Jiaying Jian*abc,
Honglong Chang*a,
Pengfan Dongc,
Zewen Baic and
Kangnian Zuoc
aSchool of Mechanical Engineering, Northwestern Polytechnical University, Baoding 710072, P. R. China. E-mail: jianjiaying@nwpu.edu.cn; changhl@nwpu.edu.cn
bSchool of Electronic Information and Engineering, Xi'an Technological University, Xi'an 710021, P. R. China. E-mail: jianjiaying@xatu.edu.cn
cShaanxi Key Laboratory of Photoelectric Functional Materials and Devices, Xi'an Technological University, Xi'an 710021, P. R. China
First published on 28th January 2021
Two-dimensional transition-metal dichalcogenides are considered as promising candidates for next-generation flexible nanoelectronics owing to their compelling properties. The photoelectric performance of a photodetector based on CVD-grown 2D MoS2 was studied. It is found that annealing treatment can make the photoresponsivity and specific detectivity of the CVD-grown 2D MoS2 based photodetector increase from 0.1722 A W−1 and 1014.65 Jones to 0.2907 A W−1 and 1014.84 Jones, respectively, while vulcanization can make the rise response time and fall response time decrease from 0.9013 s and 2.173 s to 0.07779 s and 0.08616 s, respectively. A method to determine the O-doping concentration in the CVD-grown 2D MoS2 has been obtained. The criterion for the CVD-grown 2D MoS2 to transition from an oxygen-doped state to a pure state has been developed. A mechanism explaining the variation in the photoelectric performance of the CVD-grown 2D MoS2 has been proposed. The CVD-grown 2D MoS2 and the annealed CVD-grown 2D MoS2 are oxygen-doped MoS2 while the vulcanized CVD-grown 2D MoS2 is pure MoS2. The variation in the photoelectric performance of CVD-grown 2D MoS2 results from differences in the O-doping concentration and the bandgap.
Due to the direct bandgap in monolayer forms, TMDs exhibit great opportunities for the preparation of optoelectronic devices.9–11 Moreover, the properties of TMDs can be improved by chemical doping,12 surface plasmonic enhancement,13 defect engineering,14 surface nano-rugging,15 vertical growth,16 heterojunctions17 and so on.
As a type of TMD, molybdenum disulfide (MoS2) has aroused extensive attention in recent studies due to its outstanding electronic, photoelectric and energy harvesting performances. MoS2 has a bandgap of 1.3 to 1.8 eV for the bulk and the monolayer crystal with the possibility of achieving a variable bandgap by tuning its layers.18 Monolayer MoS2 features a direct bandgap, high exciton binding energy, and remarkable photoluminescence excitation (PLE) spectroscopy, enabling it to be widely applied in the photoelectric industry, such as in the preparation of photodetectors, light emission diodes, phototransistors, solar cells,19–25 etc. Moreover, combining MoS2 with other semiconductors results in an efficient charge separation,26 high electron transfer rate, and increases the solar light absorption.27
Various methods for the synthesis of MoS2 nano-structures have been reported, including chemical vapor deposition (CVD),17,28 hydrothermal,29 sputtering methods,25 etc. Among them, the hydrothermal method is suitable for the large-scale preparation of few layered MoS2 and a MoS2 nanosphere, while the sputtering method can grow wafer-scale MoS2 layers. Compared with the hydrothermal and sputtering methods, the chemical vapor deposition method is favored by its capacity to cultivate high-quality and large-size monolayer MoS2. A lot of experimental research has been carried out on the photoelectric performances of photodetectors based on the CVD-grown monolayer MoS2. The reported photoelectric performances of CVD-grown monolayer MoS2 vary greatly.30–36 However, the mechanism for the variation in the photoelectric performance of the photodetector based on the CVD-grown monolayer MoS2 are unclear at present.
In this paper, the mechanism for the variation in the photoelectric performance of the CVD-grown 2D MoS2 has been revealed by studying the effects of annealing treatment and vulcanization on the photoelectric performance as well as the O-doping concentration in the CVD-grown 2D MoS2 and the CVD-grown 2D MoS2 after annealing treatment and vulcanization.
Raman spectra based on the inelastic scattering of photons can be used to investigate the internal structural properties of solids, liquids and gases. The peak of MoS2 on the Raman spectrum is related to the layer number.37,38 As the layer number of MoS2 decreases, the peak E12g corresponding to the horizontal in-plane vibration of the sulfur atoms is seen with a blue shift, and the peak A1g corresponding to the vertical vibration of sulfur atoms presents a red shift.39 Consequently, the wave number difference between A1g and E12g decreases with a decreasing layer number of MoS2. The values of the wave number difference between A1g and E12g for monolayer MoS2, two layer MoS2 and three layer MoS2 are 18.8–19.2, 22.3–23.9 and 24.4–24.5 cm−1, respectively.40,41 Fig. 1b shows the Raman spectrum of the sample. The peak at the wave number of 416.5 cm−1 is the characteristic peak for the sapphire substrate. The E12g peak of the sample occurs at the wave number of 384.7 cm−1, while the A1g peak occurs at the wave number of 403.8 cm−1. It can be calculated that the wave number difference between A1g and E12g for the sample is 19.1 cm−1, which is well in the region for monolayer MoS2. So, the sample can be identified as the monolayer MoS2.
The photoluminescence (PL) is used to characterize the defects, impurities and luminescent properties of semiconductors. The bulk MoS2 is a semiconductor with an indirect bandgap of about 1.2 eV and no fluorescence characteristic peaks.42 As MoS2 changes from the bulk into few-layers or even a monolayer, the indirect bandgap evolves into a direct bandgap, with the fluorescence efficiency greatly enhanced. The PL spectrum of monolayer MoS2 consists of a single narrow feature of 50 meV width, centered at 1.90 eV.42 Fig. 1c shows the PL spectrum of the sample. It can be seen that a strong peak presents at a wavelength of 692.7 nm, which is the characteristic peak of the sapphire substrate. Another peak presents at a wavelength of 662.4 nm. According to the equation between wavelength λ and photon energy E (E = ℏc/λ, where c is the speed of light, ℏ is Planck constant), we can obtain that the value of E corresponding to λ = 662.4 nm is 1.87 eV, which is very close to that of the reported photon energy42 for monolayer MoS2 (1.90 eV). The result of the PL spectrum further confirms that the MoS2 crystal obtained in this paper is monolayer MoS2.
With regression analysis of the data of currents tested at different bias voltages using Excel, the relational expression between the current I and the bias voltage V can be acquired:
I = aVb | (1) |
Fig. 2c and d show the ratios of the light current Il to the dark current Id at a wavelength of 405 nm and an illumination intensity of 15.73 W cm−2, respectively. It can be seen that the ratio of Il to Id increases with increasing bias voltage and illumination intensity. As the illumination intensity is constant, the ratio of Il to Id decreases with increasing wavelength in the bias voltage region from 0 V to 1.1 V. As the bias voltage is higher than 1.1 V, the ratio at a wavelength of 405 nm is higher than those at wavelengths of 520 nm of 650 nm. However, the ratio at a wavelength of 520 nm is lower than that of 650 nm.
The photoresponsivity (R) and specific detectivity (D*) are the key parameters to predict the sensitivity of a photodetector. The photoresponsivity and specific detectivity can be calculated according to the following equations:
(2) |
(3) |
As shown in Fig. 2c and d, when the bias voltage is higher than 0.2 V, the ratio of Il to Id is higher than 103. When the ratios of Il to Id are higher than 103, it can be calculated that the error is lower than 0.1% if the term Il − Id in eqn (2) is replaced by Il. Therefore, for the photodetector based on 2D MoS2 and a PDMS substrate, the term Il − Id in eqn (2) can be replaced by Il when the bias voltage is higher than 0.2 V, thus
(4) |
Substituting eqn (4) into eqn (3), the following expression can be obtained:
(5) |
The effective illuminated area for the photodetector in Fig. 2 is 65 μm2. Substituting the fitting expressions between the current I and the bias voltage V into eqn (4) and (5), we can obtain the photoresponsivity R and specific detectivity D* of the photodetector based on 2D MoS2 and the PDMS substrate. Fig. 3 shows dependences of R and D* on V for the photodetector. It can be seen that the values R and D* increase with increasing V and Ei. As the illumination intensity is a constant, R and D* decrease with increasing wavelength when the bias voltage is lower than 1.1 V. As the bias voltage is higher than 1.1 V, the values of R and D* at a wavelength of 405 nm is higher than those at wavelengths of 520 nm and 650 nm. However, the values of R and D* at a wavelength of 520 nm are lower than those at 650 nm.
Fig. 4a shows the current–time curves for the photodetector at a bias voltage of 9 V. In terms of the current–time curves, we can obtain not only the photoresponsivity R and the specific detectivity D* but also the rise response time trise and the fall response time tfall. Table 1 lists the values of R, D*, trise and tfall calculated according to Fig. 4a, and eqn (4) and (5). It can be seen that R, D*, trise and tfall increase with decreasing Ei.
Ei (W cm−2) | R (mA W−1) | D* (Jones) | trise (s) | tfall (s) |
---|---|---|---|---|
26.21 | 0.1508 | 1014.60 | 0.5505 | 1.615 |
20.97 | 0.1548 | 1014.61 | 0.9013 | 2.173 |
15.73 | 0.1807 | 1014.67 | 1.151 | 2.461 |
10.48 | 0.2230 | 1014.77 | 1.503 | 2.748 |
5.242 | 0.2661 | 1014.84 | 2.307 | 3.868 |
1.747 | 0.3922 | 1015.01 | 3.202 | 5.332 |
0.1747 | 1.127 | 1015.47 | 6.011 | 9.639 |
From regression analysis of the tested data of R, D*, trise and tfall at different illumination intensities by Excel, the following expressions can be acquired:
R = 0.5354E−0.3980i (A W−1) [R2 = 0.9942] | (6) |
D* = 1015.15E−0.3980i (Jones) [R2 = 0.9942] | (7) |
trise = −1.063ln(Ei) + 4.037(s) [R2 = 0.9960] | (8) |
tfall = −1.559ln(Ei) + 6.623(s) [R2 = 0.9903] | (9) |
Fig. 4b and c show the dependences of R, D*, trise and tfall on Ei. In Fig. 4b and c, the blank solid circles are the calculated results according to Fig. 4a and eqn (4) and (5), while the red solid circles are the calculated results according to Fig. 2a and eqn (1) and (2). The curves are the predicted results from the fitting expressions (6) and (7) based on Fig. 4a and eqn (4) and (5). Obviously, the calculated values of R and D* according to eqn (1) and (2) and the current-bias voltage curves in Fig. 4a are in agreement with the predicted results from the fitting expressions (6) and (7) are based on the current–time curves.
According to the current–time curve in Fig. 5a, the light current Il, the rise response time trise and the fall response time tfall can be determined. In terms of the light current Il in Fig. 5a and the dark current Id in the current-bias voltage curve, the photoresponsivity R and the specific detectivity D* can be obtained. Table 2 lists the determined values of Il, Id, R, D*, trise and tfall of the photodetector based on the annealed 2D MoS2. It can be seen that the annealing treatment can significantly increase the light current, the photoresponsivity and the specific detectivity of the photodetector based on the CVD-grown 2D MoS2. After the annealing treatment, the light current, photoresponsivity and specific detectivity of the photodetector based on the CVD-grown 2D MoS2 had increased from 2347 μA, 0.1722 A/W and 1014.65 Jones to 3962 μA, 0.2907 A/W and 1014.84 Jones, respectively. However, the annealing treatment has an adverse effect on the response time. After the annealing treatment, the rise response time trise and the fall response time tfall increased from 0.9013 s and 2.173 s to 3.593 and 6.531 s, respectively.
Conditions | Il (μA) | Id (μA) | R (W cm−2) | D* (Jones) | trise (s) | tfall (s) |
---|---|---|---|---|---|---|
CVD-grown | 2347 | 2.996 | 0.1722 | 1014.65 | 0.9013 | 2.173 |
Annealing | 3962 | 5.894 | 0.2907 | 1014.84 | 3.593 | 6.531 |
Vulcanization | 0.05508 | 10−2.444 | 10−5.407 | 1010.38 | 0.07779 | 0.08616 |
The vulcanization of the photodetector based on the CVD-grown 2D MoS2 was performed in the same furnace as that in the annealing treatment experiment. After the furnace was evacuated to 10−1.881 atm by a mechanical pump and then filled with argon (purity = 99.9%) at a constant rate of 70 sccm to maintain the argon pressure in the furnace over atmospheric pressure, the photodetector based on the CVD-grown 2D MoS2 and 0.3 g of sublimed sulfur was heated to 473 K and held for 30 minutes. Fig. 5b shows the current–time curve of the photodetector based on the vulcanized 2D MoS2 at a bias voltage of 9 V and an illumination intensity of 20.97 W cm−2.
In terms of the current–time curve in Fig. 5b, the light current Il, the dark current Id, the photoresponsivity R, the specific detectivity D*, the rise response time trise and the fall response time tfall of the photodetector based on the vulcanized 2D MoS2 can be determined. The determined values of Il, Id, R, D*, trise and tfall of the photodetector based on the vulcanized 2D MoS2 are listed in Table 2. It shows that vulcanization can decrease the light current, dark current, photoresponsivity and specific detectivity of the photodetector based on the 2D CVD-grown MoS2. After vulcanization, the light current, dark current, photoresponsivity and specific detectivity of the photodetector decreased from 2347 μA, 2.996 μA, 0.1722 A/W and 1014.65 Jones to 0.05508 μA, 10−2.444 μA, 10−5.407 A W−1 and 1010.38 Jones, respectively. However, vulcanization has a beneficial effect on the response time. After vulcanization, the rise response time and fall response time of the photodetector based on CVD-grown 2D MoS2 decreased from 0.9013 s and 2.173 s to 0.07779 s and 0.08616 s, respectively.
For comparison, the photoelectric performance metrics of reported 2D MoS2-based flexible photodetectors18,34–36 are listed in Table 3. Clearly, the photoresponsivities of the flexible photodetectors based on the CVD-grown 2D MoS2-PDMS substrate and the annealed 2D MoS2–PDMS substrate in the present work are higher than those of the monolayer MoS2/ODTS-PET substrate,34 the few-layer MoS2-cellulose paper substrate,35 the few layer MoS2/CQS-PI substrate36 and the few-layer MoS2/V2O5 nanowire-aluminum foil substrate.18 Though the photoresponsivity of the flexible photodetector based on the vulcanized 2D MoS2-PDMS substrate in the present work is lower than those of the reported flexible photodetectors, its response times are much lower than those in the reports.18,34–36 It should be noted that the response times of the vulcanized 2D MoS2 photodetector based on a flexible PDMS substrate are still higher than those of the 2D MoS2 photodetector based on a non-flexible substrate such as Si.17,25 The reason for this should be related to the connection between the 2D MoS2 and the substrate for the 2D MoS2 photodetector based on a flexible substrate implemented with compacting while the connection for the 2D MoS2 photodetector based on non-flexible substrate results from depositing.
Architecture | R (mW cm−2) | trise (s) | tfall (s) | Ref. |
---|---|---|---|---|
CVD-grown 2D MoS2 based on PDMS | 1.70 × 102 | 0.90 | 2.17 | Present |
Annealed 2D MoS2 based on PDMS | 2.91 × 102 | 3.59 | 6.53 | Present |
Vulcanized 2D MoS2 based on PDMS | 3.92 × 10−3 | 7.78 × 10−2 | 8.62 × 10−2 | Present |
Monolayer MoS2/ODTS based on PET | 1.60 | 0.70 | 34 | |
Few-layer MoS2 based on cellulose paper | 20.0 | 12.0 | 19.0 | 35 |
Few layer MoS2/CQS based on PI | 18.1 | 0.57 | 36 | |
Few-layer MoS2/V2O5 nanowires based on aluminum foil | 65.1 | 18 |
Fig. 5c and d show the effects of annealing and vulcanization on the photoluminescence spectrum of the 2D MoS2. In Fig. 5c and d, the peak at a wavelength of about 666.5 nm (E corresponding to λ = 666.5 nm is 1.88 eV, which is very close to the reported photon energy42 for monolayer MoS2) is the characteristic peak of the monolayer MoS2 while a peak at a wavelength of 692.7 nm is the characteristic peak of the sapphire substrate. It can be seen that annealing can enhance the PL intensity of the 2D MoS2 while vulcanization makes it lower, which is in agreement with the results reported by Mouri et al.41 They reported that the PL intensity of monolayer MoS2 was drastically enhanced by the adsorption of p-type dopants and the intensity enhancement was explained by the switching of the dominant PL process from the recombination of negative trions to the recombination of excitons through extraction of the unintentionally highly doped electrons.41 Moreover, they found the PL intensity was reduced by the adsorption of n-type dopants, which they attributed to the suppression of exciton PL through injection of the excess electrons.41
If the CVD-grown 2D MoS2 is sulfur vacancy-induced (N-doping) MoS2, the vulcanization should make the PL intensity higher. Therefore, the CVD-grown 2D MoS2 is the oxygen-doped (P-doping) MoS2 since the vulcanization makes the PL intensity lower. Consequently, it can be predicted that the variation in the photoelectric performance of the photodetector based on the CVD-grown 2D MoS2 at different states (CVD-grown, annealing, vulcanization) should result from the difference of the O-doping concentration.
MoS2(s) + O2(g) = MoO2(s) + S2(g) | (10) |
Under given conditions, whether a reaction can occur or not is predicted according to the change of Gibbs free energy for the reaction. The change of Gibbs free energy ΔG for reaction (10) can be calculated in terms of the following equation:
(11) |
Mo(s) + O2(g) = MoO2(s) | (12) |
Mo(s) + S2(g) = MoS2(s) | (13) |
Since we are discussing the oxygen-doping of MoS2, the concentration of MoO2 in MoS2 is very little and xMoS2 can be taken as 1 (i.e.: xMoS2 = 1), thus:
(14) |
The expressions for PS2, and as a function of temperature can be found in the literature:
log(PS2) = 1.934 × 10−2 − 8.189 × 10−6T + 7.112/T (atm),43 | (15) |
(16) |
(17) |
(18) |
Substituting eqn (15)–(18) into eqn (14), the expression for the molar fraction of MoO2 in MoS2 (xMoO2) as a function of temperature can be determined. Fig. 6a shows the determined xMoO2 as a function of the growth temperature for the growth of 2D MoS2 when the temperature of the sulfur source is 503 K while Fig. 6b shows the determined xMoO2 as a function of the vulcanization temperature for the vulcanization of 2D MoS2. It can be seen that xMoO2 for the growth and vulcanization of 2D MoS2 increases with increasing growth and vulcanization temperatures. At the same temperature, xMoO2 for growth is higher than that for vulcanization.
(19) |
(20) |
When NO is equal to 1, there is one O-doping atom in the 2D MoS2. So, the critical condition for the CVD-grown 2D MoS2 to transit from the pure state to the oxygen-doped state is:
(21) |
Fig. 6c shows the determined NO as a function of the growth temperature for the growth of 2D MoS2 when the temperature of the sulfur source is 503 K while Fig. 6d shows the determined NO as a function of the vulcanization temperature for the vulcanization of 2D MoS2. It can be seen that the temperature corresponding to NO = 1 for the growth of 2D MoS2 is in the region 935.3 K to 1072.1 K, while the temperature corresponding to NO = 1 for the vulcanization of 2D MoS2 is in the region 1164.7 K to 1309.6 K. As the temperature for the growth of 2D MoS2 is 1083 K in this work, which is higher than the temperature corresponding to NO = 1 for the growth of 2D MoS2, the O-doping number in a CVD-grown 2D MoS2 is higher than 1. This means that the CVD-grown 2D MoS2 in this work is the oxygen-doped MoS2.
With regression analysis of the data in Fig. 8c by Excel, the relational expression between NO and T can be acquired:
NMaxO = 10−7.751exp(0.0193T) [R = 0.9947] | (22) |
NMinO = 10−8.036exp(0.0193T) [R = 0.9947] | (23) |
Since the partial pressure of S2 is 0 at the time when the annealing treatment starts, part of the sulfur atoms should escape from the CVD-grown 2D MoS2 to the furnace atmosphere. The partial pressure of sulfur PS2 depends on the furnace volume V, the temperature T and the number of the escaped sulfur atoms NS. PS2 can be calculated in terms of the following equation:
(24) |
The escape of sulfur can make the O-doping number increase. After the annealing treatment, the increased O-doping number NAnnealingO should be equal to the decreased sulfur atom number NAnnealingS. Accord to eqn (24), an expression of the increased O-doping number NAnnealingO as the function of the partial pressure of sulfur PAnnealingS2 can be obtained:
(25) |
Substituting eqn (25) into eqn (19), we can obtain eqn (26):
(26) |
After annealing treatment, the total molar fraction of MoO2 in MoS2 is:
(27) |
Substituting eqn (27) into eqn (14), the following equation can be obtained:
(28) |
In terms of eqn (28), the equilibrium partial pressure of sulfur at the annealing temperature can be predicted. Fig. 7a shows the dependence of the equilibrium partial pressure of sulfur on the annealing temperature.
Substituting into eqn (25), the increased O-doping number NAnnealingO at different annealing temperatures can be determined. Fig. 7b shows the dependence of the determined NAnnealingO on the annealing temperature. Fig. 7c shows the dependences of the ratio of the determined NAnnealingO to the O-doping number in the CVD-grown 2D MoS2 on the annealing temperature while Fig. 7d shows the dependence of the molar fraction of the total O-doping number xtotalO in the annealed CVD-grown 2D MoS2.
It can be seen that the equilibrium partial pressure of sulfur increases with increasing annealing temperature while the increased O-doping number after annealing treatment is almost a constant when the annealing temperature is lower than 400 K. When the annealing temperature is higher than 400 K, the increased O-doping number after annealing treatment increases with increasing annealing temperature. In the annealing temperature region from 350 K to 550 K, the increased O-doping number after annealing treatment is 106.740–108.962 times higher than the O-doping number of the CVD-grown 2D MoS2 while the molar fraction of the total O-doping atom for the annealed CVD-grown 2D MoS2 is in the region 10−3.543–10−2.929.
The O-doping concentration in 2D MoS2 can influence the electronic properties. Kong et al.45 found that the substitution of oxygen for a sulfur atom in 2D MoS2 leads to a transition from a direct K–K bandgap to an indirect Γ–K bandgap. And the value of the bandgap decreases with increasing doping concentration. The above PL experimental results and the thermodynamic calculation show that the CVD-grown 2D MoS2 is the oxygen-doped MoS2, so its bandgap is the indirect Γ–K one, which is the same as that of the annealed CVD-grown 2D MoS2. Conversely, the bandgap of the vulcanized CVD-grown 2D MoS2 is the direct K–K one since vulcanization makes the CVD-grown 2D MoS2 transit from the oxygen-doped state to the pure state. Accordingly, the bandgap value of the vulcanized CVD-grown 2D MoS2 is the same as that of pure 2D MoS2, which is higher than that of the CVD-grown 2D MoS2. The bandgap value of the annealed CVD-grown 2D MoS2 is the lowest as it has the highest O-doping concentration. Fig. 8b shows the conduction band minimum, valence band maximum, bandgap and Fermi level (EFi) of the CVD-grown 2D MoS2 and the CVD-grown 2D MoS2 after annealing treatment and vulcanization.
Under the same conditions, the carrier concentration of a semiconductor increases with increasing doping concentration and decreasing bandgap while the conductivity increases with increasing carrier concentration. The O-doping concentration in the vulcanized CVD-grown 2D MoS2 is 0 and its bandgap is the highest, so its dark and light currents are significantly smaller than those of the CVD-grown 2D MoS2. Compared with the CVD-grown 2D MoS2, the annealed CVD-grown 2D MoS2 has a higher O-doping concentration, a lower bandgap and a higher carrier concentration, which is the reason why its dark and light currents are higher than those of the CVD-grown 2D MoS2. Therefore, the difference in the photoelectric performance of the CVD-grown 2D MoS2 and CVD-grown 2D MoS2 after annealing treatment and vulcanization is attributed to the difference in the O-doping concentration. Similarly, the difference in the reported photoelectric performances30–36 should also result from the difference in doping concentration in the CVD-grown 2D MoS2.
Heating was carried out in a stepwise manner. First, MoO3 was heated to 873 K within 30 min to trigger evaporation. Then, the temperature of MoO3 was slowly raised up to the reaction temperature, 1083 K, at a rate of 5 K min; in the meantime, sulfur powder was heated for evaporation. The sulfur vapor was carried by the argon gas flow to approach and react with the molybdenum source. The reaction proceeded for 30 min under these conditions to allow a sufficient reduction of the sulfur vapor by the MoO3 vapor. After the reaction ceased, argon injection was kept to allow the samples to naturally cool down to the ambient temperature and then the sample was taken out of the tube furnace.
During transferring, to start the surface of the PDMS-2 with nothing bound was attached to the glass slide, which was then placed into the holder of the material displacement table. The optical microscope focused on the substrate of a flexible PDMS-1 substrate with surface-patterned metal electrodes; the target substrate displacement table was adjusted to search for the position of the target metal electrode for specific transferring. Moreover, the microscope focused on the surface of the MoS2 crystal film adhering to the surface of PDMS-2, while adjusting the material displacement table to search for the specific MoS2 crystal film requiring transfer. After this, the positions of the metal electrode and the MoS2 crystal film were aligned. The holder was slowly lowered down, during which the focal distance was constantly adjusted in order to precisely adjust the relative positions of the metal electrode and MoS2 crystal film, ensuring that the target transfer material and target transfer position were always within the same vertical plane until they coincided with each other. At last, the glass slide was retrieved from the holder, on which, from top to bottom, lie the support layer PDMS-2, MoS2 crystal film and flexible substrate PDMS-1.
Footnote |
† Electronic supplementary information (ESI) available: The accompanying supporting information includes the dependence of PO2 (the partial pressure of O2) on tA (the subsequent aeration time from the moment when the inert gas was injected to 1 atm), a schematic diagram of the preparation of metal electrodes on a flexible PDMS substrate, and the operating procedure for transferring MoS2 onto the flexible substrate. See DOI: 10.1039/d0ra10302k |
This journal is © The Royal Society of Chemistry 2021 |