopant to be doped in [57]. The atomic weights of Bi, Se, In, and Sn are 209.0, 78.76, 114.8, and 118.7 (g/mole), and their electronegativities are 2.02, 2.55, 1.96, and 1.78, respectively. The Raman peaks of the doped Bi2Se3 nanoplatelets thus exhibit a redshift. As shown in Table 4, Eg2, A1g2, and especially A1g1 peaks are significantly redshifted by the addition of different dopants. The redshift is attributed to the substitution of Bi with dopant In or Sn, which has less weight and a lower electronegativity.

3.5. Photocurrent under UV and Red Light

3.5.1. Analysis under UV and Red Illumination

Figure 6a shows the photocurrents in undoped, as well as In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets under UV light. All samples pass a photocurrent (Iphoto) that increases rapidly from Idark (~0.62 × 10−11–1.65 × 10−11 A) to the maximum Imax (~8 × 10−10–1 × 10−9 A) and then suddenly falls to a stable value (Istable) when the light is turned on. Istable is clearly independent of the exposure time and, for the undoped, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets, it is 0.40 × 10−10, 0.65 × 10−10, 1.60 × 10−10, and 5.20 × 10−10 A, respectively, indicating that the dopants can increase the photocurrent of the pristine Bi2Se3 nanoplatelets. In particular, the co-dopants In and Sn increase it by a factor of more than 13 to, for example, 5.20 × 10−10 A. The Iphoto/Idark ratios in the undoped, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets are, respectively, estimated as 7.66, 9.29, 15.8, and 30.7, as shown in Figure S4a and listed in Table 5, implying that the co-dopants of In and Sn enhance the photoresponsibility of the Bi2Se3 nanoplatelets 4.01 times, which is higher than the pristine one. The decay of the current ΔIdecay (Imax–Istable) in the undoped, as well as In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets, is 8.33 × 10−10, 9.33 × 10−10, 7.76 × 10−10, and 3.21 × 10−10 A, respectively. A smaller ΔIdecay is attributed to a longer electron lifetime and a higher concentration of electrons. Both the rise time (τr) and fall time (τf) are taken by the photocurrent to rise or to fall from 10% to 90% or from 90% to 10%, respectively, of its maximum photocurrent value, as an example of the (In,Sn)-doped Bi2Se3 in Figure 6c. The average τr for the undoped, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets are, respectively, evaluated as 0.053, 0.051, 0.044, and 0.041 sec; the average τf are 0.049, 0.050, 0.054, and 0.066 sec. A shorter τr is attributable to the higher photosensitivity. The decay time (tdecay) is taken from the Imax to Istable, which depends on the recombination rate of the photo-induced electrons and holes [59]. The average tdecay of the undoped, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets are estimated as 0.091, 0.097, 0.099, and 0.105 sec. A longer tdecay corresponds to slower recombination. Therefore, the co-dopants of In and Sn suppress the recombination rate of the photoinduced electrons and holes in the pristine Bi2Se3 nanoplatelets. The detailed variations of the log10(time) versus the photocurrents, which are recorded in the first run of the light-on/light-off cycle, between the pristine, and In-, Sn-, and (In,Sn)-doped Bi2Se3 nanoplatelets are presented in Figure 6c. Table 5 presents relevant details. These results suggest that the dopants, and especially the co-dopants In and Sn, in Bi2Se3 nanoplatelets, have various favorable effects, which are (1) extending the electron lifetime, (2) increasing the electron concentration, (3) promoting surface electronic transportation, and (4) improving the photo-sensitivity of the Bi2Se3 nanoplatelets.
Figure 6b presents variations of the photocurrent of the undoped, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets under red light. Both undoped and In-doped Bi2Se3 generate no photocurrent, whereas Sn- and (In, Sn)-doped Bi2Se3 generate a photocurrent of 0.5 × 10−10 and 3.5 × 10−10 A when the red light is turned on. These results reveal that the photosensitivity of Bi2Se3 nanoplatelets to a red light is greatly improved by the dopants, and especially by the co-dopants In and Sn. The Iphoto/Idark ratios of the undoped, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets are, respectively, estimated as 1, 1, 20.9, and 52.2, as shown in Figure S4b. The co-dopants of In and Sn enhance the Bi2Se3 nanoplatelets 52.2 times higher than that of the undoped one.

3.5.2. Effects of the Defect Structure and Optical Bandgap on Photocurrent

Figure 6d shows that the optical bandgaps, estimated from the Tauc plot [60], of the undoped, as well as In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets, are 0.973, 0.641, 0.717, and 0.548 eV, respectively. These bandgaps are estimated by the following equation of α h ν n = A h ν E g , where α is the absorption coefficient, h is the Planck’s constant, ν is the light frequency, n is the characteristic coefficient of materials, A is a constant, and E g is the bandgap. For the direct bandgap of the Bi2Se3, n is 2. Their absorbance spectra, which were recorded by FPA–FTIR, are shown in Figure S5. Each individual Tauc plot of the pristine, In-, Sn-, and (In,Sn)-doped Bi2Se3 nanoplatelets is demonstrated in Figure S6. These small bandgaps show that the incident UV (~3.4 eV) and red (~1.37–1.77 eV) light easily generate photo-induced electrons and holes. A photocurrent is therefore detectable in all of the samples of interest. However, the variously doped Bi2Se3 nanoplatelets exhibit significantly different photocurrents. Defect structures are the dominant factor that affects the photocurrent. Possible defects are ( I n 3 + B i 3 + ), ( I n 3 + V 0 ), ( S n 4 + B i 3 + ), ( V 0 B i 3 + ), and ( S n 2 + B i 3 + ), where ( I n 3 + B i 3 + ) a neutral defect; ( I n 3 + V 0 ) and ( S n 4 + B i 3 + ) are donor defects and supply additional electrons; and ( V 0 B i 3 + ) and ( S n 2 + B i 3 + ) are acceptor defects and supply additional holes. ( V 0 B i 3 + ) is the main defect in undoped Bi2Se3, whose photocurrent under UV or red light is, therefore, the lowest or ~0 A. ( I n 3 + B i 3 + ) and ( I n 3 + V 0 ) instead of ( V 0 B i 3 + ) are the main defects in In-doped Bi2Se3, and therefore, a photocurrent under UV is detected therein. However, no photocurrent is detected under red light, owing to the lower electron concentration and faster recombination of electrons and holes. ( S n 4 + B i 3 + ) rather than ( S n 2 + B i 3 + ) and ( V 0 B i 3 + ) is the main defect in Sn-doped Bi2Se3 nanoplatelets; hence, more electrons than undoped and In-doped Bi2Se3 nanoplatelets can be supplied and generate a high photocurrent under UV or red light. Accordingly, Sn-doped Bi2Se3 nanoplatelets exhibit a higher photocurrent than undoped and In-doped Bi2Se3 nanoplatelets. ( V 0 B i 3 + ) and ( S n 2 + B i 3 + ) are not the main defects. Rather, ( I n 3 + V 0 ), ( S n 4 + B i 3 + ), and ( I n 3 + B i 3 + ) are the main defects in (In, Sn)-doped Bi2Se3 nanoplatelets. ( I n 3 + V 0 ) and ( S n 4 + B i 3 + ) can supply more additional electrons than undoped, as well as In- and Sn-doped Bi2Se3 nanoplatelets, to generate the highest photocurrent. Therefore, the (In, Sn)-doped Bi2Se3 nanoplatelets have the highest photocurrents under UV and red light than do undoped, as well as In-, and Sn-doped Bi2Se3 nanoplatelets. Based on the above discussion, the reduced optical bandgap of the doped Bi2Se3 nanoplatelets is a minor factor that affects the photocurrent. The type of defect that is generated by doping has a greater effect on the photodetection sensitivity than the corresponding reduction of the optical bandgap.

4. Conclusions

(In, Sn)-doped Bi2Se3 nanoplatelets under UV (8 W) and red light (5 W) have a higher Iphoto/Idark ratio of 30.7 and 52.2 and a stable photocurrent of 5.20 × 10−10 and 0.35 × 10−10 A, respectively, higher than that of the undoped Bi2Se3 nanoplatelets (UV light: 7.66, 0.4 × 10−10 A; red light: 1, 2.38 × 10−13 A). (In, Sn)-doped Bi2Se3 nanoplatelets have a shorter photocurrent rise time (0.041 s) and a longer decay time (0.105 s) than undoped Bi2Se3 nanoplatelets (0.053 and 0.091 s) by about 22.6% and 15.3%, respectively. These results suggest that photodetection under UV and red light by pristine Bi2Se3 nanoplatelets can be improved by doping with In and Sn. The optical bandgap of pristine Bi2Se3 nanoplatelets is 0.973 eV; it can be reduced to 0.641, 0.717, and 0.548 eV, corresponding to reductions of 34.1%, 26.3%, and 43.7% by doping with In, Sn, and both In and Sn. Based on XRD, XPS, FESEM–EDS, and Raman spectra, ( I n 3 + V 0 ), ( S n 4 + B i 3 + ), ( I n 3 + B i 3 + ), ( V 0 B i 3 + ), and ( S n 2 + B i 3 + ) were formed during the synthesis of nanoplatelets, and structural defects ( I n 3 + V 0 ) and ( S n 4 + B i 3 + ) significantly improved the photocurrent of (In, Sn)-doped Bi2Se3 nanoplatelets under UV and red light. This work also reveals that In or Sn dopant has no effect on the crystal structure of rhombohedral Bi2Se3. These results suggest that the photodetection sensitivity of Bi2Se3 nanoplatelets is dominated by the defect structures that are generated by doping, as well as by the consequent reduction of optical bandgap energy. The co-dopants In and Sn further enhance the ability of Bi2Se3 nanoplatelets to respond to UV and red light.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/nano11051352/s1, Figure S1: Schematic of the furnace, crucible, substrate, and operating conditions, Figure S2: Schematic diagram of the photocurrent measurement system, Figure S3: FESEM cross-sectional images of (a) pristine, (b) In-, (c) Sn-, and (d) (In,Sn)-doped Bi2Se3 nanoplatelets, Figure S4: Comparisons of the current in logarithmic scale vs. time under the (a) UV and (b) red light, Figure S5: FPA–FTIR absorbance and (b) estimated optical bandgaps of the pristine, and In-, Sn-, and (In,Sn)-doped Bi2Se3 nanoplatelets, Figure S6: The estimated optical bandgaps of (a) pristine, and (b) In-, (c) Sn-, and (d) (In,Sn)-doped Bi2Se3 nanoplatelets, Table S1: Lists of the Bi2Se3-based photodetectors, Table S2: Lists of the Bi2Se3 doped with various elements and the variation of the bandgap energy.

Author Contributions

Conceptualization, F.-S.S. and H.C.S.; methodology, F.-S.S., H.C.S., and C.-C.W.; software, C.-C.W.; validation, C.-C.W.; formal analysis, C.-C.W.; investigation, C.-C.W.; resources, H.C.S. and F.-S.S.; data curation, C.-C.W.; writing—original draft preparation, C.-C.W.; writing—review and editing, H.C.S. and F.-S.S.; visualization, C.-C.W.; supervision, H.C.S. and F.-S.S.; project administration, H.C.S.; funding acquisition, H.C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology (MOST) of Taiwan, the Republic of China, Grant Number MOST 108-2221-E-034-010.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors gratefully acknowledge the financial support of the Ministry of Science and Technology (MOST) of Taiwan, the Republic of China, under contract MOST 108-2221-E-034-010.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lawal, A.; Shaari, A.; Ahmed, R.; Jarkoni, N. First-principles many-body comparative study of Bi2Se3 crystal: A promising candidate for broadband photodetector. Phys. Lett. A 2017, 381, 2993–2999. [Google Scholar] [CrossRef]
  2. Ren, Z.; Taskin, A.A.; Sasaki, S.; Segawa, K.; Ando, Y. Large bulk resistivity and surface quantum oscillations in the topological insulator. Phys. Rev. B 2010, 82, 241306. [Google Scholar] [CrossRef] [Green Version]
  3. Zhang, H.; Liu, C.X.; Qi, X.L.; Dai, X.; Fang, Z.; Zhang, S.C. Topological insulators in Bi2Se3, Bi2Te3 and Sb2Te3 with a single Dirac cone on the surface. Nat. Phys. 2009, 5, 438–442. [Google Scholar] [CrossRef]
  4. Zhang, M.; Liu, L.G.; Wang, D.; An, X.Y.; Yang, H. Enhancement of surface state contribution in cadmium doped Bi2Se3 single crystal. J. Alloys Compd. 2019, 806, 180–186. [Google Scholar] [CrossRef]
  5. Salvato, M.; Scagliotti, M.; De Crescenzi, M.; Castrucci, P.; De Matteis, F.; Crivellari, M.; Cresi, S.P.; Catone, D.; Bauch, T.; Lombardi, F. Stoichiometric Bi2Se3 topological insulator ultra-thin films obtained through a new fabrication process for optoelectronic applications. Nanoscale 2020, 12, 12405. [Google Scholar] [CrossRef]
  6. Irfan, B.; Sahoo, S.; Gaur, A.P.S.; Ahmadi, M.; Guinel, M.J.F.; Katiyar, R.S.; Chatterjee, R. Temperature dependent Raman scattering studies of three dimensional topological insulators Bi2Se3. J. Appl. Phys. 2014, 115, 173506. [Google Scholar] [CrossRef]
  7. Schönherr, P.; Collins-McIntyre, L.J.; Zhang, S.; Kusch, P.; Reich, S.; Giles, T.; Daisenberger, D.; Prabhakaran, D.; Hesjedal, T. Vapour-liquid-solid growth of ternary Bi2Se2Te nanowires. Nanoscale Res. Lett. 2014, 9, 127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Meyer, N.; Geishendorf, K.; Walowski, J.; Thomas, A.; Munzenberg, M. Photocurrent measurements in topological insulator Bi2Se3 nanowires. Appl. Phys. Lett. 2020, 116, 172402. [Google Scholar] [CrossRef]
  9. Yue, C.; Jiang, S.; Zhu, H.; Chen, L.; Sun, Q.; Zhang, D.W. Device applications of synthetic topological insulator nanostructures. Electronics 2018, 7, 225. [Google Scholar] [CrossRef] [Green Version]
  10. Tian, W.; Yu, W.; Shi, J.; Wang, Y. The property, preparation and application of topological insulators: A Review. Materials 2017, 10, 814. [Google Scholar] [CrossRef] [Green Version]
  11. Mishra, S.K.; Satpathy, S.; Jepsen, O. Electronic structure and thermoelectric properties of bismuth telluride and bismuth selenide. J. Phys. Condens. Matter. 1997, 9, 461–470. [Google Scholar] [CrossRef]
  12. Hsieh, D.; Xia, Y.; Wray, L.; Qian, D.; Pal, A.; Dil, J.H.; Meier, F.; Osterwalder, J.; Bihlmayer, G.; Kane, C.L.; et al. First direct observation of spin-textures in topological insulators: Spin-resolved ARPES as a probe of topological quantum spin Hall effect and Berry’s phase. Science 2009, 323, 919–933. [Google Scholar] [CrossRef] [Green Version]
  13. Fei, F.; Zhang, S.; Zhang, M.; Shah, S.A.; Song, F.; Wang, X.; Wang, B. The material efforts for quantized Hall devices based on topological insulators. Adv. Mater. 2020, 32, 1904593. [Google Scholar] [CrossRef] [PubMed]
  14. Yao, J.; Shao, J.; Wang, Y.; Zhao, Z.; Yang, G. Ultra-broadband and high response of the Bi2Te3-Si heterojunction and its application as a photodetector at room temperature in harsh working environments. Nanoscale 2015, 7, 12535–12541. [Google Scholar] [CrossRef]
  15. Jung, M.; Lee, J.; Koo, J.; Park, J.; Song, Y.W.; Lee, K.; Lee, S.; Lee, J.H. A femtosecond pulse fiber laser at 1935 nm using a bulk-structured Bi2Te3 topological insulator. Opt. Express 2014, 22, 7865–7874. [Google Scholar] [CrossRef]
  16. Liu, B.; Xie, W.; Li, H.; Wang, Y.; Cai, D.; Wang, D.; Wang, L.; Liu, Y.; Li, Q.; Wang, T. Surrounding sensitive electronic properties of Bi2Te3 nanoplates-potential sensing applications of topological insulators. Sci. Rep. 2014, 4, 4639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Götte, M.; Joppe, M.; Dahm, T. Pure spin current devices based on ferromagnetic topological insulators. Sci. Rep. 2016, 6, 36070. [Google Scholar] [CrossRef] [PubMed]
  18. Fan, Y.; Upadhyaya, P.; Kou, X.; Lang, M.; Takei, S.; Wang, Z.; Tang, J.; He, L.; Chang, L.T.; Montazeri, M.; et al. Magnetization switching through giant spin-orbit torque in a magnetically doped topological insulator heterostructure. Nat. Mater. 2014, 13, 699–704. [Google Scholar] [CrossRef]
  19. Stern, A.; Lindner, N.H. Topological quantum computation-from basic concepts to first experiments. Science 2013, 339, 179–1184. [Google Scholar] [CrossRef] [PubMed]
  20. Qi, X.L.; Zhang, S.C. Topological insulators and superconductors. Rev. Mod. Phys. 2011, 83, 1057–1110. [Google Scholar]
  21. Lee, C.W.; Kim, G.H.; Kang, S.G.; Kang, M.A.; An, K.S.; Kim, H.; Lee, Y.K. Growth behavior of Bi2Te3 and Sb2Te3 thin films on graphene substrate grown by plasma-enhanced chemical vapor deposition. Phys. Status Solidi RRL 2017, 11, 1600369. [Google Scholar] [CrossRef]
  22. Hwang, T.H.; Kim, H.S.; Kim, H.; Kim, J.S.; Doh, Y.J. Electrical detection of spin-polarized current in topological insulator Bi1.5Sb0.5Te1.7Se1.3. Curr. Appl. Phys. 2019, 19, 917–923. [Google Scholar] [CrossRef] [Green Version]
  23. Fei, F.; Wei, Z.; Wang, Q.; Lu, P.; Wang, S.; Qin, Y.; Pan, D.; Zhao, B.; Wang, X.; Sun, J.; et al. Solvothermal synthesis of lateral heterojunction Sb2Te3/Bi2Te3 nanoplates. Nano Lett. 2015, 15, 5905–5911. [Google Scholar] [CrossRef] [PubMed]
  24. Bansal, N.; Cho, M.R.; Brahlek, M.; Koirala, N.; Horibe, Y.; Jing, C.; Wu, W.; Yun, D.P.; Oh, S. Transferring MBE-grown topological insulator films to arbitrary substrates and metal-insulator transition via Dirac gap. Nano Lett. 2014, 14, 1343–1348. [Google Scholar] [CrossRef] [Green Version]
  25. Zastrow, S.; Gooth, J.; Boehnert, T.; Heiderich, S.; Toellner, W.; Heimann, S.; Schulz, S.; Nielsch, K. Thermoelectric transport and Hall measurements of low defect Sb2Te3 thin films grown by atomic layer deposition. Semicond. Sci. Technol. 2013, 28, 035010. [Google Scholar] [CrossRef]
  26. Bendt, G.; Zastrow, S.; Nielsch, K.; Mandal, P.S.; Sánchezbarriga, J.; Rader, O.; Schulz, S. Deposition of topological insulator Sb2Te3 films by an MOCVD process. J. Mater. Chem. A 2014, 2, 8215–8222. [Google Scholar] [CrossRef] [Green Version]
  27. Le, P.H.; Wu, K.H.; Luo, C.W.; Leu, J. Growth and characterization of topological insulator Bi2Se3 thin films on SrTiO3 using pulsed laser deposition. Thin Solid Films 2013, 534, 659–665. [Google Scholar] [CrossRef]
  28. Fang, B.; Zeng, Z.; Yan, X.; Hu, Z. Effects of annealing on thermoelectric properties of Sb2Te3 thin films prepared by radio frequency magnetron sputtering. J. Mater. Sci. Mater. Electron. 2013, 24, 1105–1111. [Google Scholar] [CrossRef]
  29. Nam, H.; Xu, Y.; Miotkowski, I.; Tian, J.; Chen, Y.P.; Liu, C.; Hasan, M.Z.; Zhu, W.; Fiete, G.A.; Shih, C.K. Microscopic investigation of Bi2-xSbxTe3-xSey systems: On the origin of a robust intrinsic topological insulator. J. Phys. Chem. Solids 2019, 128, 251–257. [Google Scholar] [CrossRef]
  30. Bhattacharyya, B.; Sharma, A.; Kaur, M.; Singh, B.P.; Husale, S. Highly responsive broadband photodetection in topological insulator-Carbon nanotubes based heterostructure. J. Alloys Compd. 2021, 851, 156759. [Google Scholar] [CrossRef]
  31. Zhang, H.; Song, Z.; Li, D.; Xu, Y.; Li, J.; Bai, C.; Man, B. Near-infrared photodetection based on topological insulator P-N heterojunction of SnTe/Bi2Se3. Appl. Surf. Sci. 2020, 509, 145290. [Google Scholar] [CrossRef]
  32. Li, X.M.; Zhao, K.; Ni, H.; Zhao, S.Q.; Xiang, W.F.; Lu, Z.Q.; Yue, Z.J.; Wang, F.; Kong, Y.C.; Wong, H.K. Voltage tunable photodetecting properties of La0.4Ca0.6MnO3 films grown on miscut LaSrAlO4 substrates. Appl. Phys. Lett. 2010, 97, 044104. [Google Scholar] [CrossRef]
  33. Huang, S.M.; Huang, S.J.; Yan, Y.J.; Yu, S.H.; Chou, M.; Yang, H.W.; Chang, Y.S.; Chen, R.S. Extremely high-performance visible light photodetector in the Sb2SeTe2 nanoflake. Sci. Rep. 2017, 7, 45413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Sharma, A.; Bhattacharyya, B.; Srivastava, A.K.; Senguttuvan, T.D.; Husale, S. High performance broadband photodetector using fabricated nanowires of bismuth selenide. Sci. Rep. 2016, 6, 19138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Yang, M.; Han, Q.; Liu, X.; Han, J.; Zhao, Y.; He, L.; Gou, J.; Wu, Z.; Wang, X.; Wang, J. Ultrahigh stability 3D TI Bi2Se3/MoO3 thin film heterojunction infrared photodetector at optical communication waveband. Adv. Funct. Mater. 2020, 30, 1909659. [Google Scholar] [CrossRef]
  36. Pejova, B.; Grozdanov, I. Chemical deposition and characterization of glassy bismuth(III) selenide thin films. Thin Solid Films 2002, 408, 6–10. [Google Scholar] [CrossRef]
  37. Pramanik, P.; Bhattacharya, R.N.A. Mondal, A chemical method for the deposition of thin films of Bi2Se3. J. Electrochem. Soc. 1980, 127, 1857. [Google Scholar] [CrossRef]
  38. Garcia, V.M.; Nair, M.T.S.; Nair, P.K.; Zingaro, R.A. Chemical deposition of bismuth selenide thin films using N,N-dimethylselenourea. Semicond. Sci. Technol. 1997, 12, 645–653. [Google Scholar]
  39. Manjulavalli, T.E.; Balasubramanian, T.; Nataraj, D. Structural and optical properties of thermally evaporated Bi2Se3 thin film. Chalcogenide Lett. 2008, 5, 297–302. [Google Scholar]
  40. Augustine, S.; Ampili, S.; Kang, J.K.; Mathai, E. Structural, electrical and optical properties of Bi2Se3 and Bi2Se(3−x)Tex thin films. Mater. Res. Bull. 2005, 40, 1314–1325. [Google Scholar] [CrossRef]
  41. Alemi, A.; Babalou, A.; Dolatyari, M.; Klein, A.; Meyer, G. Hydrothermal synthesis of NdIII doped Bi2Se3 nanoflowers and their physical properties. Z. Anorg. Allg. Chem. 2009, 635, 2053–2057. [Google Scholar] [CrossRef]
  42. Zheng, F.; Zhang, Q.; Meng, Q.; Wang, B.; Fan, L.; Zhu, L.; Song, F.; Wang, G. Electronic structures and magnetic properties of rare-earth (Sm, Gd) doped Bi2Se3. Chalcogenide Lett. 2017, 14, 551–560. [Google Scholar]
  43. Xin, X.; Guo, C.; Pang, R.; Zhang, M.; Shi, X.; Yang, X.; Zhao, Y. Theoretical and experimental studies of spin polarized carbon doped Bi2Se3. Appl. Phys. Lett. 2019, 115, 042401. [Google Scholar] [CrossRef]
  44. Kobayashi, K.; Ueno, T.; Fujiwara, H.; Yokoya, T.; Akimitsu, J. Unusual upper critical field behavior in Nb-doped bismuth selenides. Phys. Rev. B 2017, 95, 180503. [Google Scholar] [CrossRef]
  45. Zhang, G.; Qin, H.; Teng, J.; Guo, J.; Guo, Q.; Dai, X.; Fang, Z.; Wu, K. Quintuple-layer epitaxy of thin films of topological insulator Bi2Se3. Appl. Phys. Lett. 2009, 95, 053114. [Google Scholar] [CrossRef] [Green Version]
  46. Wagner, C.D.; Riggs, W.M.; Davis, L.E.; Moulder, J.F.; Muilenberg, G.E. Hand Book of X-ray Photoelectron Spectroscopy; Perkin-Elmer Corporation: Eden Prairie, MN, USA, 1979; p. 162. [Google Scholar]
  47. Meng, A.; Yuan, X.; Shen, T.; Li, Z.; Jiang, Q.; Xue, H.; Lin, Y.; Zhao, J. One-step synthesis of flower-like Bi2O3/Bi2Se3 nanoarchitectures and NiCoSe2/Ni0.85Se nanoparticles with appealing rate capability for the construction of high-energy and long-cycle-life asymmetric aqueous batteries. J. Mater. Chem. A 2019, 7, 17613. [Google Scholar] [CrossRef]
  48. Green, A.J.; Dey, S.; An, Y.Q.; O′Brien, B.; O′Mullane, S.; Thiel, B.; Diebold, A.C. Surface oxidation of the topological insulator Bi2Se3. J. Vac. Sci. Technol. A 2016, 34, 061403. [Google Scholar] [CrossRef] [Green Version]
  49. Hobbs, R.G.; Schmidt, M.; Bolger, C.T.; Georgiev, Y.M.; Fleming, P.; Morris, M.A.; Petkov, N.; Holmes, J.D.; Xiu, F.; Wang, K.L.; et al. Resist–substrate interface tailoring for generating high-density arrays of Ge and Bi2Se3 nanowires by electron beam lithography. J. Vac. Sci. Technol. B 2012, 30, 041602. [Google Scholar] [CrossRef] [Green Version]
  50. Botcha, V.D.; Hong, Y.; Huang, Z.; Li, Z.; Liu, Q.; Wu, J.; Lu, Y.; Liu, X. Growth and thermal properties of various In2Se3 nanostructures prepared by single step PVD technique. J. Alloys Compd. 2019, 773, 698–705. [Google Scholar] [CrossRef]
  51. Zhang, X.; Liu, Y.; Zhang, G.; Wang, Y.; Zhang, H.; Huang, F. Thermal decomposition of bismuth oxysulfide from photoelectric Bi2O2S to superconducting Bi4O4S3. ACS Appl. Mater. Interfaces 2015, 7, 4442–4448. [Google Scholar] [CrossRef]
  52. Lu, D.; Yue, C.; Luo, S.; Li, Z.; Xue, W.; Qi, X.; Zhong, J. Phase controllable synthesis of SnSe and SnSe2 films with tunable photoresponse properties. Appl. Surf. Sci. 2021, 541, 148615. [Google Scholar] [CrossRef]
  53. Mukhokosi, E.P.; Krupanidhi, S.B.; Nanda, K.K. Band gap engineering of hexagonal SnSe2 nanostructured thin films for infra-red photodetection. Sci. Rep. 2017, 7, 15215. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, H.; Zhang, X.; Liu, C.; Lee, S.T.; Jie, J. High-responsivity, high-detectivity, ultrafast topological insulator Bi2Se3/silicon heterostructure broadband photodetectors. ACS Nano 2016, 10, 5113–5122. [Google Scholar] [CrossRef]
  55. Kimura, K.; Hayashi, K.; Yashina, L.V.; Happo, N.; Nishioka, T.; Yamamoto, Y.; Ebisu, Y.; Ozaki, T.; Hosokawa, S.; Matsushita, T.; et al. Local structural analysis of In-doped Bi2Se3 topological insulator using X-ray fluorescence holography. Surf. Interface Anal. 2019, 51, 51–55. [Google Scholar] [CrossRef] [Green Version]
  56. Yuan, J.; Zhao, M.; Yu, W.; Lu, Y.; Chen, C.; Xu, M.; Li, S.; Loh, K.P.; Bao, Q. Raman spectroscopy of two-dimensional Bi2TexSe3 − x platelets produced by solvothermal method. Materials 2015, 8, 5007–5017. [Google Scholar] [CrossRef]
  57. Pu, X.Y.; Zhao, K.; Liu, Y.; Wei, Z.T.; Jin, R.; Yang, X.S.; Zhao, Y. Structural and transport properties of iridium-doped Bi2Se3 topological insulator crystals. J. Alloys Compd. 2017, 694, 272–275. [Google Scholar] [CrossRef]
  58. Shahil, K.M.F.; Hossain, M.Z.; Goyal, V.; Balandin, A.A. Micro-Raman spectroscopy of mechanically exfoliated few-quintuple layers of Bi2Te3, Bi2Se3, and Sb2Te3 materials. J. Appl. Phys. 2012, 111, 054305. [Google Scholar] [CrossRef] [Green Version]
  59. Basumatary, P.; Agarwal, P. Photocurrent transient measurements in MAPbI3 thin films. J. Mater. Sci. Mater. El. 2020, 31, 10047–10054. [Google Scholar] [CrossRef]
  60. Arumugam, J.; Raj, A.D.; Irudayaraj, A.A.; Thambidurai, M. Solvothermal synthesis of Bi2S3 nanoparticles and nanorods towards solar cell application. Mater. Lett. 2018, 220, 28–31. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns and (b) deviations of the lattice constants a and c of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
Figure 1. (a) XRD patterns and (b) deviations of the lattice constants a and c of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
Nanomaterials 11 01352 g001
Figure 2. FESEM images and EDS results of (a) pristine, (b) In-, (c) Sn-, and (d) (In, Sn)-doped Bi2Se3 nanoplatelets.
Figure 2. FESEM images and EDS results of (a) pristine, (b) In-, (c) Sn-, and (d) (In, Sn)-doped Bi2Se3 nanoplatelets.
Nanomaterials 11 01352 g002
Figure 3. HRTEM images and SAD patterns of (a) Bi2Se3, (b) In-, (c) Sn-, and (d) (In, Sn)-doped Bi2Se3 nanoplatelets.
Figure 3. HRTEM images and SAD patterns of (a) Bi2Se3, (b) In-, (c) Sn-, and (d) (In, Sn)-doped Bi2Se3 nanoplatelets.
Nanomaterials 11 01352 g003
Figure 4. XPS spectra of (a) Bi 4f, (b) Se 3d, (c) In 3d, and (d) Sn 3d in pristine, and Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
Figure 4. XPS spectra of (a) Bi 4f, (b) Se 3d, (c) In 3d, and (d) Sn 3d in pristine, and Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
Nanomaterials 11 01352 g004aNanomaterials 11 01352 g004b
Figure 5. (a) Raman spectra of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets; the inset is the schematic vibration modes of A1g1, Eg2, and A1g2 [56,58]. (b) Schematic layered structure of Bi2Se3 [55]. (c) Variations of Raman shift at A1g1, Eg2, and A1g2 with different dopants.
Figure 5. (a) Raman spectra of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets; the inset is the schematic vibration modes of A1g1, Eg2, and A1g2 [56,58]. (b) Schematic layered structure of Bi2Se3 [55]. (c) Variations of Raman shift at A1g1, Eg2, and A1g2 with different dopants.
Nanomaterials 11 01352 g005
Figure 6. Photocurrent plots of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets under (a) UV (8 W) and (b) red light (5 W). (c) The variations of the log10(time) vs. the photocurrents between the pristine, and In-, Sn-, and (In,Sn)- doped Bi2Se3 nanoplatelets in the first run of the light-on/light-off cycle. (d) The estimated optical bandgap energies.
Figure 6. Photocurrent plots of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets under (a) UV (8 W) and (b) red light (5 W). (c) The variations of the log10(time) vs. the photocurrents between the pristine, and In-, Sn-, and (In,Sn)- doped Bi2Se3 nanoplatelets in the first run of the light-on/light-off cycle. (d) The estimated optical bandgap energies.
Nanomaterials 11 01352 g006
Table 1. Lattice constants a and c, c/a ratios of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
Table 1. Lattice constants a and c, c/a ratios of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
SampleLattice Constantc/a RatioConcentration (×10−4 Mole)
a (=b)cBiSeInSn
Bi2Se30.41422.86776.92364.7812.700
In-doped Bi2Se30.40882.87677.03784.7812.72.170
Sn-doped Bi2Se30.41402.80066.76414.7812.702.11
(In, Sn)-doped Bi2Se30.40682.87197.05934.7812.71.091.05
Table 2. Average diameter, average thickness, and FESEM–EDS of Bi, Se, In, and Sn (atomic percent, at.%) of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
Table 2. Average diameter, average thickness, and FESEM–EDS of Bi, Se, In, and Sn (atomic percent, at.%) of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
SampleAverage Diameter
(μm)
Average Thickness
(nm)
Bi (at.%)Se (at.%)In (at.%)Sn (at.%)Bi: Se
Bi2Se31.31917.543.0156.990.000.001:1.325
In-doped Bi2Se30.96521.832.0059.338.050.001:1.854
Sn-doped Bi2Se30.91239.828.8156.900.0014.281:1.975
(In, Sn)-doped Bi2Se30.31731.529.7656.6410.982.621:1.903
Table 3. The d-spacings and diffracted planes of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
Table 3. The d-spacings and diffracted planes of pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
SampleHigh-Resolution ImagesSAD Patterns
d-Spacings (nm)PlanesDiffracted Planes
Bi2Se30.3470(0 1 2)(1 0 1)
0.3225(1 0 4)(0 1 2)
(1 0 4)
In-doped Bi2Se30.3404(0 1 2)(1 0 1)
0.2958(0 1 5)(0 1 5)
(0 1 11)
Sn-doped Bi2Se30.3011(0 1 5)(0 1 5)
(1 0 7)
(1 0 10)
(1 1 0)
(In, Sn)-doped Bi2Se30.3546(1 0 1)(1 0 1)
0.3096(0 1 5)(1 0 4)
(0 1 11)
(1 1 3)
Table 4. Characteristic Raman peaks of the pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
Table 4. Characteristic Raman peaks of the pristine, and In-, Sn-, and (In, Sn)-doped Bi2Se3 nanoplatelets.
SampleRaman Shift (cm−1)
A1g1Eg2A1g2
Bi2Se387.21133.21176.36
In-doped Bi2Se382.65131.10174.43
Sn-doped Bi2Se382.02130.54171.90
(In, Sn)-doped Bi2Se381.93133.64177.48
Table 5. Details of photocurrent measurements of the Imax, Istable, ΔIdecay, τr, tdecay, τf, and Iphoto/Idark ratios under the UV light.
Table 5. Details of photocurrent measurements of the Imax, Istable, ΔIdecay, τr, tdecay, τf, and Iphoto/Idark ratios under the UV light.
SampleImax (× 10−10 A)Istable (× 10−10 A)ΔIdecay (× 10−10 A)τr (sec)tdecay (sec)τf (sec)Iphoto/Idark
(0 V Bias Voltage)
Bi2Se38.730.408.330.0530.0910.0497.66
In-doped Bi2Se39.980.659.330.0510.0970.0509.29
Sn-doped Bi2Se39.361.607.760.0440.0990.05415.8
(In, Sn)-dopedBi2Se38.415.203.210.0410.1050.06630.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, C.-C.; Shieu, F.-S.; Shih, H.C. Photosensing and Characterizing of the Pristine and In-, Sn-Doped Bi2Se3 Nanoplatelets Fabricated by Thermal V–S Process. Nanomaterials 2021, 11, 1352. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11051352

AMA Style

Wang C-C, Shieu F-S, Shih HC. Photosensing and Characterizing of the Pristine and In-, Sn-Doped Bi2Se3 Nanoplatelets Fabricated by Thermal V–S Process. Nanomaterials. 2021; 11(5):1352. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11051352

Chicago/Turabian Style

Wang, Chih-Chiang, Fuh-Sheng Shieu, and Han C. Shih. 2021. "Photosensing and Characterizing of the Pristine and In-, Sn-Doped Bi2Se3 Nanoplatelets Fabricated by Thermal V–S Process" Nanomaterials 11, no. 5: 1352. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11051352

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop