You are on page 1of 8

www.acsami.

org Research Article

Asymmetric 2D MoS2 for Scalable and High-Performance


Piezoelectric Sensors
Wonbong Choi,* Junyoung Kim, Eunho Lee, Gayatri Mehta, and Vish Prasad
Cite This: ACS Appl. Mater. Interfaces 2021, 13, 13596−13603 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Piezoelectricity in two-dimensional (2D) transition-metal dichalcogenides


Downloaded via UNIV PRESBITERIANA MACKENZIE on August 11, 2021 at 18:59:49 (UTC).

(TMDs) has attracted significant attention due to their unique crystal structure and the
lack of inversion centers when the bulk TMDs thin down to monolayers. Although the
piezoelectric effect in atomic-thickness TMDs has been reported earlier, they are
exfoliated 2D TMDs and are therefore not scalable. Here, we demonstrate a superior
piezoelectric effect from large-scale sputtered, asymmetric 2D MoS2 using meticulous
defect engineering based on the thermal-solvent annealing of the MoS2 layer. This yields
an output peak current and voltage of 20 pA and 700 mV (after annealing at 450 °C),
respectively, which is the highest piezoelectric strength ever reported in 2D MoS2.
Indeed, the piezoelectric strength increases with the defect density (sulfur vacancies),
which, in turn, increases with the annealing temperature at least up to 450 °C. Moreover,
our piezoelectric MoS2 device array shows an exceptional piezoelectric sensitivity of 262
mV/kPa with a high level of uniformity and excellent performance under ambient
conditions. A detailed study of the sulfur vacancy-dependent property and its resultant asymmetric structure-induced piezoelectricity
is reported. The proposed approach is scalable and can produce advanced materials for flexible piezoelectric devices to be used in
emerging bioinspired robotics and biomedical applications.
KEYWORDS: TMDs, MoS2, piezoelectric sensors, asymmetric, defects

1. INTRODUCTION coefficient. This membrane is packed in a hexagonal lattice


The piezoelectricity of thin films, nanowires, and bulk crystals with one atomic layer of Mo sandwiched between two S layers.
has been extensively studied for its applications in sensors, Two S atoms asymmetrically order this unit cell between Mo
electronics, nanogenerators, and energy conversion.1,2 Con- atoms, such that an external electric field is generated from the
sequently, two-dimensional (2D) materials have attracted S to the Mo site in the hexagonal lattice. This enables the
considerable interest due to their high crystallinity as well as generation of internal piezoelectricity of the 2D MoS2
ability to withstand tremendous strain while presenting their membrane when it is strained.11,12 On the other hand, the
piezoelectric properties at the atomic level. Indeed, two- piezoelectricity from alternating layers is canceled out when
dimensional (2D) transition-metal dichalcogenides (TMDs) the membranes are even-layered. Evidently, the piezoelectric
effect of 2D MoS2 is a number of layer-dependent.
have been the subject of intense research due to their unique
As is well established, a conventional flexible piezoelectric
structural features of thin atomic layers held together via van
sensor measures the dipole moment to obtain its strain-
der Waals (vdW) interactions.3,4 Any qualitative change in
induced piezoelectric field.12−14 The internal electric dipoles of
their atomic layer of 2D TMDs can affect their physical and
piezoelectric materials are coupled to the material lattice;
chemical properties because of the large surface-area-to-
therefore, any changes in the lattice, such as point defects,
thickness ratio.5−8
impurities, and vacancies, can result in a change in the
Interestingly, bulk TMDs, including MoS2, do not possess a
spontaneous polarization.15−17 Prezhdo et al. have observed
piezoelectric effect due to the opposite orientations of adjacent
that the defects introduce charge traps and accelerate
atomic layers, centrosymmetric in their bulk. However, when
electron−hole recombination by a factor of 1.7 since the
the same 2D TMDs are thinned down to a few layers,
inversion symmetry disappears and the piezoelectric effect
appears.8−10 In free-standing MoS2 membranes, it has been Received: January 12, 2021
found that piezoelectricity exists when there are odd number Accepted: March 3, 2021
layers in the 2D crystal because the inversion symmetry breaks Published: March 12, 2021
down.11 The monolayer of MoS2 has a D3h crystalline
symmetry, which can adequately describe the anisotropic
electromechanical coupling by the single piezoelectric

© 2021 American Chemical Society https://dx.doi.org/10.1021/acsami.1c00650


13596 ACS Appl. Mater. Interfaces 2021, 13, 13596−13603
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 1. Thermal-solvent annealing process and characterization of MoS2. (a) Schematic illustration for the thermal-solvent annealing process
(top diagram) and its microstructure change. (b) Cross-sectional high-resolution transmission electron microscope (HRTEM) image of MoS2,
which confirms three layers of MoS2. (c) Intensity profile of HRTEM analysis that confirms the interlayer spacing of the layer as 0.65 nm. (d)
atomic force microscope (AFM) analysis of the MoS2 device; thickness of the device is 1.32 nm. (e) Raman spectra of pristine (magenta) and
thermal-solvent annealed MoS2 at 350 °C (blue), 400 °C (black), and 450 °C (red). (f) XPS analysis of MoS2, pristine (bottom) and annealed at
450 °C (top); the nonstoichiometric area fraction of the MoxSy peak appeared due to the sulfur-deficient MoS2. (g) Calculated interdefect areal
fractions of the MoxSy; the areal fraction gradually increases from 0 to 31.51% for 450-MoS2.

sulfur vacancy in MoS2 creates a deep electron trap and a redistribution occurs in conduction- or valence bands, which
shallow hole trap that couple to the free hole.18 Moreover, a can cause an asymmetrical modulation of carrier transport and
sulfur vacancy with unsaturated chemical bonds at the Mo higher piezoelectric response from the p-type MoS2 monolayer.
atoms attempts to interact with the adjacent S atoms, resulting It has also been shown that the intrinsic p- or n-type dopants
in a perturbed MoS2 symmetry. can be introduced by generating structural defects during the
As reported by Wu et al., the strain-induced polarization sample fabrication process.24,25 Indeed, these defects can be
charges in MoS2 can modulate charge carrier transport at the presented as the chalcogenide vacancies or surface in the
MoS2 interface and enhance its strain-sensing property. The structure, resulting in the local nonstoichiometry as well as
undoped MoS2 possesses a relatively finite intrinsic charge varying doping regions.26−28 This inhomogeneous electronic
carrier density with a limited piezoelectric response.12 landscape plays a critical role in electron transport in these
Therefore, if the distribution of free carriers in MoS2 is materials.
modified, an additional polarization effect may arise by the Interestingly, by creating defects in the graphene nanoribbon
asymmetrical charge distribution and can result in higher of the proper symmetry, Sharma et al. have found that even
piezoelectricity.11,19,20 The additional doping in the MoS2 centrosymmetric graphene nanoribbon can act as a piezo-
structure can further improve the sensitivity of strain-induced electric material.29 The implication, therefore, is that the
polarization in the MoS2 membrane and enhance its precise control of the defects in the 2D MoS2 structure can
piezoelectric effect.9,10,21,22 Nakamura et al. have reported lead to fine-tuning of the piezoelectrical property. The
that based on the density functional theory (DFT), a p-type difference in atomic bonding lengths gives rise to inequivalent
MoS2 monolayer possesses high longitudinal gauge factors of charge distribution, resulting in a broken reflection symmetry
94.2 (armchair) and 87.4 (zigzag) at room temperature,23 along the vertical direction and vertical piezoelectric polar-
which are 800 times higher than those of the n-type MoS2 ization. In the same manner, sulfur-deficient MoS2 exhibits
monolayer. Moreover, upon uniaxial tensile strain, the carrier similar characteristics due to the lack of inversion symmetry. In
13597 https://dx.doi.org/10.1021/acsami.1c00650
ACS Appl. Mater. Interfaces 2021, 13, 13596−13603
ACS Applied Materials & Interfaces www.acsami.org Research Article

this regard, conventional etching methods such as plasma or corresponding to a dispersive, polar, and hydrogen-bonding
electron irradiation etching suffer from severe limitations due parameter, respectively36,37
to their complexity and limited controllability. To control
precise defects over large-scale 2D TMDs, a new technique δ T = (δ D2 + δ P2 + δ H2)1/2
needed to be devised.26,30−32 The close predicted values of a total parameter, δT, for MoS2,
Recently, the authors have developed a facile and precise 18.54−24.02, and toluene, 18.16, using HSP calculations
defect-engineering process, using a thermal-solvent etching suggest that the sulfur can be efficiently etched on the MoS2
method, to fabricate large-scale asymmetrically engineered 2D surface using toluene. The full parameter values for MoS2 and
WS2 piezoelectric structures/devices.33 These 2D WS2-based toluene are summarized in Table S1.
devices have demonstrated a 3-fold higher piezoelectric A cross-sectional high-resolution transmission electron
response voltage (96.74 mV) than the pristine WS2 under a microscope (HRTEM) was used to determine the number of
3 kPa compression. In this paper, we extend the thermal- layers of MoS2 as well as the interlayer distance to confirm its
solvent method to synthesize an asymmetrical 2D MoS2. This uniformity at the atomic scale. As shown in Figure 1b, the
defect-engineered MoS2 shows an exceptional piezoelectric trilayered MoS2 was uniformly synthesized onto the SiO2/Si
output of current and voltage of 20 pA and 700 mV, substrate. Moreover, its intensity profile confirms that the
respectively, which is 20 times higher than that of the pristine interlayer distance was 0.65 nm, as shown in Figure 1c.
MoS2 device and much higher than that of the WS2-based Additionally, the thickness of the MoS2 device was measured
devices.33 Furthermore, it shows excellent uniformity and using an atomic force microscope (AFM), as shown in Figure
piezoelectric sensitivity, 262 mV/kPa, in an ambient 1d. The pristine MoS2 possessed a thickness of 1.32 nm, which
atmosphere. Note that unlike plasma or irradiation etching after thermal-solvent annealing did not change in any
techniques, the thermal-solvent etching method uses a significant manner.
temperature-controlled organic solvent vapor to etch the sulfur Raman spectroscopy was used to determine the sulfur-
with atomic precision and with no damage to the sample. deficient MoS2 crystal structure after thermal-solvent anneal-
Moreover, the produced sulfur vacancies lead to excessive ing. Sulfur deficiencies created during this process lead to the
holes in MoS2, rendering strong piezoelectric properties. formation of defects in the MoS2 structure. The two-vibration
Consequently, we have established a scalable process for the mode difference, Δk, was linearly increased from 20.47 to
synthesis of precise, defect-engineered 2D MoS2 that can 23.95 cm−1, indicating a systematic decrease in the out-of-
provide a comprehensive methodology/guidance for the plane binding energy due to the generation of defects.38,39 The
design of future flexible piezoelectronics of high capacities. density functional theory analysis suggests that the sulfur-
deficient structure weakens the restoring force constant and
2. RESULTS AND DISCUSSION allows the static Mo atom to vibrate out-of-plane.38,39 This
A large-scale polycrystalline MoS2 was grown by the two-step behavior was clearly observed in the present case, where the
process developed by Choi and co-workers,3,4 the details of out-of-plane phonon vibration mode, A1g, was blue-shifted, as
which can be found in the Experimental Section. Essentially, shown in Figure 1e. The observation further supports the
Mo was sputtered for 2 s on top of the bottom electrode, Ti/ hypothesis that the generated defects in the present samples
Au, followed by its sulfurization at 600 °C in a low-pressure are sulfur vacancies, and their density can be increased with the
chemical vapor deposition system. A schematic illustration of annealing temperature to generate more asymmetry in the
the two-step process and its scalability is described in Figure MoS2 structure.
S1. In addition, the surface film quality and scalable uniformity
By devising this process, we have established a novel were measured using Raman mapping and XPS-EDX. The two
thermal-solvent etching method to synthesize the non- major phonon vibration mode differences, Δk, for pristine,
stoichiometric 2D MoS2 by controlling the sulfur defect 350-MoS2, 400-MoS2, and 450-MoS2 were mapped, as shown
density.33 An illustration of the thermal-solvent etching process in Figure S3. The Mo and S distributions of five randomly
is presented in Figure 1a. Briefly, the vaporized organic solvent selected different zones were mapped using XPS-EDX to
partially dissolves the sulfur of MoS2 to create sulfur defects on confirm the scalable uniformity of the film surface, as shown in
MoS2. This process was repeated at three different temper- Figure S4. This Raman mapping and XPS-EDX analysis
atures, 350, 400, and 450 °C, to vary the etched sulfur confirm that the thermal-solvent annealing process gradually
concentration. The samples were termed as 350-MoS2, 400- etches the sulfur while it maintains its scalable uniform film
MoS2, and 450-MoS2 corresponding hereafter to different surface.
annealing temperatures used. An X-ray photoelectron spec- The details of sulfur vacancies on electronic and chemical
troscopy-energy dispersive spectroscopy (XPS-EDX) mapping states of MoS2 were further characterized by X-ray photo-
confirmed the uniform film surface before and after the electron spectroscopy (XPS). As compared to other vacancies,
thermal-solvent annealing, as shown in Figure S2. sulfur vacancies have low formation energy, and their
The usefulness of a solvent for etching TMD materials can formation is more favored during the thermal-solvent
be determined by the Hansen solubility parameter (HSP), as annealing.26,27 As shown in Figure 1f, the two major Mo 3d
presented in the previous paper on WS2-based piezoelectric peaks were observed for the pristine and 450-MoS2 samples.
sensors.33 A nonpolar circularity feature and high density of π- The most intense binding energy peaks were observed in the
bonds of the solvent molecule facilitate the dissolution of a pristine sample at 232.06 (Mo 3d3/2) and 228.95 eV (Mo
cyclic crown-shaped sulfur of the MoS2 surface.34 A semi- 3d5/2), corresponding to the Mo4+ state.1,2 In contrast, the
empirical correlation explaining the dissolution behavior of thermal-solvent-annealed-MoS2 sample shows the blue-shifted
sulfur in etchant can be expressed using the Hansen solubility Mo4+ state at 232.13 (Mo 3d3/2) and 229.07 eV (Mo 3d5/2)
parameter (HSP).35 Each solubility parameter represents the corresponding to the Mo4+ state.39 In addition, nonstoichio-
character of a solvent or solute by δD, δP, and δH, metric MoxSy peaks with lower binding energy in the Mo 3d
13598 https://dx.doi.org/10.1021/acsami.1c00650
ACS Appl. Mater. Interfaces 2021, 13, 13596−13603
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 2. Electrical characterization of MoS2. (a) Relative potential peak shift in MoS2, pristine (black) and thermal-solvent annealed at 350 °C
(red), 400 °C (blue), and 450 °C (magenta). (b) KPFM image of the pristine MoS2 (left) and thermal-solvent-annealed MoS2 (right). (c)
Illustration of the field-effect transistor (FET) device and a top-view optical image of the FET device. (d) Threshold voltage shift of MoS2 devices;
the initial threshold voltage of 1.2 V was shifted to 12 V, which demonstrates the enhanced p-type behavior after thermal-solvent annealing due to
the deep electron trap. (e) Summary plot of the threshold voltage and electron mobility of the MoS2 device.

Figure 3. Piezoelectric characterization of MoS2. (a, d) Output voltage of the pristine MoS2, 350-MoS2, 400-MoS2, and 450-MoS2 samples; the
thermal-solvent-annealed-MoS2 sample shows 35 times higher output voltage and 20 times higher output current than the pristine device. (b, e)
Output voltage and current data of 450-MoS2. (c) Output voltage of pristine and 10 different samples annealed between 350 and 450 °C, indicating
that the systematically controlled sulfur vacancy has an output voltage enhancement. (f) Output current of pristine and 10 samples annealed
between 350 and 450 °C indicates that the systematically controlled sulfur vacancy has an output current enhancement.

were observed near 230.81 and 227.98 eV, representing 31.51%, as shown in Figure 1g. Likewise, the ratio of Mo and S
intermediate defect states.40−43 significantly decreased from 1:2 to 1:1.45 with increased
The areal fraction of MoxSy indicates that the density of annealing temperature. The ratio of Mo vs S, i.e., S defect
sulfur vacancies increases with the annealing temperature. A concentration, with annealing temperature variation, is
detailed XPS analysis of different temperature samples, 350- summarized in Table S2. Their Raman and XPS analyses
MoS2 and 400-MoS2, is summarized in Figure S5. As the confirm that the sulfur defects can be efficiently generated to
annealing temperature was increased, the areal fraction of create asymmetry in the MoS2 structure, which can then be
nonstoichiometric MoxSy peaks gradually increased from 0 to used for piezoelectric applications. In HRTEM images (Figure
13599 https://dx.doi.org/10.1021/acsami.1c00650
ACS Appl. Mater. Interfaces 2021, 13, 13596−13603
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 4. Large-area array and stability demonstration. (a) Schematic illustration of the piezoelectricity measurement setup and tested device. (b)
Durability test data from the 450-MoS2 device showing excellent stability over 1000 s. (c) Voltage output map measured on 5 cm × 5 cm. MoS2
piezoelectric sensor showing the uniform distribution of piezoelectricity with a value of 0.659 ± 0.17 V under a compression pressure at 3 kPa. (d)
Piezoelectric sensitivity of pristine (black) MoS2 and thermal-solvent-annealed at 350 °C (red), 400 °C (blue), and 450 °C (pink); the sensitivity
increases gradually from 7.72 to 262 mV/kPa. (e) Stability of the MoS2 piezoelectric sensor; our device shows good stability in an ambient
atmosphere of 23 ± 1 °C and 33.4 ± 3% humidities for a month.

S6), sulfur defects are clearly seen from the thermal-solvent- shifted from 0.8 to 12.42 V and the drain−source current (Ids)
annealed sample, and the atomic structure with defects is not decreased from 1.298 to 0.017 nA. Also, the mobility decreased
centrosymmetric. Therefore, it can be concluded that the from 1.36 × 10−3 to 2.24 × 10−5 cm2/(V s). Evidently, the
defects generated by the present thermal-solvent annealing annealed-MoS2 samples exhibit a p-type characteristic with an
method result in asymmetric MoS2 structures. increase in positive charges in the MoS2 crystal structure.24,46,47
To understand the electron trapping during the thermal A behavior as demonstrated by the threshold voltage, Vth, in
release process and differentiate the role of the defects in active Figure 2e was also followed by the carrier density, n2D, which
MoS2, the surface potential and carrier density before and after was calculated to compare the doping levels of the pristine and
the thermal-solvent annealing were analyzed using Kelvin thermal-solvent-annealed MoS2 using the following equation47
probe force microscopy (KPFM). The relative surface
potential of pristine MoS2 was set to zero to compare the n2D = (IDSL)/(qWVDSμ)
surface potential with thermal-solvent-annealed MoS2. As a
result, the relative potential of 450-MoS2 decreased by 0.175 where q is the electron charge, L is the channel length, W is the
eV, indicating a Fermi level shift toward the intrinsic Fermi channel width, and μ is the mobility at VDS = −1 V. The
level, as shown in Figure 2a. The binding energy shift observed detailed calculation is provided in the Supporting data (Figure
in XPS results in Figure 1g corresponds to the Fermi level of S8). The carrier (electron) density of the pristine MoS2 device
the system, where the shift of 0.15 eV to lower binding energy was 1.83 × 1011 cm−2. However, the carrier density of the
can be attributed to the shift in the Fermi level.44,45 A KPFM annealed device decreased to 6.69 × 1010 cm−2, about a 2.73
image of the pristine and 450-MoS2, as shown in Figure 2b, times reduction after the annealing. Note that the hole density
confirms the uniformity of the defect-engineered MoS2 surface did not decrease as much as the electron density during the
and its surface potential change with thermal-solvent annealing annealing process.
temperature. The piezoelectric performance of the pristine sample and
Furthermore, we fabricated the MoS2 field-effect transistor thermal-solvent-annealed samples is shown in Figure 3. The
(FET) device to characterize the carrier density before and piezoelectric output voltage responses of pristine and 450-
after the thermal-solvent annealing, as shown in Figure 2c. Ids− MoS2 were measured under an applied compressive force of 3
Vgs curves of the pristine and annealed-MoS2 film-based FET kPa. Interestingly, the piezoelectric output voltage and current
are p-type, as presented in Figure 2d,e. They revealed that after increased with thermal-solvent annealing temperature (Figure
annealing, the threshold voltage (Vth) of the MoS2 device 3a,d). It is noted that the pristine MoS2 device generated an
13600 https://dx.doi.org/10.1021/acsami.1c00650
ACS Appl. Mater. Interfaces 2021, 13, 13596−13603
ACS Applied Materials & Interfaces www.acsami.org Research Article

output voltage of around 40 mV. Notably, 450-MoS2 generated sensitivity than that by the pristine MoS2 regardless of the
an output voltage of around 700 mV, which is over 18 times pressure, which makes it extremely suitable for its use as an
higher than that of the pristine MoS2 device and the highest effective piezoelectric sensor in bioinspired robots and
output voltage reported from 2D MoS2 thus far. In the case of biomedical applications.
the output current, the 450-MoS2 device generated 20 pA, Finally, all of the MoS2 devices were exposed to the ambient
which is about 20 times higher than that of the pristine MoS2 temperature for 1 month to investigate their stability in normal
device, as shown in Figure 3d. It is evident that the level of the operating conditions for a long time; the temperature and
holes is similar in all types of MoS2 devices, as shown in Figure humidity were maintained at 23 ± 1 °C and 33.4 ± 3%,
2c. However, an electron density in 450-MoS2 was significantly respectively. Indeed, all of the devices showed environmental
more suppressed than that of any other types of MoS2 devices, stability in the atmosphere even after a month within a 5%
which resulted in limiting the charge carrier recombination. deviation, which is comparable to other reported values.22
This allows excessive holes to behave as free carriers for These results demonstrate that the present thermal-solvent
current transport in a device consistent with the previously annealing process can be an effective method to engineer
reported data.18,25,46 defect concentrations in 2D MoS2 layers, thereby facilitating
Indeed, Inoue et al. have found that the MoS2 device the fabrication of large-scale stable MoS2 piezoelectric sensors
behaves as a relatively metallic-like feature due to the sulfur with excellent performance.
defects, as predicted by the DFT calculations.48 By removing
sulfur atoms from a MoS2 surface, electronic states are formed 3. CONCLUSIONS
in a mid-gap state assigned to the Mo d-orbital under the sulfur We have demonstrated that the sulfur-deficient 2D MoS2,
vacancy.49 Therefore, a high density of surface vacancy is found synthesized using a thermal-solvent etching method, can
to increase the electronic density of states in the band gap, produce a precisely defect-engineered structure with a p-type
resulting in a metallic feature. Moreover, these defects provide characteristic and provide a synergistic effect toward piezo-
low-resistance conduction paths in MoS2-based nanodevi- electricity. This can lead to the piezoelectric output current
ces.44,48 The difference in atomic bonding lengths gives rise to and voltage of the defect-controlled MoS2 device by more than
inequivalent charge distribution, resulting in a broken 20 times compared to a pristine device, a phenomenon
reflection symmetry in vertical piezoelectric polarization.50 In attributed to the formation of the local asymmetric structure by
the same manner, sulfur-deficient MoS2 exhibits piezoelectric the created defects and abundance of p-type dopants within
characteristics due to the lack of inversion symmetry. A the MoS2 structure. Our results also demonstrate that the
summary of the piezoelectric output voltage and current is piezoelectricity increases with the defect density, increasing
presented in Figure 3c,f, respectively; the piezoelectric with the thermal-solvent annealing temperature within the
performance of other samples is described in Figure S7. range of present experiments (≤450 °C). Indeed, the MoS2
From these experiments, we can conclude that the precise device created by us shows an outstanding piezoelectric
defect control in 2D MoS2 enhances its piezoelectric strength sensitivity of 262 mV/kPa while retaining its performance in an
by creating local asymmetry and increasing p-type dopants ambient atmosphere for a long time, over a month in the
within the MoS2 structure. Here, the piezoelectric coefficient present experiment. This work also establishes that the
was calculated using the method of Lefki and Dormans51 thermal-solvent etching method is a facile and effective
Q method to engineer piezoelectricity in atomic-thickness MoS2
d33 =
F
= ∫ I dt / F layers by precisely creating an asymmetrical charge distribu-
tion.
The piezoelectric coefficient, d33, was calculated to be 16.6 pC/
N (pristine MoS2), 49.8 pC/N (350-MoS2), 83 pC/N (400- 4. EXPERIMENTAL SECTION
MoS2), and 332 pC/N (450-MoS2); evidently, piezoelectric 4.1. Synthesis of MoS2. For piezoelectric device fabrication, a
coefficients increase with the thermal-solvent annealing bottom electrode was patterned and sputtered on a SiO2/Si substrate
temperature. Further studies will be carried out using by using titanium (99.99%, Plasmaterials) and gold (99.99%,
piezoresponse force microscopy (PFM) to visualize the Plasmaterials) targets. Then, the 2D MoS2 film was synthesized via
electrochemical motion of the dipole moment to provide an a two-step synthesis method.1,2 First, the bottom electrode-patterned
insight into switchable electric polarization.52 substrate was transferred to a radiofrequency magnetron plasma
Figure 4a shows the fabricated scalable MoS2 piezoelectric sputtering system. Molybdenum (Mo) (99.99%, Plasmaterials) was
sensor and its schematic with the measurement setup. A sputtered under Ar plasma (99.99% Airgas) at a deposition of 60 W
for 2 s. Then, the Mo-sputtered sample was transferred to the low
representative voltage durability data from the 450-MoS2 pressure chemical vapor deposition (LPCVD) chamber for sulfuriza-
device shows exceptional reliability over 1000 s, as shown in tion at 600 °C for 1 h. Finally, a top electrode, Cu (99.99%
Figure 4b. A large-area, 5 cm × 5 cm piezoelectric mapping Plasmaterials), was patterned and deposited on top of the MoS2 layer.
taken from the fabricated MoS 2 piezoelectric sensor 4.2. Thermal-Solvent Annealing Method. The defect-engi-
undoubtedly shows the outstanding film uniformity and neered MoS2 was processed via LPCVD with organic solvent vapor.
sensitivity on the entire surface area, as shown in Figure 4c. First, the as-deposited MoS2 film on top of the quartz sample was
The output voltage average, measured at 3 kPa, was 0.659 ± placed in the middle of the LPCVD furnace connected to an attached
0.17 V. The sensitivity of the pristine and thermal-solvent- supplementary chamber filled with toluene. Second, once the vacuum
annealed samples with respect to the pressure variation from 1 level and the annealing temperature were reached, the supplementary
chamber was opened for 7 min to allow the toluene vapors to etch the
to 5 kPa was also investigated and is presented in Figure 4d. sulfur of the MoS2 surface. This process was repeated at three
The pristine MoS2 device had 7.72 mV/kPa of sensitivity, and different temperatures to vary the concentration of the etched sulfur:
this gradually increased to 262 mV/kPa for the 450 °C 350, 400, and 450 °C.
annealed sample. This sensitivity data shows that our thermal- 4.3. Field-Effect Transistor (FET) Fabrication. An electrical
solvent-annealed device can generate 34 times higher property of defect-engineered MoS2 was characterized by the MoS2-

13601 https://dx.doi.org/10.1021/acsami.1c00650
ACS Appl. Mater. Interfaces 2021, 13, 13596−13603
ACS Applied Materials & Interfaces www.acsami.org Research Article

FET device. First, a 300 nm thickness of SiO2/Si was prepared and Author Contributions
cleaned. A few-layered MoS2 was prepared by the aforementioned All of the work was designed and performed under the
method on top of SiO2/Si, followed by the thermal-solvent annealing supervision of W.C. J.K. performed material synthesis and
process. Second, a patterned shadow mask was placed on top of the characterizations, and E.L performed device characterization.
MoS2 layer, and then gold (99.99%, Plasmaterials) was sputtered and
used as an electrode. W.C. and V.P. wrote the manuscript. G.M. analyzed data and
4.4. Microscopic and Electrical Characterization. The edited the manuscript. All authors have discussed the results,
crystalline property of the MoS2 was analyzed by Raman spectroscopy read the manuscript, and agreed with its content.
(WiTec Alpha 300R) with an excitation wavelength of 532 nm. The Notes
thickness of the MoS2 film was confirmed by atomic force microscopy The authors declare no competing financial interest.


(Parks NX-10). The structural composition of the MoS2 was
determined by X-ray photoelectron spectroscopy (PHI 5000
VersaProbe Scanning XPS) at 15 KV under an ultrahigh vacuum ACKNOWLEDGMENTS
(10−8 Torr). The electrochemical properties were analyzed by a This work was supported by the National Science Foundation
Kelvin probe force microscope (Bruker MultiMode 8). The (grant No. CBET-1936255).


piezoelectricity was measured using an automated tapping machine,
ESM-303 (Mark-10); the compressive force was applied vertically REFERENCES
from 1 to 5 kPa. ESM-303 was connected directly with the computer
and precision source/measure unit electrometer (Agilent B2912A) to (1) Kingon, A. I.; Srinivasan, S. Lead Zirconate Titanate Thin Films
read the piezoelectric responses. TEM analysis was performed using a Directly on Copper Electrodes for Ferroelectric, Dielectric and
JSM-7800F Prime instrument (JEOL Ltd.) at an operational voltage Piezoelectric Applications. Nat. Mater. 2005, 4, 233−237.
of 200 kV, fitted with a Cs corrector (CEOS GmbH) FEG-STEM/ (2) Wang, Z. L.; Song, J. Piezoelectric Nanogenerators Based on
TEM unit. The images were reconstructed using a Gatan Digital Zinc Oxide Nanowire Arrays. Science 2006, 312, 242−246.
Micrograph computational package. (3) Choudhary, N.; Park, J.; Hwang, J. Y.; Choi, W. Growth of


Large-Scale and Thickness-Modulated MoS2 Nanosheets. ACS Appl.
ASSOCIATED CONTENT Mater. Interfaces 2014, 6, 21215−21222.
(4) Choi, W.; Choudhary, N.; Park, J.; Akinwande, D.; Lee, Y.; et al.
*
sı Supporting Information Recent development of 2D materials and their applications. Mater.
The Supporting Information is available free of charge at Today 2017, 20, 116−130.
https://pubs.acs.org/doi/10.1021/acsami.1c00650. (5) Iannaccone, G.; Bonaccorso, F.; Colombo, L.; Fiori, G.
Quantum Engineering of Transistors Based on 2D Materials
Schematic illustration of our two-step process of MoS2 Heterostructures. Nat. Nanotechnol. 2018, 13, 183−191.
synthesis (Figure S1); XPS-EDX mapping confirmed the (6) Yang, H.; Kim, S. W.; Chhowalla, M.; Lee, Y. H. Structural and
uniform film surface (Figure S2); Raman mapping Quantum-State Phase Transition in van Der Waals Layered Materials.
analysis of peak differences (Figure S3); XPS-EDX Nat. Phys. 2017, 13, 931−937.
analysis of Mo and S atomic distributions (Figure S4); (7) Xia, F.; Wang, H.; Xiao, D.; Dubey, M.; Ramasubramaniam, A.
XPS analysis of MoS2 (Figure S5); high-resolution TEM Two-Dimensional Material Nanophotonics. Nat. Photonics 2014, 8,
images (Figure S6); temperature-dependent piezo- 899−907.
electric output current−voltage (Figure S7); carrier (8) Ong, M. T.; Reed, E. J. In Engineered Piezoelectricity in Graphene
density calculations (Figure S8); Hansen solubility by Chemical Doping, Technical Proceedings of the 2012 NSTI
Nanotechnology Conference and Expo; NSTI-Nanotech 2012; 2012;
parameters (Table S1); and summarized Mo and S pp 161−164.
ratios (Table S2) (PDF) (9) Duerloo, K.-A. N. D.; Ong, M. T.; Reed, E. J. Intrinsic


Piezoelectricity in Two-Dimensional Materials. J. Phys. Chem. Lett.
AUTHOR INFORMATION 2012, 3, 2871−2876.
(10) Michel, K. H.; Verberck, B. Theory of Elastic and Piezoelectric
Corresponding Author Effects in Two-Dimensional Hexagonal Boron Nitride. Phys. Rev. B
Wonbong Choi − Department of Materials Science and 2009, 80, No. 224301.
Engineering and Department of Mechanical Engineering, (11) Zhu, H.; Wang, Y.; Xiao, J.; Liu, M.; Xiong, S.; Wong, Z. J.; Ye,
University of North Texas, Denton, Texas 76203, Unites Z.; Ye, Y.; Yin, X.; Zhang, X. Observation of Piezoelectricity in Free-
States; orcid.org/0000-0002-7896-7655; Standing Monolayer MoS2. Nat. Nanotechnol. 2015, 10, 151−155.
Email: Wonbong.choi@unt.edu (12) Wu, W.; Wang, L.; Li, Y.; Zhang, F.; Lin, L.; Niu, S.; Chenet,
D.; Zhang, X.; Hao, Y.; Heinz, T. F.; James, H.; Wang, Z. L.
Authors Piezoelectricity of Single-Atomic-Layer MoS2 for Energy Conversion
Junyoung Kim − Department of Materials Science and and Piezotronics. Nature 2014, 514, 470−474.
Engineering, University of North Texas, Denton, Texas (13) Kim, S. K.; Bhatia, R.; Kim, T. H.; Seol, D.; Kim, J. H.; Kim, H.;
Seung, W.; Kim, Y.; Lee, Y. H.; Kim, S. W. Directional Dependent
76203, Unites States; orcid.org/0000-0001-5023-4030 Piezoelectric Effect in CVD Grown Monolayer MoS2 for Flexible
Eunho Lee − Department of Mechanical Engineering, Piezoelectric Nanogenerators. Nano Energy 2016, 22, 483−489.
University of North Texas, Denton, Texas 76203, United (14) Zhou, Y.; Liu, W.; Huang, X.; Zhang, A.; Zhang, Y.; Wang, Z. L.
States; orcid.org/0000-0002-8564-6999 Theoretical Study on Two-Dimensional MoS2 Piezoelectric Nano-
Gayatri Mehta − Department of Electrical Engineering, generators. Nano Res. 2016, 9, 800−807.
University of North Texas, Denton, Texas 76203, United (15) Ren, X. Large Electric-Field-Induced Strain in Ferroelectric
States Crystals by Point-Defect-Mediated Reversible Domain Switching.
Vish Prasad − Department of Mechanical Engineering, Nat. Mater. 2004, 3, 91−94.
University of North Texas, Denton, Texas 76203, United (16) Apte, A.; Mozaffari, K.; Samghabadi, F. S.; Hachtel, J. A.;
States Chang, L.; Susarla, S.; Idrobo, J. C.; Moore, D. C.; Glavin, N. R.;
Litvinov, D.; Sharma, P.; Puthirath, A. B.; Ajayan, P. M. 2D Electrets
Complete contact information is available at: of Ultrathin MoO2 with Apparent Piezoelectricity. Adv. Mat. 2020, 32,
https://pubs.acs.org/10.1021/acsami.1c00650 No. 200006.

13602 https://dx.doi.org/10.1021/acsami.1c00650
ACS Appl. Mater. Interfaces 2021, 13, 13596−13603
ACS Applied Materials & Interfaces www.acsami.org Research Article

(17) Lohkämper, R.; Neumann, H.; Arlt, G. Internal Bias in Two-Dimensional Nanosheets Produced by Liquid Exfoliation of
Acceptor-Doped BaTiO3 Ceramics: Numerical Evaluation of Increase Layered Materials. Science 2011, 331, 568−571.
and Decrease. J. Appl. Phys. 1990, 68, 4220−4224. (37) Zhou, K.-G.; Mao, N.-N.; Wang, H.-X.; Peng, Y.; Zhang, H.-L.
(18) Li, L.; Long, R.; Bertolini, T.; Prezhdo, O. V. Sulfur Adatom A Mixed-Solvent Strategy for Efficient Exfoliation of Inorganic
and Vacancy Accelerate Charge Recombination in MoS2 but by Graphene Analogues. Angew. Chem., Int. Ed. 2011, 50, 10839−10842.
Different Mechanisms: Time-Domain Ab Initio Analysis. Nano Lett. (38) Wu, K.; Li, Z.; Tang, J.; Lv, X.; Wang, H.; Luo, R.; Liu, P.;
2017, 17, 7962−7967. Qian, L.; Zhang, S.; Yuan, S. Controllable Defects Implantation in
(19) Tsai, M. Y.; Tarasov, A.; Hesabi, Z. R.; Taghinejad, H.; MoS2 Grown by Chemical Vapor Deposition for Photoluminescence
Campbell, P. M.; Joiner, C. A.; Adibi, A.; Vogel, E. M. Flexible MoS2 Enhancement. Nano Res. 2018, 11, 4123−4132.
Field-Effect Transistors for Gate-Tunable Piezoresistive Strain (39) Parkin, W. M.; Balan, A.; Liang, L.; Das, P. M.; Lamparski, M.;
Sensors. ACS Appl. Mater. Interfaces 2015, 7, 12850−12855. Naylor, C. H.; Rodríguez-Manzo, J. A.; Johnson, A. T. C.; Meunier,
(20) Manzeli, S.; Allain, A.; Ghadimi, A.; Kis, A. Piezoresistivity and V.; Drndić, M. Raman Shifts in Electron-Irradiated Monolayer MoS2.
Strain-Induced Band Gap Tuning in Atomically Thin MoS2. Nano ACS Nano 2016, 10, 4134−4142.
Lett. 2015, 15, 5330−5335. (40) Bhoyate, S.; Kim, J.; Lee, E.; Park, B.; Lee, E.; Park, J.; Oh, S.
(21) Michel, K. H.; Verberck, B. Phonon Dispersions and H.; Kim, J.; Choi, W. Mixed Phase 2D Mo0.5W0.5S2 Alloy as Multi-
Piezoelectricity in Bulk and Multilayers of Hexagonal Boron Nitride. Functional Electrocatalyst for the High-Performance Cathode in Li-S
Phys. Rev. B 2011, 83, No. 115328. Batteries. J. Mater. Chem. A 2020, 8, 12436−12445.
(22) Zhang, D.; Yang, Z.; Li, P.; Pang; Xue, Q. Flexible self-powered (41) Kondekar, N. P.; Boebinger, M. G.; Woods, E. V.; McDowell,
high-performance ammonia sensor based on Au-decorated MoSe2 M. T. In Situ XPS Investigation of Transformations at Crystallo-
graphically Oriented MoS2 Interfaces. ACS Appl. Mater. Interfaces
nanoflowers driven by single layer MoS2-flake piezoelectric nano-
2017, 9, 32394−32404.
generator. Nano Energy 2019, 65, No. 103974.
(42) Baker, M. A.; Gilmore, R.; Lenardi, C.; Gissler, W. XPS
(23) Nakamura, K. First-Principles Simulation of Piezoresistivity of
Investigation of Preferential Sputtering of S from MoS2 and
Transition Metal Dichalcogenide Monolayers. Sens. Mater. 2018, 30,
Determination of MoSx Stoichiometry from Mo and S Peak Positions.
2073−2083. Appl. Surf. Sci. 1999, 150, 255−262.
(24) Yang, J.; Kawai, H.; Wong, C. P. Y.; Goh, K. E. J. Electrical (43) Bhoyate, S.; Mhin, S.; Jeon, J.; Park, K. R.; Kim, J.; Choi, W.
Doping Effect of Vacancies on Monolayer MoS2. J. Phys. Chem. C Stable and High-Energy-Density Zn Ion Rechargeable Batteries Based
2019, 123, 2933−2939. on MoS2 Coated Zn Anode. ACS Appl. Mater. Interfaces 2020, 12,
(25) Noh, J.-Y.; Kim, H.; Kim, Y.-S. Stability and Electronic 27249−27257.
Structures of Native Defects in Single-Layer MoS2. Phys. Rev. B 2014, (44) Fang, H.; Tosun, M.; Seol, G.; Chang, T. C.; Takei, K.; Guo, J.;
18, No. 205417. Javey, A. Degenerate N-Doping of Few-Layer Transition Metal
(26) Roy, S.; Choi, W.; Jeon, S.; Kim, D. H.; Kim, H.; Yun, S. J.; Lee, Dichalcogenides by Potassium. Nano Lett. 2013, 13, 1991−1995.
Y.; Lee, J.; Kim, Y. M.; Kim, J. Atomic Observation of Filling (45) Han, S. A.; Kim, T. H.; Kim, S. K.; Lee, K. H.; Park, H. J.; Lee,
Vacancies in Monolayer Transition Metal Sulfides by Chemically J. H.; Kim, S. W. Point-Defect-Passivated MoS2 Nanosheet-Based
Sourced Sulfur Atoms. Nano Lett. 2018, 18, 4523−4530. High Performance Piezoelectric Nanogenerator. Adv. Mater. 2018, 30,
(27) Zhou, W.; Zou, X.; Najmaei, S.; Liu, Z.; Shi, Y.; Kong, J.; Lou, No. 1800342.
J.; Ajayan, P. M.; Yakobson, B. I.; Idrobo, J. C. Intrinsic Structural (46) Chuang, S.; Battaglia, C.; Azcatl, A.; McDonnell, S.; Kang, J. S.;
Defects in Monolayer Molybdenum Disulfide. Nano Lett. 2013, 13, Yin, X.; Tosun, M.; Kapadia, R.; Fang, H.; Wallace, R. M.; Javey, A.
2615−2622. MoS2 P-Type Transistors and Diodes Enabled by High Work
(28) Lee, E.; Kim, J.; Bhoyate, S.; Cho, K.; Choi, W. Realizing Function MoOx Contacts. Nano Lett. 2014, 14, 1337−1342.
Scalable Two-Dimensional MoS2 Synaptic Devices for Neuromorphic (47) Sim, D. M.; Kim, M.; Yim, S.; Choi, M. J.; Choi, J.; Yoo, S.;
Computing. Chem. Mater. 2020, 32, 10447−10455. Jung, Y. S. Controlled Doping of Vacancy-Containing Few-Layer
(29) Chandratre, S.; Sharma, P. Coaxing Graphene to Be MoS2 via Highly Stable Thiol-Based Molecular Chemisorption. ACS
Piezoelectric. Appl. Phys. Lett. 2012, 100, No. 023114. Nano 2015, 9, 12115−12123.
(30) Dhall, R.; Neupane, M. R.; Wickramaratne, D.; Mecklenburg, (48) Inoue, A.; Komori, T.; Shudo, K. I. Atomic-Scale Structures and
M.; Li, Z.; Moore, C.; Lake, R. K.; Cronin, S. Direct Bandgap Electronic States of Defects on Ar+-Ion Irradiated MoS2. J. Electron
Transition in Many-Layer MoS2 by Plasma-Induced Layer Decou- Spectrosc. Relat. Phenom. 2013, 189, 11−18.
pling. Adv. Mater. 2015, 27, 1573−1578. (49) Kodama, N.; Hasegawa, T.; Tsuruoka, T.; Joachim, C.; Aono,
(31) Lee, Y.; Kim, J. Controlling Lattice Defects and Inter-Exciton M. Electronic State Formation by Surface Atom Removal on a MoS2
Interactions in Monolayer Transition Metal Dichalcogenides for Surface. Jpn. J. Appl. Phys. 2012, 51, No. 06FF07.
Efficient Light Emission. ACS Photonics 2018, 5, 4187−4194. (50) McDonnell, S.; Addou, R.; Buie, C.; Wallace, R. M.; Hinkle, C.
(32) Wang, S.; Lee, G.-D.; Lee, S.; Yoon, E.; Warner, J. H. Detailed L. Defect-Dominated Doping and Contact Resistance in MoS2. ACS
Atomic Reconstruction of Extended Line Defects in Monolayer MoS2. Nano 2014, 8, 2880−2888.
ACS Nano 2016, 10, 5419−5430. (51) Lefki, K.; Dormans, G. J. M. Measurement of Piezoelectric
(33) Kim, J.; Lee, E.; Mehta, G.; Chi, W. Stable and high- Coefficients of Ferroelectric Thin Films. J. Appl. Phys. 1994, 76, 1764.
performance piezoelectric sensor via CVD grown WS2. Nano- (52) Hong, S. Single frequency vertical piezoresponse force
technology 2020, 31, No. 445203. microscopy. J. Appl. Phys. 2021, 129, No. 051101.
(34) Ren, Y.; Shui, H.; Peng, C.; Liu, H.; Hu, Y. Solubility of
Elemental Sulfur in Pure Organic Solvents and Organic Solvent-Ionic
Liquid Mixtures from 293.15 to 353.15K. Fluid Phase Equilib. 2011,
312, 31−36.
(35) Hansen, C. M. Hansen Solubility Parameters: A User’s
Handbook: Second Edition, 2007; https://doi.org/10.1201/
9781420006834.
(36) Coleman, J. N.; Lotya, M.; O’Neill, A.; Bergin, S. D.; King, P. J.;
Khan, U.; Young, K.; Gaucher, A.; De, S.; Smith, R. J.; Shvets, I.V.;
Arora, S. K.; Stanton, G.; Kim, Hy.; Lee, K.; Kim, G. T.; Duesberg, G.
S.; Hallad, T.; Nicholls, R. J.; Perkins, J. M.; Grieveson, E. M.;
Theuwissen, K.; McComb, D. W.; Nellist, P. D.; Nicolosi, V.; et al.

13603 https://dx.doi.org/10.1021/acsami.1c00650
ACS Appl. Mater. Interfaces 2021, 13, 13596−13603

You might also like