ACS Publications. Most Trusted. Most Cited. Most Read
Peroxide-Induced Tuning of the Conductivity of Nanometer-Thick MoS2 Films for Solid-State Sensors
My Activity

Figure 1Loading Img
  • Open Access
Article

Peroxide-Induced Tuning of the Conductivity of Nanometer-Thick MoS2 Films for Solid-State Sensors
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

ACS Applied Nano Materials

Cite this: ACS Appl. Nano Mater. 2020, 3, 11, 10864–10877
Click to copy citationCitation copied!
https://doi.org/10.1021/acsanm.0c02135
Published November 6, 2020

Copyright © 2020 American Chemical Society. This publication is licensed under these Terms of Use.

Abstract

Click to copy section linkSection link copied!

Applications of molybdenum disulfide (MoS2) in energy storage devices, solar cells, electrocatalysts, and sensors require good electrical conductivity. However, neither of the current ways to prepare conductive MoS2 (lithium intercalation and hydrothermal processes) is easily amenable to scale-up. A possible alternative pathway is the modulation of the electronic properties of the semiconducting form of MoS2 through structural defects. Here, we report the preparation of nanoscale conductive MoS2 flakes by treating exfoliated 2H-MoS2 with dilute aqueous hydrogen peroxide at room temperature. Sheet resistance measurements as well as Raman and photoelectron spectroscopy reveal the partial formation of hydrogen molybdenum bronze (HxMoO3) and substoichiometric MoO3–y, which help tune the conductivity of the nanometer-scale thin films without impacting the sulfur-to-molybdenum ratio. We have cast the material into thin film networks to fabricate highly stable chemiresistive pH sensors. Our work introduces a straightforward and safe way of preparing a conductive form of MoS2 and its application as a low-cost solid-state sensor.

Copyright © 2020 American Chemical Society

Introduction

Click to copy section linkSection link copied!

Two-dimensional (2D) materials have garnered much interest over the past decade due to their demonstrated high performance in nanoelectronic devices. (1,2) The discovery of graphene opened up many opportunities to investigate and explore other 2D materials. Since then, a wide range of materials have been discovered or predicted, (3,4) with molybdenum disulfide (MoS2) being of particular interest. (5,6) The semiconducting phase of MoS2 (2H-MoS2) is one of the most commonly studied among the transition metal dichalcogenides. The direct band gap of monolayer 2H-MoS2 enables its use as a channel material for field-effect transistors (FET), where a high on/off current ratio is desired. (7,8) The presence of a band gap means that the use of 2H-MoS2 requires a relatively high voltage bias to operate. For applications in batteries, (9) supercapacitors, (10) electrocatalytic reactions, (11) and solar cells, (12) however, a substantially increased conductivity is essential in order to achieve practical current densities. The most common conductive form of MoS2 is metallic MoS2 (1T-MoS2) that has been prepared via the lithium intercalation process involving inert atmosphere processing and elaborate safety procedures. (13,14) This procedure requires elevated temperatures (∼100 °C) over long durations (more than 2 days). Another synthesis procedure of conductive MoS2 was reported by using a hydrothermal process (15) in an autoclave under careful temperature control. Again, this process requires sophisticated instrumentation. A safer, cost-effective, and more efficient process to yield conductive MoS2 would be beneficial in a number of intended applications.
MoS2 is most reactive at defect sites, which thus play a very important role in modulating the electrical properties of MoS2. (16−18) Sonication of MoS2 in an appropriate solvent creates many disordered structural defects. The most common defects in MoS2 are sulfur vacancies. (16) Both sulfur and molybdenum vacancies in MoS2 introduce gap states, (19) but sulfur vacancies are energetically favorable. Sulfur vacancies are not desirable for the purpose of device applications as they can result in Fermi level pinning and eventually deteriorate the device performance. (19,20) Various strategies have been proposed to passivate the sulfur vacancy defects using thiols, (21,22) molecular and atomic oxygen, (23) and organic superacids like bis(trifluoromethane)sulfonimide (TFSI) where protons (H+) act as a passivating agent. (24) While past research has mainly focused on ways to modify the defects for photoluminescence efficiency, (25,26) they could also be controlled to improve the conductivity of MoS2 as a safer alternative for applications in batteries, supercapacitors, solar cells, electrocatalysts, and sensors. So far, however, no reliable bulk synthesis of MoS2 with defect-induced conductivity has been reported.
Nanometer-scale thin films of conductive MoS2 can be used as active material for low-cost solid-state chemiresistive pH sensors. Paper-based colorimetric pH sensors are ubiquitous and easy to fabricate and use but are inaccurate and require human intervention. For online monitoring or laboratory applications, ion selective pH electrodes are most popular. However, they are expensive and also require frequent calibration due to their reliance on reference electrodes. Solid-state chemiresistive pH sensors can become a more reliable alternative to ion selective pH electrodes since they are inexpensive to fabricate and do not need any reference electrodes or membranes. In chemiresistive sensors, conductivity changes are observed based on direct interactions between the active material and the analyte. Even though chemiresistive pH sensors based on exfoliated graphene, carbon nanotubes, or graphitic materials are available, their sensing response is limited to <20%. (27−30)
In this work, we demonstrate a facile and safe way to prepare few-layer conductive MoS2 (c-MoS2) at ambient conditions using low concentrations of aqueous hydrogen peroxide (H2O2). The samples were characterized by four-probe conductivity and Hall measurements, X-ray photoelectron spectroscopy (XPS), and Raman spectroscopy to elucidate the origin of the enhanced conductivity. It was found that the interaction of H2O2 with 2H-MoS2 causes the formation of small amounts of hydrogen molybdenum bronze (HxMoO3) and substochiometric MoO3–y which are structurally intertwined with MoS2 and act as dopants. While the chemical and structural characteristics of MoS2 are retained in the process, the conductivity of the resulting c-MoS2 material is only about an order of magnitude lower than that of molybdenum bronze or 1T-MoS2. The c-MoS2 flakes can then be cast into percolation networks, forming conductive thin films with a sheet resistance up to 7 orders of magnitude lower than that of 2H-MoS2. We also explored the possibility of surface functionalization by different thiols and demonstrate a practical application of such modified conductive films in environmental sensing by fabricating a pH sensor that is highly stable and has a significant response. Because of the high conductivity of c-MoS2 films, we were able to fabricate chemiresistive pH sensors with centimeter channel length while maintaining low measurement voltages. Our study furthers the understanding of conductive forms of MoS2 and opens a new pathway toward next-generation electronic devices.

Experimental Section

Click to copy section linkSection link copied!

Materials

Unless otherwise mentioned, all organic solvents were HPLC grade and used without further purification. Bulk molybdenum disulfide (∼6 μm to maximum 40 μm, product number 69860, batch number WXBD2352 V) powder, cysteamine (∼95%, product number M9768), undecanethiol (98%, product number 510467), and 4-chlorothiophenol (97%, product number 125237) were purchased from Sigma-Aldrich and used without further purification. Water used for experiments was ultrapure type I water (18.2 MΩ·cm) from a Millipore Simplicity water purification system. Hydrogen peroxide (H2O2) (30%) was purchased from J.T Baker. A bath sonicator (Elmasonic P60H ultrasonic cleaner) was used for sonication, and an Eppendorf MiniSpin Plus microcentrifuge was used for centrifugation.

Sample Preparation

2H-MoS2 was exfoliated by using 45% (v/v) of ethanol in water via sonication (80 kHz frequency, 100% power and sweep mode) for 12 h, and the temperature (30 °C) was controlled during sonication by cooling the bath. (31) The centrifugation process was then optimized for conditions to consist of a first step with 3500 rpm (820g) for 15 min, and then the supernatant underwent a second step of centrifugation at 4500 rpm (1700g) for 3 min. Thus, the 2H-MoS2 was collected in the form of a gray precipitate. The collected 2H-MoS2 was washed with water, and the supernatant was discarded. Three different concentrations of aqueous H2O2 (0.02, 0.06, and 0.33 vol %) were used to ultrasonically suspend the prepared 2H-MoS2. The suspension was first sonicated (37 kHz frequency, 100% power, and sweep mode) for 2, 10, or 20 min and then centrifuged first at 3500 rpm (820g) for 8 min, followed by centrifugation of the supernatant at 10000 rpm (6708g) for 15 min. Shaken samples were prepared by adding different concentrations of aqueous H2O2 into the exfoliated 2H-MoS2 instead of sonicating with aqueous H2O2 and followed the same centrifugation procedure as described above. Then the supernatant was discarded by using a glass pipet, and the precipitate was collected in the form of a gray solid for further use.
For the functionalization of c-MoS2 surfaces with cysteamine (2-aminoethanethiol) and 4-chlorothiophenol, the air-brushed (NEO for Iwata CN gravity feed dual action brush #N4500 using 20 psi of N2) and dried thin films were immersed for 20 h in either a 2.5% (v/v) aqueous cysteamine solution or a 2.5% (v/v) ethanolic solution of 4-chlorothiophenol. They were then rinsed with water (cysteamine) and pure ethanol (4-chlorothiophenol) and dried in a fume hood for 10 min at room temperature for further experiments.

Characterization

A Renishaw inVia Raman spectrometer was used over the range 100–3000 cm–1, with a spectral resolution of 2 cm–1, using a 20× objective in backscattering configuration. All spectra were obtained on three different spots of the same sample by using a fully focused 633 nm laser on a spot size of about 1.2 μm limited to 1% of laser power to avoid sample damage. A JEOL JSM-7000F scanning electron microscope (SEM) was used to obtain high-resolution images of MoS2 at 20 kV. Talos 120C and Titan LB transmission electron microscopes (TEM) were used to obtain low- and high-resolution images of MoS2 at 120 and 300 kV, respectively. The XPS analysis was performed on a Kratos AXIS Supra X-ray photoelectron spectrometer. XPS survey spectra were obtained from an area of approximately 300 × 700 μm2 by using a pass energy of 160 eV. XPS high-resolution spectra were obtained from an area of approximately 300 × 700 μm2 by using a pass energy of 20 eV. XPS was performed on 2H-MoS2 and all c-MoS2 samples to evaluate the effect of H2O2. Exfoliated 2H-MoS2 was used as a comparison. Survey scans and high-resolution spectra of C 1s, O 1s, S 2p, N 1s, Cl 2p, Mo 3d, and the valence band edge were recorded and analyzed. The thickness of the films was measured on a Bruker Alicona Infinite FocusG5 plus 3D optical measurement system using a 50× objective for height measurement (20 nm vertical resolution). Sheet resistance and Hall mobility of the films were measured by using a Nanometrics HL 5500PC Hall effect measurement system.

Device Fabrication for Sheet Resistance

To measure the conductivity of MoS2, the sheet resistance was calculated by using a four-probe geometry. Silicon wafers (Virginia Semiconductor) were thermally oxidized to grow a 1 μm thick silicon dioxide (SiO2) layer for electrical insulation. After oxidation, the wafers were cut into 1 × 1 cm2 pieces with a dicing saw. The wafer pieces were then rinsed with acetone followed by methanol and DI water, dried with nitrogen gas, and kept in a nitrogen storage box for further experiments. In order to lower the contact resistance, Cr (20 nm)/Au (200 nm) 3 × 3 mm2 pads were sputter-deposited onto the four corners of each sample. Kapton tape was used as a mask for MoS2 deposition onto a 7 × 7 mm2 area in the center of the substrate. A mixture of 95% (v/v) ethanol and 5% (v/v) water was used to suspend c-MoS2 for airbrushing so the solvent can evaporate quickly. Airbrushing was performed onto the masked area of the silicon wafer in order to make a homogeneous film by using a NEO for Iwata CN gravity feed dual action brush #N4500 with 20 psi of N2. Gas flow slowed the solvent evaporation during airbrushing as it cooled the surface of the devices; hence, the hot plate temperature was kept at 80 °C. Once c-MoS2 dried to a solid film, the mask was removed. Two devices were fabricated from each MoS2 sample to ensure reproducibility.

Device Fabrication for Chemiresistive Sensors

All two-terminal sensor devices were fabricated on the frosted end of the microscope glass slides (VWR, catalogue no. CA48323-185, dimensions 1 in. × 3 in., ground edges, precleaned twin-frosted end). Glass slides were cleaned by sonicating first in acetone for 15 min followed by sonicating in methanol for 15 min. The slides were then dried at room temperature and stored in nitrogen storage boxes for further use. Two parallel conductive pads (each 6 mm wide) were drawn about 1 cm apart (channel dimension where c-MoS2 would be drop-casted) by using a 9B pencil on the frosted part of glass slide. Kapton tape (1 mil, 1 in. × 36 yd) was used to mask the area (1.5 cm × 1.5 cm) for drop-casting MoS2. The c-MoS2 dispersion was obtained by suspending the gray solid in DI water, and 150 μL of MoS2 suspension was drop-casted on the mask and dried at 100 °C to get a continuous solid film. Once the film was dry, the Kapton tape was removed, and two strips of conductive adhesive copper (Cu) tape (Adafruit Industries, 6 mm wide) were pasted onto the pencil lines and covered with hot glue (all-purpose glue sticks, thermoplastic adhesive) with a hot glue gun (Aleene’s ultimate glue gun kit). The more involved airbrushing technique is only used for sheet resistance measurements since a homogeneous film thickness is required in order to calculate a meaningful sheet resistance value. Sensor films benefit from the higher surface area and simpler fabrication of drop-casted films despite less exacting dimensions.
For the functionalization of c-MoS2 surfaces with cysteamine (2-aminoethanethiol) and undecanethiol the drop-casted and dried thin films were immersed for 20 h in either a 2.5% (v/v) aqueous cysteamine solution or a 2.5% (v/v) toluene solution of undecanethiol. They were then rinsed with water (cysteamine) and toluene (undecanethiol) and dried in a fume hood for 10 min at room temperature for sensing experiments.

pH Measurement

The conductivity of the chemiresistive devices was measured in air by using a two-probe configuration with a Keithley 2450 source meter at room temperature. The sensing measurements were carried out by using a four-channel eDAQ EPU452 Quad Multifunction isoPod with USB (purchased from eDAQ Inc.). Three channels were used for chemiresistive devices, and one channel was used for a pH electrode. The pH electrode was purchased from eDAQ Inc. and calibrated in both pH 4 and 7 at 25 °C before each experiment. The device responses in liquid medium were recorded by using continuous two-probe measurements in Biosensor mode (100 mV applied bias, current range 2 mA, 6 decimal places, scanning time 1.0 s, and scan rate 30 points/min) at room temperature. The entire experiment was carried in an 800 mL bowl filled with 500 mL of a 200 ppm of NaCl salt solution. The bowl was kept on a stir plate with a continuous stirring. Three devices at a time were run in parallel. 1 M HCl or NaOH solutions were added dropwise into the 200 ppm solution by using glass pipettes in order to adjust the pH in 30 min intervals.

Results and Discussion

Click to copy section linkSection link copied!

Optimization for Conductivity

In order to achieve high conductivity, the morphology of the percolation network film needs to be optimized. 2H-MoS2 was first exfoliated according to a reported procedure using ethanol in water via sonication at room temperature. (31) The centrifugation process was then optimized for the fraction of well-exfoliated few-layer 2H-MoS2 flakes with minimum damage. These were collected in the form of a gray precipitate which was washed with water to remove ethanol before being used to prepare c-MoS2 by exposure to dilute aqueous H2O2 (Figure 1). The exfoliated c-MoS2 was then airbrushed onto a silicon dioxide (SiO2) substrate with four gold contacts (Figure 2a) to measure sheet resistance and Hall mobility. Scanning electron microscope (SEM) images of these films (Figure 2b) showed the presence of multilayers and homogeneously distributed flakes with sufficient overlap to ensure good conductivity of the film. This kind of film morphology cannot be achieved without first exfoliating the semiconducting form. TEM images further supported the differences in the surface properties of with and without treated with aqueous H2O2 of bulk MoS2 and exfoliated 2H-MoS2 (Figures 3a–d). Films of c-MoS2 obtained directly from bulk powder were more poorly aligned (Figure S1) and had a lower conductivity compared to those obtained from exfoliated 2H-MoS2. (15) The degree of exfoliation may also play a role in increasing the reactive surface area during sonication in aqueous H2O2. The height of the deposited solid films was typically around 9 μm as determined by high-resolution optical microscopy (Figures 2c,d).

Figure 1

Figure 1. Schematic representation of the exfoliation procedure of c-MoS2. Steps are (a) bulk MoS2 sonication in ethanol/water mixture for 12 h; (b) exfoliated 2H-MoS2 suspended in ethanol/water; (c–e) two-stage centrifugation process to collect exfoliated 2H-MoS2; (f) exfoliated 2H-MoS2 as precipitate; (g) exfoliated 2H-MoS2 shaken in 0.06% aqueous hydrogen peroxide; (h) suspended c-MoS2 after sonication; (i–k) two-stage centrifugation process to collect exfoliated c-MoS2; and (l) exfoliated c-MoS2 as precipitate after centrifugation.

Figure 2

Figure 2. Morphology of MoS2 solid films and sheet resistance as a function of process parameters. (a) Actual image of exfoliated c-MoS2 (gray patch, 7 × 7 mm2) on a SiO2 substrate (1 × 1 cm2) with Au contacts in the four corners of the substrate. (b) SEM images showing overall film distribution of final exfoliated c-MoS2 on the substrate with high magnification showing few-layer flakes of material. The scale bars on the images represent 10 μm and 100 nm, respectively. (c) Alicona optical microscope mapping image of a SiO2 substrate with c-MoS2 film edge (red line is 19.7 mm long). (d) Height profile at the location of the red line in (d). (e) Relationship between the sheet resistance of c-MoS2 samples sonicated for different times in 0.06% and 0.22% aqueous H2O2. Lines are drawn to guide the eye. Solid lines are for 0.22%. The dotted lines are for 0.06%.

Figure 3

Figure 3. TEM images showing film distribution of different MoS2 flasks: (a) bulk MoS2, (b) exfoliated 2H-MoS2, (c) bulk MoS2 treated with 0.06% aqueous H2O2, and (d) exfoliated MoS2 treated with 0.06% aqueous H2O2. The scale bar on the images represents 0.2 μm.

In order to achieve the desired degree of oxidation of MoS2 leading to increased conductivity without dissolution by the highly reactive H2O2 at higher concentrations, it is necessary to identify the optimal peroxide concentration and sonication time. Twelve distinct batches of c-MoS2 were prepared by using three different concentrations of aqueous H2O2 (0.02%, 0.06%, and 0.22%) in which exfoliated 2H-MoS2 samples were either briefly shaken or sonicated for 2, 10, or 20 min. Duplicate devices were fabricated from each batch and characterized by using four-probe sheet resistance and Hall probe measurements at room temperature. The sheet resistance data indicate that the lowest resistances were obtained from 0.06% and 0.22% of H2O2 (Figure 2e), whereas the sheet resistances obtained by treatment with 0.02% H2O2 were similar to those of semiconducting 2H-MoS2 (>109 Ω/□). The sonication time was another key factor in controlling the effect of H2O2 on 2H-MoS2. The bulk resistivity of 0.06% and 0.22% c-MoS2 samples at 20 min was 0.43 Ω·cm (the corresponding sheet resistance of a 9 μm thick film was 4.50 × 102 Ω/□) and 0.40 Ω·cm (the corresponding sheet resistance of a 9 μm thick film was 4.34 × 102 Ω/□) (Tables S1–S4). A trend in decreasing sheet resistance with sonication time was observed at 0.06% c-MoS2 samples, whereas the minimum in sheet resistance for 0.22% c-MoS2 was already reached after 10 min of sonication. The 20 min sonicated sample showed no further change in sheet resistance. The measured sheet resistance for c-MoS2 shaken in either of the two concentrations was 5 orders of magnitude less than that of 2H-MoS2, indicating that sonication is not fundamentally required in order to improve the conductivity of 2H-MoS2. Sonication, however, exfoliates and reshuffles the sheets, thus increasing the film conductivity by another 2 orders of magnitude in addition to allowing the material to retain a higher conductivity even if the surface undergoes further modification (as demonstrated below). We also tested higher concentrations of H2O2 but found that MoS2 tended to completely dissolve in those cases following the formation of hydrated molybdenum trioxide and sulfur dioxide. (32,33) Based on the above observations, samples were prepared by using 0.06% H2O2 in all subsequent experiments to maximize yield.
The Hall mobility of the same samples used to measure the sheet resistance was measured at room temperature. It varied between 226 and 355 cm2 V–1 s–1 for different sonication times (Figure S3), which was comparable to some reported values for multilayer MoS2 FETs (34) but higher than other reports for p-doped MoS2 FETs. (35) The Hall mobility depends on several factors such as number of layers, metal contacts, and surface of the materials. (36,37) The positive sign of the Hall mobility values indicates that holes were the majority charge carriers as a result of p-type doping. (36,37)

Identifying the Source of Increased Conductivity

XRD data of the c-MoS2 phase (Figures S3 and S4) show it to be a (doped) 2H-MoS2 phase (a broad (002) peak at 2θ ∼ 14.5°) rather than any of the metallic 1T phases ((001) peak at 2θ ∼ 7.3° for 1T-MoS2) reported in the literature, (38) consistent with the Hall measurements indicating a p-doped semiconductor. No changes in the peak at 2θ of bulk MoS2, 2H-MoS2, and c-MoS2 are observed in the XRD data (Figures S3 and S4). This is also borne out by Raman and XPS data. The Raman spectra for all samples display the two distinct E12g and A1g modes. The in-plane E12g mode at 385 cm–1 originates in the antiparallel vibration of sulfur atoms with respect to the molybdenum atoms, whereas the out-of-plane A1g mode at 410 cm–1 involves the vibration of the sulfur atoms in the opposite direction. These are the most prominent peaks for identification of the properties of 2H-MoS2. The absence of three characteristic Raman peaks at ∼156, ∼226, and ∼333 cm–1 further confirms that our samples do not contain any detectable amounts of 1T-MoS2. (39) The XPS binding energies for Mo 3d5/2, Mo 3d3/2, S 2p3/2, and S 2p1/2 in 1T-MoS2 have been reported to be 228.7, 231.8, 161.6, and 163.7 eV, respectively, about 1 eV lower than those for 2H-MoS2. (13,15) XPS analysis of all H2O2-treated c-MoS2 samples shows the above binding energies to be consistent with 2H-MoS2 (Figure 4). The location of the valence band edge further confirms the semiconducting nature of our samples (Figure S5).

Figure 4

Figure 4. XPS high-resolution spectra of exfoliated 2H-MoS2; shaken c-MoS2; 2 min sonicated c-MoS2; and 20 min sonicated c-MoS2 samples. Spectra are (a) Mo 3d, (b) S 2p in nonfunctionalized MoS2, and (c) S 2p in cysteamine-functionalized MoS2.

Sulfur vacancies can also have a significant impact on the electronic properties of MoS2 samples. Combined XPS survey and high-resolution data demonstrate that the S2–/Mo4+ ratio does not decrease for MoS2 upon exposure to dilute aqueous H2O2 without sonication when compared to the exfoliated 2H-MoS2 (Table 1). Since the biggest change in conductivity was seen as the result of this step, we conclude that the increase in conductivity also cannot be the result of an increase in sulfur vacancies. There is a small drop in the S2–/Mo4+ ratio for the sonicated samples, likely due to further damage inflicted during sonication, but the change is very small and not correlated with the evolution in conductivity (Table 1). Since no chemical elements beyond H, C, O, S, and Mo have been introduced into our samples during processing, the three most common explanations for conductivity in MoS2 (1T metallic phases, sulfur vacancies, and substitutional doping at the Mo sites) (13,16,40) have thus been excluded from consideration, and we have to look elsewhere to explain the nature of our c-MoS2 phase.
Table 1. Compositional Changes in the Bare Samples from XPS Dataa
sampleS2–/Mo4+Mo4+/MoMo5+/MoMo6+/Mosheet resistance (Ω /□)bulk resisivity (Ω·cm)
2H-MoS21.3590.952 0.0483.0 × 1092.7 × 106
shaken c-MoS21.3720.6840.1640.1512.5 × 10422
2 min c-MoS21.3290.8080.0940.0971.3 × 10412
10 min c-MoS21.3300.8200.0900.0904.1 × 1034.0
20 min c-MoS21.3210.949 0.0514.4 × 1020.43
a

S to Mo atomic ratio of 2H-MoS2 and c-MoS2 samples. The atomic ratio of sulfide to Mo(IV) was calculated from the total atomic percentages of Mo and S as well as high-resolution XPS spectra of S 2p and Mo 3d. The atomic ratios of Mo(IV), Mo(V), and Mo(VI) relative to the total Mo content in 2H-MoS2 and c-MoS2 samples were calculated by using high-resolution XPS spectra of Mo 3d. The bulk resistivity is calculated from the product of the sheet resistance and thickness of ∼9 μm thick films (average of two devices each).

Spectroscopic Determination of the Nature of c-MoS2

Figure 4a shows high-resolution Mo 3d spectra of c-MoS2 shaken or sonicated for 2 or 20 min (Figure S6a for 10 min sonicated c-MoS2) in 0.06% H2O2, while the Mo 3d spectrum of the exfoliated 2H-MoS2 sample is shown in Figure 4a. The Mo 3d doublet shows a Mo 3d5/2 binding energy of 229.7 eV, which is characteristic for Mo4+ in 2H-MoS2. (41,42) A Mo 3d5/2 binding energy of 233.3 eV is characteristic of Mo6+ such as in molybdenum trioxide (MoO3). (41,42) The same +6 oxidation peaks of Mo 3d5/2 are present in the H2O2-treated c-MoS2 samples (shaken, 2, 10, or 20 min sonicated in Figures 4a and Figure S6a). In addition to the +6 and +4 oxidation states, Mo 3d5/2 peaks at 232.1 eV attributable to the formation of Mo5+ were observed for shaken, 2 min, and 10 min c-MoS2 samples. This indicates that while the exfoliated 2H-MoS2 was already partially oxidized during the sonication in the ethanol/water mixture, adding H2O2 causes further oxidation of 2H-MoS2 (Figure 6).
Decomposition of H2O2 can cause formation of atomic hydrogen, (43,44) which may react further with species already present on the 2H-MoS2 surface (e.g., MoO3) and result in the formation of hydrogen molybdenum bronze HxMoO3 and substoichiometric MoO3–y. (45,46) It was reported that the reduction of MoO3 leads to the formation of substoichiometric MoO3–y where an increase in y is correlated with a decrease of the electronic bandgap of MoO3, making the material conductive. (47) The bronze also contains Mo5+, as observed in XPS. HxMoO3 is significantly more conductive than 2H-MoS2 and MoO3. (48,49) The sheet resistance of our c-MoS2 material is only about 1 order of magnitude higher than those reported for MoO2, HxMoO3–y nanobelts, and high-temperature carbon electrodes. (48,50) Hence, we propose the conductivity of the c-MoS2 samples shaken or sonicated up to 10 min in H2O2 to be due to the presence of HxMoO3 and substoichiometric MoO3–y.
The Mo5+ peak is no longer present in the spectrum of the sample sonicated for 20 min since it was the result of H2O2 interacting with the surface, and the small amount of H2O2 will have been consumed by reaction with MoS2 or ultrasonically decomposed by that point in the process. The ratio of Mo5+/Mo in the high-resolution XPS data is highest for shaken samples (0.16) and decreases to 0.09 upon sonication for both 2 and 10 min (Table 1). No Mo5+ signal remains for 20 min sonicated samples, implying that HxMoO3 and MoO3–y are unstable intermediate species that are gradually reduced to MoO2, consistent with growth of the Mo4+ signal in the spectra. (49) The high-resolution Mo 3d peaks should be broader for MoO2 than for MoS2, but the relatively small amounts of MoO2 are likely being obscured by the large Mo4+ signal from MoS2 in the 20 min samples. In addition to the formation of Mo5+, the amount of Mo6+ is found to increase upon exposure to H2O2 coupled with a significant drop in the proportion of Mo4+ (Table 1). Upon sonication, the proportion of Mo4+ gradually recovers over time while Mo5+ and Mo6+ levels gradually decrease back to resemble the 2H-MoS2 starting material, but with a much higher conductivity.
While XPS spectroscopy only probes the top 10 nm from the sample surface, Raman spectroscopy can help to further understand the bulk structure and properties of c-MoS2, which is important because the flakes will be continuously exfoliated and recombined during sonication, thus exposing fresh surfaces to the dilute aqueous peroxide. The gap between E12g and A1g modes is narrowed by 2 cm–1 due to a slight blue shift of the E12g peak combined with a slight red shift of the A1g peak for c-MoS2 as compared to the initial 2H phase (Figure 5a). The shift is very small since our samples consist of multilayers (SEM data, Figure 2b), and both modes have been reported to stiffen as the number of layers increases due to an increase in the restoring force on the atoms. (51) The peaks at 178, 423, 466, 526, 600, and 644 cm–1 for all samples (2H, shaken, and 20 min c-MoS2, Figure 5b–d) are due to resonant Raman scattering (633 nm laser) of MoS2. (52) Two peaks at 570 and 738 cm–1 can be attributed to vibrational modes of MoO2, while the peaks 230 and 492 cm–1 can be assigned to the phonon modes of MoO2. (53,54) The Raman features for MoO2 are very weak, since only a small amount of MoO2 was evidently formed in agreement with our interpretation of the XPS data. The Raman peak at 820 cm–1 is the most prominent peak attributable to MoO3. (46) It is important to note that MoO3 is an insulator and therefore does not contribute to either surface or bulk conductivity. While we did not observe any HxMoO3 peaks (204 cm–1) (46) in the Raman spectrum of either sample, some features around 780 cm–1 were instead detected for all samples. Broad features in this region have previously been attributed to substoichiometric MoO3–y. (47) While insulating MoO3 is clearly present in all samples starting with 2H-MoS2, the presence of minor MoO2 and MoO3–y impurities can be confirmed (although not quantified) by using Raman spectroscopy. The absence of a 440 cm–1 band in the Raman spectra (52) supports the notion that no oxysulfide species were formed during the exfoliation process, in agreement with the high-resolution S 2p XPS data from all MoS2 samples (Figure 4b and Figure S6b). Since the main Raman characteristic peaks of MoS2 still dominate the spectra, we conclude that even after oxidation the quality of the MoS2 material was maintained throughout the sonication process, as suggested earlier by the constant S/Mo ratio (Table 1).

Figure 5

Figure 5. Raman spectra of MoS2 samples. (a) Raman spectra proving the doping effect of H2O2 on 2H-MoS2. Black and violet color curves represent c-MoS2 (20 min sonicated) and 2H-MoS2, respectively, showing the shift in the E12g and A1g modes. (b) Raman spectrum of 2H-MoS2 sample. (c) Raman spectrum of c-MoS2 sample shaken in 0.06% H2O2. (d) Raman spectrum of 20 min sonicated c-MoS2 sonicated sample. All spectra were recorded with a 633 nm laser at 1% power. Each spectrum is normalized based on the highest peak (∼466 cm–1).

Figure 6

Figure 6. Schematic representation of structure and chemical composition of the formation of c-MoS2.

Probing the c-MoS2 Surface via Thiol Chemistry

Exposure to H2O2 helped to partially convert the MoO3 (which formed during exfoliation of 2H-MoS2 due to the oxidation of MoS2) to new compounds like hydrogen molybdenum bronze and substochiometric MoO3–y. While the S/Mo ratio did not change significantly as a result of the reaction with H2O2, reactive sulfur vacancy defects may have remained on the c-MoS2 surface or at the edges of the flakes. The c-MoS2 surface was titrated with different thiol molecules, namely cysteamine and 4-chlorothiophenol, to determine the number of residual reactive sites.
The high-resolution S 2p XPS spectra of all 4-chlorothiophenol-treated c-MoS2 samples are dominated by sulfide peaks associated with MoS2, with a binding energy of 162.4 eV observed for 2p3/2 (Figure S8). Thiol-related 2p3/2 peaks at a binding energy of 164.0 eV were only observed in 2H-MoS2 as well as 2 and 10 min sonicated c-MoS2 samples (Figure S8). If the sulfur vacancies had been functionalized by thiols, a thiolate peak would have been expected at a binding energy of 161.8 eV, which is absent in all our spectra. (55) The presence of a small thiol peak suggests the presence of unbound thiol on the surface, (55) but the thiol to total sulfur ratio was very small (Table 2). The very small Cl/Mo ratio (Table 2) further confirms that only a negligible amount of 4-chlorothiophenol remained at the surfaces of 2H-MoS2 and all c-MoS2 samples. The samples had been exposed to ambient conditions prior to thiolation, so oxidation at the sulfur vacancy sites was expected. High-resolution XPS spectra of the Cl 2p peaks of all 4-chlorothiophenol-functionalized c-MoS2 (Figure S7) samples confirm that chlorine remained bonded to organic carbon at a binding energy of 200.7 eV for 2p3/2 and 202.3 eV for 2p1/2. (56) Overall, 4-chlorothiophenol reacted with neither the c-MoS2 nor the 2H-MoS2 surfaces under our given experimental conditions.
Table 2. Compositional Changes in Functionalized Samples from XPS Dataa
 cysteamine4-chlorothiophenol
sampleN/Mothiol/SCl/Mothiol/S
2H-MoS20.4560.1350.0040.021
shaken c-MoS20.2680.0230.007 
2 min c-MoS20.3360.0680.0090.028
10 min c-MoS20.4120.0500.0100.023
20 min c-MoS20.3840.1120.008 
a

Atomic ratios of nitrogen (N) and chlorine (Cl) after functionalization with respect to total Mo for each sample. The atomic ratios were calculated from the total atomic percentages of Mo and S as well as high-resolution XPS spectra of N 1s, Cl 2p, and Mo 3d. Atomic ratios of thiols (cysteamine and 4-chlorothiophenol) were calculated by using high-resolution XPS spectra of S (thiol) 2p with respect to total S for each sample.

In all cysteamine-treated 2H-MoS2 and c-MoS2 samples, S 2p3/2 peaks are observed at 164.0 eV for unbound thiols, and S 2p3/2 peaks associated with sulfide in MoS2 are observed at 162.4 eV (Figure 4c and Figure S10). (55) The unbound thiol to total sulfur ratio is consistently much higher in all cysteamine-treated samples compared to 4-chlorothiophenol-treated samples (Table 2). Unsurprisingly, the N/Mo atomic ratio in these samples is also higher compared to the Cl/Mo atomic ratio in 4-chlorothiophenol-treated samples. Hence, at our given experimental conditions for the respective procedures, cysteamine reacts more easily than 4-chlorothiophenol with both c-MoS2 and 2H-MoS2 samples. It is clear from the XPS data that the thiols did not attach to any reactive sulfur vacancies, as no thiolate peak was detected. Furthermore, we only observed a single N 1 s peak associated with free amines at a binding energy of 399.3 eV (Figure S9). (57) While the amine group may participate in noncovalent interactions with the surface, there is no spectroscopic evidence of it. While the sulfide to Mo(IV) ratios as given in Table 1 for all samples are in the range 1.3–1.4 rather than close to the ideal value of 2.0, the defects have been passivated with oxygen species under the given circumstances and are not accessible to thiol functionalization in organic solvents. The aqueous conditions of cysteamine functionalization appear to have facilitated surface interactions, even though neither the thiol nor the amine group was found to covalently react with the surface. The availability of lone electron pairs at oxygen and sulfur surface sites may nevertheless enable hydrogen-bonding interactions or acid–base chemistry.

Direct In-Solution Measurement of pH Using a c-MoS2-Based Chemiresistive Sensor

A potential application for surface-functionalized thin conductive films (such as those fabricated from c-MoS2) is chemiresistive sensing, for example, of the pH of aqueous solutions (Figure 7). Chemiresistive sensors are solid-state electrical devices based on measuring the change in conductivity of a thin film as a result of the interaction between the active layer and an analyte. (58) While this sensor geometry is generally easy to fabricate and use, MoS2-based chemiresistive sensors operating at low enough voltages to be safe for aqueous analytes are rare due to their low conductivity. We fabricated nanometer thick pristine and cysteamine-functionalized c-MoS2 films on glass substrates as active layers for chemiresistive sensors (Figure S11). The thickness of the c-MoS2 film was calculated based on the bulk resistivity of c-MoS2 material. The sensor films are estimated to be at least 100 nm thick considering their geometry and the reported bulk resistivity of c-MoS2 but neglecting contact resistance and film roughness. The contact resistance cannot be determined in a two-terminal geometry (and will be different from our four terminal measurements due to the different nature of the contacts); the roughness of the drop-cast films will add uncertainty to any film thickness determination, but we can state that the sensor films are a few 100 nm in thickness at most.

Figure 7

Figure 7. Schematic representation of the fabrication and potential applications of exfoliated c-MoS2.

Before using those chemiresistive devices to measure pH, we recorded current vs voltage graphs in ambient conditions after each fabrication step, including film deposition, passivation of the metal contacts with hot glue, and functionalization with cysteamine (Figure S12). We observed a decrease in current (from 38 to 12 μA at 1 V) after passivation of the metal contacts, whereas after cysteamine functionalization the current dropped by about an order of magnitude. The same devices were then immersed into water to record their pH sensing responses. We used 200 ppm of sodium chloride as a background supporting electrolyte to avoid structural changes in the electrical double layer during acid and base addition for pH adjustment. (59,60)
The transition of the devices from air into an aqueous environment caused another decrease in film conductivity by about 1 order of magnitude, depending on whether deionized water or salt solution was used (Figure S13). This is due to electrostatic gating by the electrical double layer. Current changes over time were recorded for all devices starting at 0 min upon immersion in water. Figure 8e shows the pH sensing response of a pristine c-MoS2 film. While an overall pH ranging from 2.7 to 10 (Figure S14) was tested, no changes in current were observed above pH 7. Even though the pristine device shows a marked response to changes in pH of almost 100% increase in sensor current at pH 4 relative to the baseline at pH 6.8, the sensor response is not very stable upon exposure to alternating higher and lower pH values (Figure 8e). Additional pristine c-MoS2 devices were also fabricated, showing comparable sensor responses and baseline drift (Figure S15). The drift may be the result of pH-related instability of the molybdenum oxide phases. (61) We thus examined the relationship between solution pH and electrochemical open circuit potential of the sensor films against a reference electrode (Ag/AgCl) (Figure S16). This relationship can then be overlaid with established Pourbaix diagrams of molybdenum (oxide, sulfide)/water systems to predict the thermodynamic stability of different species as a function of pH and electrochemical potential. (62,63) The result is that no phase changes are predicted in the potential region transversed by the c-MoS2 films at open circuit potential and relevant pH. Further, Raman spectra were collected of two chemiresistive devices that were exposed to pH 2.5 (Figure S17) or pH 9 (Figure S18) for 40 min, immediately dried with nitrogen gas and transferred to the spectrometer within a minute. Neither sample shows any changes in the spectral features compared to a freshly prepared film of c-MoS2 (Figure 4d). Hence, it is further confirmed that bulk phase changes do not occur during the protonation and deprotonation process, leaving modulations of the doping level due to protonation and deprotonation of surface oxide species as the origin of the sensing response.

Figure 8

Figure 8. pH sensing response of c-MoS2 chemiresistive devices. 100 mV potential bias was applied across the c-MoS2 film to measure the current changes. Cysteamine-functionalized c-MoS2 chemiresisitive pH responses at (a) 3.7, (b) 6.5, (c) 4, and (d) 3.5. (e) Pristine c-MoS2 chemiresistive response to different pH values between 4 and 6.5. (f) Cysteamine-functionalized c-MoS2 chemiresistive response to different pH values between 3 and 6.5. (g) Calibration curve (linear fitting) of functionalized c-MoS2 chemiresistive device response (b). (h) Calibration curve from (g) replotted as a function of pH.

The pH response of a cysteamine-functionalized device was significantly higher (ranging from 20% at pH 6.5 over 124% at pH 4 to almost 642% at pH 3) (Figures 8a–d,f) and significantly more stable compared to the pristine devices. Because of the previously mentioned increase in film resistivity upon functionalization with cysteamine, the baseline currents of the functionalized devices are noticeably lower than those of the pristine devices. The functionalization of MoS2 with cysteamine has been reported to result in n-type doping. (21) Since the pristine c-MoS2 films are p-doped, this would lead in effect to a reduction in charge carrier density and thus the observed decrease in the conductivity upon functionalization (Figure S12). Protonation of the amino group in the cysteamine at low pH reduces its ability to act as an electron donor (i.e., n-dopant) (Figure 9). The effective doping level of the c-MoS2 films will thus be increased again, leading to the significant sensor response. In contrast, undecanethiol-functionalized devices (containing a thiol group but no amino group) did not exhibit any reproducible pH responses. Moreover, the obtained response was much lower (15% at pH 6.5, 20% at pH 4.2, etc., Figure S19) compared to even pristine devices (30% at pH 6.5, 90% at pH 4, etc.) devices. Not only does the amino group serve to amplify the pH response, but the response of the pristine films can be suppressed by passivation of the defect sites with thiols.

Figure 9

Figure 9. Schematic representation of protonation and deprotonation process of cysteamine-functionalized c-MoS2.

The sensor response of the cysteamine-functionalized device in Figure 6b was linearly related to the concentration of hydrogen ions in the solution with an R2 value of 0.9864 (Figure 8g). This linear concentration dependence results in a logarithmic sensor response to pH (Figure 8f). Additional devices functionalized with cysteamine were tested and showed qualitatively similar behavior (Figure S20). The simple fabrication method via drop-casting of the c-MoS2 resulted in an uneven distribution of the c-MoS2 flakes on the glass slides and widely varying film resistances. Nevertheless, stable and high sensor responses were observed from all devices upon repeated exposure to pH 4 and 6.5 (Figures S20). The long-term stability of the sensors was further investigated. Functionalized c-MoS2 devices were kept in the 200 ppm of NaCl solution after the pH measurement for about 1 week, and no significant baseline drift was observed (Figure S21), indicating that functionalized devices are stable in aqueous environments over long periods of time.
The new form of conductive MoS2 presented here is a semiconducting 2H-MoS2 phase very heavily p-doped by hydrogen molybdenum bronze (HxMoO3) and substoichiometric MoO3–y, which are formed due to the interaction of very dilute H2O2 with the MoS2 surface. XRD, XPS, Raman, and Hall measurements all corroborate this mechanism. Doping of MoS2 by H2O2 itself cannot be the origin of the conductivity since it would lead to an increase of the band gap by ∼40 meV. (32) The interaction of dilute H2O2 with exfoliated 2H-MoS2 causes HxMoO3 and MoO3–y formation, while sonication in dilute aqueous H2O2 further reduces the sheet resistance of the bulk material down to about 440 Ω /□ (Hall mobilities as high as 355 cm2 V–1 s–1) without compromising the structure of the p-doped 2H-MoS2 phase.
The resulting c-MoS2 material has a wide range of possible applications in batteries, supercapacitors, solar cells, electrocatalysts, or sensors. We have chosen to demonstrate the use of c-MoS2 films in a chemiresistive geometry as pH sensors within a pH range of 3–6.5 for both pristine and functionalized c-MoS2 devices. Functionalized devices showed a stable sensor response of up to 2000% at pH 3 compared to a neutral solution baseline, depending linearly on the hydrogen ion concentration. Pristine devices still have a significant response of over 100%, but the directly exposed substoichiometric oxide phases are not stable under acidic conditions, causing a drift in the sensor baseline. This instability is avoided by passivation of the defect sites with thiols. For unfunctionalized thiols, the passivation also suppresses the pH response, while amino-functionalized thiols (e.g., cysteamine) serve to enhance the pH response. Chemiresistive solid-state sensor devices are cheaper to fabricate, simpler to use, and lower in maintenance than electrochemical sensors since they do not require any reference electrode or gate electrode. (58,64) Most MoS2 sensors reported to date are FET-based gas sensors. (7,65,66) Recently, a MoS2 chemiresistive sensor has been reported to detect the cadmium cations in aqueous environments, but the resistances of the reported 2H-MoS2 devices were much higher the c-MoS2 devices presented here, thus requiring a higher voltage bias that may interfere with measurements in aqueous electrolytes. (67) The lower sheet resistance of c-MoS2 allows for simpler fabrication, more favorable operating parameters (lower voltage and higher current), and a higher sensor response of the chemiresistive films for applications in aqueous environments.

Conclusions

Click to copy section linkSection link copied!

In summary, a stable conductive form of MoS2 was exfoliated in solution by using a two-step ultrasonication procedure using an ethanol/water mixture in the first step and dilute aqueous H2O2 in the second step. Centrifugation then results in a slurry of micrometer-sized few-layer p-doped 2H-MoS2 flakes that can be processed into conductive nanometer-scale films with <1 Ω·cm in bulk resistivity and good carrier mobility. Chemiresistive devices that perform well and are stable in aqueous environments over a wide pH range have been demonstrated after passivation with cysteamine. While stability of the c-MoS2 under a wider range of conditions still has to be studied, the process is easily scalable for research and industrial applications and may be followed by surface passivation steps.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsanm.0c02135.

  • Additional SEM, additional TEM, sheet resistance, Hall measurement, XRD, XPS, Raman, and pH sensor data (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Dipankar Saha - Department of Chemistry and Chemical Biology, McMaster University, 1280 Main Street West, Hamilton, Ontario L8S 4M1, CanadaOrcidhttp://orcid.org/0000-0002-6268-2807
    • Ponnambalam Ravi Selvaganapathy - Department of Mechanical Engineering and School of Biomedical Engineering, McMaster University, 1280 Main Street West, Hamilton, Ontario L8S 4L7, CanadaOrcidhttp://orcid.org/0000-0003-2041-7180
  • Notes
    The authors declare the following competing financial interest(s): The authors have filed for a provisional patent covering the fabrication process described in this manuscript.

Acknowledgments

Click to copy section linkSection link copied!

We are grateful to Dr. Mark Biesinger (Surface Science Western) for help with XPS; Mohammad Shariful Islam Chowdhury and Ali Aliakbari Khoei (McMaster Manufacturing Research Institute, McMaster University) for help with optical microscopy; Dr. Ross Anthony (Centre for Emerging Device Technologies, McMaster University) for help with oxidation of the silicon wafers; Dr. Shahram Tavakoli and Doris Stevanovic (both from Centre for Emerging Device Technologies, McMaster University) for help with sputtering, sheet resistance, and Hall measurements; Dr. James F. Britten and Victoria Jarvis (both from McMaster University) for help with XRD measurement; Dr. Carmen Andrei (CCEM-McMaster University) for help with TEM; and Vinay Patel, Sreekant Damodara, Jayasree Biswas, Dr. Aditya Aryasomayajula, Ana Zubiarrain Laserna, Marcia Reid, Chris Butcher, Maryam Darestani-Farahani, and Johnson Dalmieda (all from McMaster University) for fruitful discussions. Electronic microscopy was carried out at the Electron Microscopy Facility of the Faculty of Health Sciences (McMaster University) and at the Canadian Centre for Electron Microscopy (CCEM) McMaster University, a national facility. The work was financially supported by the Natural Sciences and Engineering Research Council of Canada through the Discovery Grant Program as well as the Canada First Research Excellence Fund project “Global Water Futures”.

References

Click to copy section linkSection link copied!

This article references 67 other publications.

  1. 1
    Dragoman, M.; Dinescu, A.; Dragoman, D. 2D Materials Nanoelectronics: New Concepts, Fabrication, Characterization from Microwaves up to Optical Spectrum. Phys. Status Solidi A 2019, 216 (8), 1800724,  DOI: 10.1002/pssa.201800724
  2. 2
    Akinwande, D.; Petrone, N.; Hone, J. Two-Dimensional Flexible Nanoelectronics. Nat. Commun. 2014, 5 (1), 5678,  DOI: 10.1038/ncomms6678
  3. 3
    Chhowalla, M.; Shin, H. S.; Eda, G.; Li, L.-J.; Loh, K. P.; Zhang, H. The Chemistry of Two-Dimensional Layered Transition Metal Dichalcogenide Nanosheets. Nat. Chem. 2013, 5 (4), 263275,  DOI: 10.1038/nchem.1589
  4. 4
    Eftekhari, A. Tungsten Dichalcogenides (WS2, WSe2, and WTe2): Materials Chemistry and Applications. J. Mater. Chem. A 2017, 5 (35), 1829918325,  DOI: 10.1039/C7TA04268J
  5. 5
    Late, D. J.; Liu, B.; Matte, H. S. S. R.; Dravid, V. P.; Rao, C. N. R. Hysteresis in Single-Layer MoS2 Field Effect Transistors. ACS Nano 2012, 6 (6), 56355641,  DOI: 10.1021/nn301572c
  6. 6
    Divigalpitiya, W. M. R.; Morrison, S. R.; Frindt, R. F. Thin Oriented Films of Molybdenum Disulphide. Thin Solid Films 1990, 186 (1), 177192,  DOI: 10.1016/0040-6090(90)90511-B
  7. 7
    Kiriya, D.; Tosun, M.; Zhao, P.; Kang, J. S.; Javey, A. Air-Stable Surface Charge Transfer Doping of MoS2 by Benzyl Viologen. J. Am. Chem. Soc. 2014, 136 (22), 78537856,  DOI: 10.1021/ja5033327
  8. 8
    Sarkar, D.; Liu, W.; Xie, X.; Anselmo, A. C.; Mitragotri, S.; Banerjee, K. MoS2 Field-Effect Transistor for Next-Generation Label-Free Biosensors. ACS Nano 2014, 8 (4), 39924003,  DOI: 10.1021/nn5009148
  9. 9
    Stephenson, T.; Li, Z.; Olsen, B.; Mitlin, D. Lithium Ion Battery Applications of Molybdenum Disulfide (MoS2) Nanocomposites. Energy Environ. Sci. 2014, 7 (1), 209231,  DOI: 10.1039/C3EE42591F
  10. 10
    Cao, L.; Yang, S.; Gao, W.; Liu, Z.; Gong, Y.; Ma, L.; Shi, G.; Lei, S.; Zhang, Y.; Zhang, S.; Vajtai, R.; Ajayan, P. M. Direct Laser-Patterned Micro-Supercapacitors from Paintable MoS2 Films. Small 2013, 9 (17), 29052910,  DOI: 10.1002/smll.201203164
  11. 11
    Deng, Z. H.; Li, L.; Ding, W.; Xiong, K.; Wei, Z. D. Synthesized Ultrathin MoS2 Nanosheets Perpendicular to Graphene for Catalysis of Hydrogen Evolution Reaction. Chem. Commun. 2015, 51 (10), 18931896,  DOI: 10.1039/C4CC08491H
  12. 12
    Tsai, M.-L.; Su, S.-H.; Chang, J.-K.; Tsai, D.-S.; Chen, C.-H.; Wu, C.-I.; Li, L.-J.; Chen, L.-J.; He, J.-H. Monolayer MoS2 Heterojunction Solar Cells. ACS Nano 2014, 8 (8), 83178322,  DOI: 10.1021/nn502776h
  13. 13
    Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.; Chhowalla, M. Photoluminescence from Chemically Exfoliated MoS2. Nano Lett. 2011, 11 (12), 51115116,  DOI: 10.1021/nl201874w
  14. 14
    Xia, J.; Wang, J.; Chao, D.; Chen, Z.; Liu, Z.; Kuo, J.-L.; Yan, J.; Shen, Z. X. Phase Evolution of Lithium Intercalation Dynamics in 2H-MoS2. Nanoscale 2017, 9 (22), 75337540,  DOI: 10.1039/C7NR02028G
  15. 15
    Geng, X.; Sun, W.; Wu, W.; Chen, B.; Al-Hilo, A.; Benamara, M.; Zhu, H.; Watanabe, F.; Cui, J.; Chen, T.-p. Pure and Stable Metallic Phase Molybdenum Disulfide Nanosheets for Hydrogen Evolution Reaction. Nat. Commun. 2016, 7 (1), 10672,  DOI: 10.1038/ncomms10672
  16. 16
    Dabral, A.; Lu, A. K. A.; Chiappe, D.; Houssa, M.; Pourtois, G. A Systematic Study of Various 2D Materials in the Light of Defect Formation and Oxidation. Phys. Chem. Chem. Phys. 2019, 21 (3), 10891099,  DOI: 10.1039/C8CP05665J
  17. 17
    Xie, Y.; Liang, F.; Chi, S.; Wang, D.; Zhong, K.; Yu, H.; Zhang, H.; Chen, Y.; Wang, J. Defect Engineering of Mos2 for Room-Temperature Terahertz Photodetection. ACS Appl. Mater. Interfaces 2020, 12 (6), 73517357,  DOI: 10.1021/acsami.9b21671
  18. 18
    Saha, D.; Kruse, P. Editors’ Choice─Review─Conductive Forms of MoS2 and Their Applications in Energy Storage and Conversion. J. Electrochem. Soc. 2020, 167 (12), 126517,  DOI: 10.1149/1945-7111/abb34b
  19. 19
    Kc, S.; Longo, R. C.; Addou, R.; Wallace, R. M.; Cho, K. Impact of Intrinsic Atomic Defects on the Electronic Structure of MoS2 Monolayers. Nanotechnology 2014, 25 (37), 375703,  DOI: 10.1088/0957-4484/25/37/375703
  20. 20
    McDonnell, S.; Addou, R.; Buie, C.; Wallace, R. M.; Hinkle, C. L. Defect-Dominated Doping and Contact Resistance in MoS2. ACS Nano 2014, 8 (3), 28802888,  DOI: 10.1021/nn500044q
  21. 21
    Sim, D. M.; Kim, M.; Yim, S.; Choi, M.-J.; Choi, J.; Yoo, S.; Jung, Y. S. Controlled Doping of Vacancy-Containing Few-Layer MoS2 Via Highly Stable Thiol-Based Molecular Chemisorption. ACS Nano 2015, 9 (12), 1211512123,  DOI: 10.1021/acsnano.5b05173
  22. 22
    Förster, A.; Gemming, S.; Seifert, G.; Tománek, D. Chemical and Electronic Repair Mechanism of Defects in MoS2 Monolayers. ACS Nano 2017, 11 (10), 99899996,  DOI: 10.1021/acsnano.7b04162
  23. 23
    Kc, S.; Longo, R. C.; Wallace, R. M.; Cho, K. Surface Oxidation Energetics and Kinetics on Mos2 Monolayer. J. Appl. Phys. 2015, 117 (13), 135301,  DOI: 10.1063/1.4916536
  24. 24
    Lu, H.; Kummel, A.; Robertson, J. Passivating the Sulfur Vacancy in Monolayer Mos2. APL Mater. 2018, 6 (6), 066104,  DOI: 10.1063/1.5030737
  25. 25
    Nan, H.; Wang, Z.; Wang, W.; Liang, Z.; Lu, Y.; Chen, Q.; He, D.; Tan, P.; Miao, F.; Wang, X.; Wang, J.; Ni, Z. Strong Photoluminescence Enhancement of MoS2 through Defect Engineering and Oxygen Bonding. ACS Nano 2014, 8 (6), 57385745,  DOI: 10.1021/nn500532f
  26. 26
    Verhagen, T.; Guerra, V. L. P.; Haider, G.; Kalbac, M.; Vejpravova, J. Towards the Evaluation of Defects in MoS2 Using Cryogenic Photoluminescence Spectroscopy. Nanoscale 2020, 12 (5), 30193028,  DOI: 10.1039/C9NR07246B
  27. 27
    Mohtasebi, A.; Broomfield, A. D.; Chowdhury, T.; Selvaganapathy, P. R.; Kruse, P. Reagent-Free Quantification of Aqueous Free Chlorine Via Electrical Readout of Colorimetrically Functionalized Pencil Lines. ACS Appl. Mater. Interfaces 2017, 9 (24), 2074820761,  DOI: 10.1021/acsami.7b03968
  28. 28
    Hoque, E.; Hsu, L. H. H.; Aryasomayajula, A.; Selvaganapathy, P. R.; Kruse, P. Pencil-Drawn Chemiresistive Sensor for Free Chlorine in Water. IEEE Sensors Letters 2017, 1 (4), 14,  DOI: 10.1109/LSENS.2017.2722958
  29. 29
    Dalmieda, J.; Zubiarrain-Laserna, A.; Ganepola, D.; Selvaganapathy, P. R.; Kruse, P. Chemiresistive Detection of Silver Ions in Aqueous Media. Sens. Actuators, B 2021, 328, 129023,  DOI: 10.1016/j.snb.2020.129023
  30. 30
    Gou, P.; Kraut, N. D.; Feigel, I. M.; Bai, H.; Morgan, G. J.; Chen, Y.; Tang, Y.; Bocan, K.; Stachel, J.; Berger, L.; Mickle, M.; Sejdić, E.; Star, A. Carbon Nanotube Chemiresistor for Wireless pH Sensing. Sci. Rep. 2015, 4 (1), 4468,  DOI: 10.1038/srep04468
  31. 31
    Zhou, K.-G.; Mao, N.-N.; Wang, H.-X.; Peng, Y.; Zhang, H.-L. A Mixed-Solvent Strategy for Efficient Exfoliation of Inorganic Graphene Analogues. Angew. Chem., Int. Ed. 2011, 50 (46), 1083910842,  DOI: 10.1002/anie.201105364
  32. 32
    Su, W.; Dou, H.; Li, J.; Huo, D.; Dai, N.; Yang, L. Tuning Photoluminescence of Single-Layer MoS2 Using H2O2. RSC Adv. 2015, 5 (101), 8292482929,  DOI: 10.1039/C5RA12450F
  33. 33
    Dong, L.; Lin, S.; Yang, L.; Zhang, J.; Yang, C.; Yang, D.; Lu, H. Spontaneous Exfoliation and Tailoring of MoS2 in Mixed Solvents. Chem. Commun. 2014, 50 (100), 1593615939,  DOI: 10.1039/C4CC07238C
  34. 34
    Pradhan, N. R.; Rhodes, D.; Zhang, Q.; Talapatra, S.; Terrones, M.; Ajayan, P. M.; Balicas, L. Intrinsic Carrier Mobility of Multi-Layered Mos2 Field-Effect Transistors on SiO2. Appl. Phys. Lett. 2013, 102 (12), 123105,  DOI: 10.1063/1.4799172
  35. 35
    Laskar, M. R.; Nath, D. N.; Ma, L.; Lee, E. W.; Lee, C. H.; Kent, T.; Yang, Z.; Mishra, R.; Roldan, M. A.; Idrobo, J.-C.; Pantelides, S. T.; Pennycook, S. J.; Myers, R. C.; Wu, Y.; Rajan, S. P-Type Doping of MoS2 Thin Films Using Nb. Appl. Phys. Lett. 2014, 104 (9), 092104,  DOI: 10.1063/1.4867197
  36. 36
    Werner, F. Hall Measurements on Low-Mobility Thin Films. J. Appl. Phys. 2017, 122 (13), 135306,  DOI: 10.1063/1.4990470
  37. 37
    Rai, A.; Movva, H. C. P.; Roy, A.; Taneja, D.; Chowdhury, S.; Banerjee, S. K. Progress in Contact, Doping and Mobility Engineering of MoS2: An Atomically Thin 2D Semiconductor. Crystals 2018, 8 (8), 316,  DOI: 10.3390/cryst8080316
  38. 38
    Acerce, M.; Voiry, D.; Chhowalla, M. Metallic 1T Phase MoS2 Nanosheets as Supercapacitor Electrode Materials. Nat. Nanotechnol. 2015, 10 (4), 313318,  DOI: 10.1038/nnano.2015.40
  39. 39
    Attanayake, N. H.; Thenuwara, A. C.; Patra, A.; Aulin, Y. V.; Tran, T. M.; Chakraborty, H.; Borguet, E.; Klein, M. L.; Perdew, J. P.; Strongin, D. R. Effect of Intercalated Metals on the Electrocatalytic Activity of 1T-MoS2 for the Hydrogen Evolution Reaction. ACS Energy Letters 2018, 3 (1), 713,  DOI: 10.1021/acsenergylett.7b00865
  40. 40
    Yin, Y.; Han, J.; Zhang, Y.; Zhang, X.; Xu, P.; Yuan, Q.; Samad, L.; Wang, X.; Wang, Y.; Zhang, Z.; Zhang, P.; Cao, X.; Song, B.; Jin, S. Contributions of Phase, Sulfur Vacancies, and Edges to the Hydrogen Evolution Reaction Catalytic Activity of Porous Molybdenum Disulfide Nanosheets. J. Am. Chem. Soc. 2016, 138 (25), 79657972,  DOI: 10.1021/jacs.6b03714
  41. 41
    Scanlon, D. O.; Watson, G. W.; Payne, D. J.; Atkinson, G. R.; Egdell, R. G.; Law, D. S. L. Theoretical and Experimental Study of the Electronic Structures of MoO3 and MoO2. J. Phys. Chem. C 2010, 114 (10), 46364645,  DOI: 10.1021/jp9093172
  42. 42
    Afanasiev, P.; Lorentz, C. Oxidation of Nanodispersed Mos2 in Ambient Air: The Products and the Mechanistic Steps. J. Phys. Chem. C 2019, 123 (12), 74867494,  DOI: 10.1021/acs.jpcc.9b01682
  43. 43
    Ziembowicz, S.; Kida, M.; Koszelnik, P. Sonochemical Formation of Hydrogen Peroxide. Proceedings 2018, 2 (5), 188,  DOI: 10.3390/ecws-2-04957
  44. 44
    Riesz, P.; Kondo, T. Free Radical Formation Induced by Ultrasound and Its Biological Implications. Free Radical Biol. Med. 1992, 13 (3), 247270,  DOI: 10.1016/0891-5849(92)90021-8
  45. 45
    Hu, X. K.; Qian, Y. T.; Song, Z. T.; Huang, J. R.; Cao, R.; Xiao, J. Q. Comparative Study on MoO3 and HxMoO3 Nanobelts: Structure and Electric Transport. Chem. Mater. 2008, 20 (4), 15271533,  DOI: 10.1021/cm702942y
  46. 46
    Ou, J. Z.; Campbell, J. L.; Yao, D.; Wlodarski, W.; Kalantar-Zadeh, K. In Situ Raman Spectroscopy of H2 Gas Interaction with Layered MoO3. J. Phys. Chem. C 2011, 115 (21), 1075710763,  DOI: 10.1021/jp202123a
  47. 47
    Balendhran, S.; Deng, J.; Ou, J. Z.; Walia, S.; Scott, J.; Tang, J.; Wang, K. L.; Field, M. R.; Russo, S.; Zhuiykov, S.; Strano, M. S.; Medhekar, N.; Sriram, S.; Bhaskaran, M.; Kalantar-Zadeh, K. Enhanced Charge Carrier Mobility in Two-Dimensional High Dielectric Molybdenum Oxide. Adv. Mater. 2013, 25 (1), 109114,  DOI: 10.1002/adma.201203346
  48. 48
    Zhang, H.; Wang, H.; Yang, Y.; Hu, C.; Bai, Y.; Zhang, T.; Chen, W.; Yang, S. Hx moo3-Y Nanobelts: An Excellent Alternative to Carbon Electrodes for High Performance Mesoscopic Perovskite Solar Cells. J. Mater. Chem. A 2019, 7 (4), 14991508,  DOI: 10.1039/C8TA10892G
  49. 49
    Borgschulte, A.; Sambalova, O.; Delmelle, R.; Jenatsch, S.; Hany, R.; Nüesch, F. Hydrogen Reduction of Molybdenum Oxide at Room Temperature. Sci. Rep. 2017, 7 (1), 40761,  DOI: 10.1038/srep40761
  50. 50
    Yang, L.; Zhou, W.; Hou, D.; Zhou, K.; Li, G.; Tang, Z.; Li, L.; Chen, S. Porous Metallic MoO2-Supported MoS2 Nanosheets for Enhanced Electrocatalytic Activity in the Hydrogen Evolution Reaction. Nanoscale 2015, 7 (12), 52035208,  DOI: 10.1039/C4NR06754A
  51. 51
    Lee, C.; Yan, H.; Brus, L. E.; Heinz, T. F.; Hone, J.; Ryu, S. Anomalous Lattice Vibrations of Single- and Few-Layer MoS2. ACS Nano 2010, 4 (5), 26952700,  DOI: 10.1021/nn1003937
  52. 52
    Li, H.; Zhang, Q.; Yap, C. C. R.; Tay, B. K.; Edwin, T. H. T.; Olivier, A.; Baillargeat, D. From Bulk to Monolayer MoS2: Evolution of Raman Scattering. Adv. Funct. Mater. 2012, 22 (7), 13851390,  DOI: 10.1002/adfm.201102111
  53. 53
    Dieterle, M.; Mestl, G. Raman Spectroscopy of Molybdenum Oxides Part II. Resonance Raman Spectroscopic Characterization of the Molybdenum Oxides Mo4O11 and MoO2. Phys. Chem. Chem. Phys. 2002, 4 (5), 822826,  DOI: 10.1039/b107046k
  54. 54
    Zhang, Q.; Li, X.; Ma, Q.; Zhang, Q.; Bai, H.; Yi, W.; Liu, J.; Han, J.; Xi, G. A Metallic Molybdenum Dioxide with High Stability for Surface Enhanced Raman Spectroscopy. Nat. Commun. 2017, 8 (1), 14903,  DOI: 10.1038/ncomms14903
  55. 55
    Castner, D. G.; Hinds, K.; Grainger, D. W. X-Ray Photoelectron Spectroscopy Sulfur 2p Study of Organic Thiol and Disulfide Binding Interactions with Gold Surfaces. Langmuir 1996, 12 (21), 50835086,  DOI: 10.1021/la960465w
  56. 56
    Moonoosawmy, K. R.; Kruse, P. To Dope or Not to Dope: The Effect of Sonicating Single-Wall Carbon Nanotubes in Common Laboratory Solvents on Their Electronic Structure. J. Am. Chem. Soc. 2008, 130 (40), 1341713424,  DOI: 10.1021/ja8036788
  57. 57
    Graf, N.; Yegen, E.; Gross, T.; Lippitz, A.; Weigel, W.; Krakert, S.; Terfort, A.; Unger, W. E. S. XPS and NEXAFS Studies of Aliphatic and Aromatic Amine Species on Functionalized Surfaces. Surf. Sci. 2009, 603 (18), 28492860,  DOI: 10.1016/j.susc.2009.07.029
  58. 58
    Kruse, P. Review on Water Quality Sensors. J. Phys. D: Appl. Phys. 2018, 51 (20), 203002,  DOI: 10.1088/1361-6463/aabb93
  59. 59
    Zubiarrain-Laserna, A.; Kruse, P. Review─Graphene-Based Water Quality Sensors. J. Electrochem. Soc. 2020, 167 (3), 037539,  DOI: 10.1149/1945-7111/ab67a5
  60. 60
    Grahame, D. C. The Electrical Double Layer and the Theory of Electrocapillarity. Chem. Rev. 1947, 41 (3), 441501,  DOI: 10.1021/cr60130a002
  61. 61
    Zafir Mohamad Nasir, M.; Sofer, Z.; Pumera, M. Effect of Electrolyte Ph on the Inherent Electrochemistry of Layered Transition-Metal Dichalcogenides (MoS2, MoSe2, WS2, WSe2). ChemElectroChem 2015, 2 (11), 17131718,  DOI: 10.1002/celc.201500259
  62. 62
    Nishimoto, M.; Muto, I.; Sugawara, Y.; Hara, N. Morphological Characteristics of Trenching around Mns Inclusions in Type 316 stainless Steel: The Role of Molybdenum in Pitting Corrosion Resistance. J. Electrochem. Soc. 2019, 166 (11), C3081C3089,  DOI: 10.1149/2.0131911jes
  63. 63
    Schulman, D. S.; May-Rawding, D.; Zhang, F.; Buzzell, D.; Alem, N.; Das, S. Superior Electro-Oxidation and Corrosion Resistance of Monolayer Transition Metal Disulfides. ACS Appl. Mater. Interfaces 2018, 10 (4), 42854294,  DOI: 10.1021/acsami.7b17660
  64. 64
    Mohtasebi, A.; Kruse, P. Chemical Sensors Based on Surface Charge Transfer. Phys. Sci. Rev. 2018, 3 (2), 20170133,  DOI: 10.1515/psr-2017-0133
  65. 65
    Wan, H.; Xu, L.; Huang, W.-Q.; Zhou, J.-H.; He, C.-N.; Li, X.; Huang, G.-F.; Peng, P.; Zhou, Z.-G. Band Structure Engineering of Monolayer MoS2: A Charge Compensated Codoping Strategy. RSC Adv. 2015, 5 (11), 79447952,  DOI: 10.1039/C4RA12498G
  66. 66
    Donarelli, M.; Ottaviano, L. 2d Materials for Gas Sensing Applications: A Review on Graphene Oxide, MoS2, WS2 and Phosphorene. Sensors 2018, 18 (11), 3638,  DOI: 10.3390/s18113638
  67. 67
    Bazylewski, P.; Van Middelkoop, S.; Divigalpitiya, R.; Fanchini, G. Solid-State Chemiresistors from Two-Dimensional Mos2 Nanosheets Functionalized with L-Cysteine for in-Line Sensing of Part-Per-Billion Cd2+ Ions in Drinking Water. ACS Omega 2020, 5 (1), 643649,  DOI: 10.1021/acsomega.9b03246

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 14 publications.

  1. Nhat M. Ngo, Minh Dang Nguyen, Hung-Vu Tran, Riddhiman Medhi, Jong Moon Lee, T. Randall Lee. Photocatalytic Hydrogen Generation by Monodisperse TiO2 Nanoparticles Singly and Dually Doped with Niobium and Tantalum. ACS Applied Nano Materials 2025, 8 (9) , 4841-4851. https://doi.org/10.1021/acsanm.5c00391
  2. Dipankar Saha, Ayush Bhardwaj, Jiacheng Wang, Varun Pande, Robert Hengstebeck, Peng Bai, James J. Watkins. Probing Electrocatalytic Synergy in Graphene/MoS2/Nickel Networks for Water Splitting through a Combined Experimental and Theoretical Lens. ACS Applied Materials & Interfaces 2024, 16 (32) , 42254-42269. https://doi.org/10.1021/acsami.4c08869
  3. Dipankar Saha, Hsin-Jung Yu, Jiacheng Wang, Prateek, Xiaobo Chen, Chaoyun Tang, Claire Senger, James Nicolas Pagaduan, Reika Katsumata, Kenneth R. Carter, Guangwen Zhou, Peng Bai, Nianqiang Wu, James J. Watkins. Mesoporous Single Atom-Cluster Fe–N/C Oxygen Evolution Electrocatalysts Synthesized with Bottlebrush Block Copolymer-Templated Rapid Thermal Annealing. ACS Applied Materials & Interfaces 2024, 16 (11) , 13729-13744. https://doi.org/10.1021/acsami.3c18693
  4. Dipankar Saha, Johnson Dalmieda, Vinay Patel. Surface-Modified MXenes: Simulation to Potential Applications. ACS Applied Electronic Materials 2023, 5 (6) , 2933-2955. https://doi.org/10.1021/acsaelm.3c00076
  5. Dipankar Saha, Shayan Angizi, Maryam Darestani-Farahani, Johnson Dalmieda, Ponnambalam Ravi Selvaganapathy, Peter Kruse. Tuning the Chemical and Mechanical Properties of Conductive MoS2 Thin Films by Surface Modification with Aryl Diazonium Salts. Langmuir 2022, 38 (12) , 3666-3675. https://doi.org/10.1021/acs.langmuir.1c03061
  6. Shayan Angizi, Eugene Yat Chun Yu, Johnson Dalmieda, Dipankar Saha, P. Ravi Selvaganapathy, Peter Kruse. Defect Engineering of Graphene to Modulate pH Response of Graphene Devices. Langmuir 2021, 37 (41) , 12163-12178. https://doi.org/10.1021/acs.langmuir.1c02088
  7. Johnson Dalmieda, Ana Zubiarrain-Laserna, Dipankar Saha, Ponnambalam Ravi Selvaganapathy, Peter Kruse. Impact of Surface Adsorption on Metal–Ligand Binding of Phenanthrolines. The Journal of Physical Chemistry C 2021, 125 (38) , 21112-21123. https://doi.org/10.1021/acs.jpcc.1c04509
  8. M. Humaun Kabir, Darrius Dias, SMH Marjuban, Mohd Avais, Homero Castaneda, Hong Liang. Effects of urea-functionalized MoS2 on hydrophilic lubrication. Tribology International 2025, 203 , 110384. https://doi.org/10.1016/j.triboint.2024.110384
  9. Vasilios I. Georgakilas. Water as Solvent for the Dispersion of 2D Nanostructured Materials. ChemPhysChem 2025, 26 (3) https://doi.org/10.1002/cphc.202400904
  10. A. G. Abd-Elrahim, Deepto Roy, Muhammad Shehroze Malik, Doo-Man Chun. Low-temperature coating of Mn2O3–MoS2 micro-nano-heterostructure anode as an efficient catalyst for water splitting applications. Journal of Materials Science 2024, 59 (17) , 7332-7355. https://doi.org/10.1007/s10853-024-09620-6
  11. Halyna Klym, Ivan Hadzaman, Yuriy Kostiv, Dmytro Chalyy. Electrical properties and thermally induced aging processes in eco-friendly Cu 0.1 Ni 0.1 Co 1.6 Mn 1.2 O 4 and Cu 0.1 Ni 0.8 Co 0.2 Mn 1.9 O 4 -based thick-film structures for multisensor systems. Molecular Crystals and Liquid Crystals 2023, 767 (1) , 42-50. https://doi.org/10.1080/15421406.2023.2224987
  12. Lina N. Khandare, Dattatray J. Late, Nandu B. Chaure. Engineered two-dimensional molybdenum disulfide/carbonaceous hybrid nanostructures for supercapacitors development. Journal of Energy Storage 2023, 74 , 109336. https://doi.org/10.1016/j.est.2023.109336
  13. Ana Zubiarrain-Laserna, Shayan Angizi, Md Ali Akbar, Ranjith Divigalpitiya, Ponnambalam Ravi Selvaganapathy, Peter Kruse. Detection of free chlorine in water using graphene-like carbon based chemiresistive sensors. RSC Advances 2022, 12 (4) , 2485-2496. https://doi.org/10.1039/D1RA08264G
  14. Dipankar Saha, Vinay Patel, Ponnambalam Ravi Selvaganapathy, Peter Kruse. Facile fabrication of conductive MoS 2 thin films by sonication in hot water and evaluation of their electrocatalytic performance in the hydrogen evolution reaction. Nanoscale Advances 2021, 4 (1) , 125-137. https://doi.org/10.1039/D1NA00456E

ACS Applied Nano Materials

Cite this: ACS Appl. Nano Mater. 2020, 3, 11, 10864–10877
Click to copy citationCitation copied!
https://doi.org/10.1021/acsanm.0c02135
Published November 6, 2020

Copyright © 2020 American Chemical Society. This publication is licensed under these Terms of Use.

Article Views

1848

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Schematic representation of the exfoliation procedure of c-MoS2. Steps are (a) bulk MoS2 sonication in ethanol/water mixture for 12 h; (b) exfoliated 2H-MoS2 suspended in ethanol/water; (c–e) two-stage centrifugation process to collect exfoliated 2H-MoS2; (f) exfoliated 2H-MoS2 as precipitate; (g) exfoliated 2H-MoS2 shaken in 0.06% aqueous hydrogen peroxide; (h) suspended c-MoS2 after sonication; (i–k) two-stage centrifugation process to collect exfoliated c-MoS2; and (l) exfoliated c-MoS2 as precipitate after centrifugation.

    Figure 2

    Figure 2. Morphology of MoS2 solid films and sheet resistance as a function of process parameters. (a) Actual image of exfoliated c-MoS2 (gray patch, 7 × 7 mm2) on a SiO2 substrate (1 × 1 cm2) with Au contacts in the four corners of the substrate. (b) SEM images showing overall film distribution of final exfoliated c-MoS2 on the substrate with high magnification showing few-layer flakes of material. The scale bars on the images represent 10 μm and 100 nm, respectively. (c) Alicona optical microscope mapping image of a SiO2 substrate with c-MoS2 film edge (red line is 19.7 mm long). (d) Height profile at the location of the red line in (d). (e) Relationship between the sheet resistance of c-MoS2 samples sonicated for different times in 0.06% and 0.22% aqueous H2O2. Lines are drawn to guide the eye. Solid lines are for 0.22%. The dotted lines are for 0.06%.

    Figure 3

    Figure 3. TEM images showing film distribution of different MoS2 flasks: (a) bulk MoS2, (b) exfoliated 2H-MoS2, (c) bulk MoS2 treated with 0.06% aqueous H2O2, and (d) exfoliated MoS2 treated with 0.06% aqueous H2O2. The scale bar on the images represents 0.2 μm.

    Figure 4

    Figure 4. XPS high-resolution spectra of exfoliated 2H-MoS2; shaken c-MoS2; 2 min sonicated c-MoS2; and 20 min sonicated c-MoS2 samples. Spectra are (a) Mo 3d, (b) S 2p in nonfunctionalized MoS2, and (c) S 2p in cysteamine-functionalized MoS2.

    Figure 5

    Figure 5. Raman spectra of MoS2 samples. (a) Raman spectra proving the doping effect of H2O2 on 2H-MoS2. Black and violet color curves represent c-MoS2 (20 min sonicated) and 2H-MoS2, respectively, showing the shift in the E12g and A1g modes. (b) Raman spectrum of 2H-MoS2 sample. (c) Raman spectrum of c-MoS2 sample shaken in 0.06% H2O2. (d) Raman spectrum of 20 min sonicated c-MoS2 sonicated sample. All spectra were recorded with a 633 nm laser at 1% power. Each spectrum is normalized based on the highest peak (∼466 cm–1).

    Figure 6

    Figure 6. Schematic representation of structure and chemical composition of the formation of c-MoS2.

    Figure 7

    Figure 7. Schematic representation of the fabrication and potential applications of exfoliated c-MoS2.

    Figure 8

    Figure 8. pH sensing response of c-MoS2 chemiresistive devices. 100 mV potential bias was applied across the c-MoS2 film to measure the current changes. Cysteamine-functionalized c-MoS2 chemiresisitive pH responses at (a) 3.7, (b) 6.5, (c) 4, and (d) 3.5. (e) Pristine c-MoS2 chemiresistive response to different pH values between 4 and 6.5. (f) Cysteamine-functionalized c-MoS2 chemiresistive response to different pH values between 3 and 6.5. (g) Calibration curve (linear fitting) of functionalized c-MoS2 chemiresistive device response (b). (h) Calibration curve from (g) replotted as a function of pH.

    Figure 9

    Figure 9. Schematic representation of protonation and deprotonation process of cysteamine-functionalized c-MoS2.

  • References


    This article references 67 other publications.

    1. 1
      Dragoman, M.; Dinescu, A.; Dragoman, D. 2D Materials Nanoelectronics: New Concepts, Fabrication, Characterization from Microwaves up to Optical Spectrum. Phys. Status Solidi A 2019, 216 (8), 1800724,  DOI: 10.1002/pssa.201800724
    2. 2
      Akinwande, D.; Petrone, N.; Hone, J. Two-Dimensional Flexible Nanoelectronics. Nat. Commun. 2014, 5 (1), 5678,  DOI: 10.1038/ncomms6678
    3. 3
      Chhowalla, M.; Shin, H. S.; Eda, G.; Li, L.-J.; Loh, K. P.; Zhang, H. The Chemistry of Two-Dimensional Layered Transition Metal Dichalcogenide Nanosheets. Nat. Chem. 2013, 5 (4), 263275,  DOI: 10.1038/nchem.1589
    4. 4
      Eftekhari, A. Tungsten Dichalcogenides (WS2, WSe2, and WTe2): Materials Chemistry and Applications. J. Mater. Chem. A 2017, 5 (35), 1829918325,  DOI: 10.1039/C7TA04268J
    5. 5
      Late, D. J.; Liu, B.; Matte, H. S. S. R.; Dravid, V. P.; Rao, C. N. R. Hysteresis in Single-Layer MoS2 Field Effect Transistors. ACS Nano 2012, 6 (6), 56355641,  DOI: 10.1021/nn301572c
    6. 6
      Divigalpitiya, W. M. R.; Morrison, S. R.; Frindt, R. F. Thin Oriented Films of Molybdenum Disulphide. Thin Solid Films 1990, 186 (1), 177192,  DOI: 10.1016/0040-6090(90)90511-B
    7. 7
      Kiriya, D.; Tosun, M.; Zhao, P.; Kang, J. S.; Javey, A. Air-Stable Surface Charge Transfer Doping of MoS2 by Benzyl Viologen. J. Am. Chem. Soc. 2014, 136 (22), 78537856,  DOI: 10.1021/ja5033327
    8. 8
      Sarkar, D.; Liu, W.; Xie, X.; Anselmo, A. C.; Mitragotri, S.; Banerjee, K. MoS2 Field-Effect Transistor for Next-Generation Label-Free Biosensors. ACS Nano 2014, 8 (4), 39924003,  DOI: 10.1021/nn5009148
    9. 9
      Stephenson, T.; Li, Z.; Olsen, B.; Mitlin, D. Lithium Ion Battery Applications of Molybdenum Disulfide (MoS2) Nanocomposites. Energy Environ. Sci. 2014, 7 (1), 209231,  DOI: 10.1039/C3EE42591F
    10. 10
      Cao, L.; Yang, S.; Gao, W.; Liu, Z.; Gong, Y.; Ma, L.; Shi, G.; Lei, S.; Zhang, Y.; Zhang, S.; Vajtai, R.; Ajayan, P. M. Direct Laser-Patterned Micro-Supercapacitors from Paintable MoS2 Films. Small 2013, 9 (17), 29052910,  DOI: 10.1002/smll.201203164
    11. 11
      Deng, Z. H.; Li, L.; Ding, W.; Xiong, K.; Wei, Z. D. Synthesized Ultrathin MoS2 Nanosheets Perpendicular to Graphene for Catalysis of Hydrogen Evolution Reaction. Chem. Commun. 2015, 51 (10), 18931896,  DOI: 10.1039/C4CC08491H
    12. 12
      Tsai, M.-L.; Su, S.-H.; Chang, J.-K.; Tsai, D.-S.; Chen, C.-H.; Wu, C.-I.; Li, L.-J.; Chen, L.-J.; He, J.-H. Monolayer MoS2 Heterojunction Solar Cells. ACS Nano 2014, 8 (8), 83178322,  DOI: 10.1021/nn502776h
    13. 13
      Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.; Chhowalla, M. Photoluminescence from Chemically Exfoliated MoS2. Nano Lett. 2011, 11 (12), 51115116,  DOI: 10.1021/nl201874w
    14. 14
      Xia, J.; Wang, J.; Chao, D.; Chen, Z.; Liu, Z.; Kuo, J.-L.; Yan, J.; Shen, Z. X. Phase Evolution of Lithium Intercalation Dynamics in 2H-MoS2. Nanoscale 2017, 9 (22), 75337540,  DOI: 10.1039/C7NR02028G
    15. 15
      Geng, X.; Sun, W.; Wu, W.; Chen, B.; Al-Hilo, A.; Benamara, M.; Zhu, H.; Watanabe, F.; Cui, J.; Chen, T.-p. Pure and Stable Metallic Phase Molybdenum Disulfide Nanosheets for Hydrogen Evolution Reaction. Nat. Commun. 2016, 7 (1), 10672,  DOI: 10.1038/ncomms10672
    16. 16
      Dabral, A.; Lu, A. K. A.; Chiappe, D.; Houssa, M.; Pourtois, G. A Systematic Study of Various 2D Materials in the Light of Defect Formation and Oxidation. Phys. Chem. Chem. Phys. 2019, 21 (3), 10891099,  DOI: 10.1039/C8CP05665J
    17. 17
      Xie, Y.; Liang, F.; Chi, S.; Wang, D.; Zhong, K.; Yu, H.; Zhang, H.; Chen, Y.; Wang, J. Defect Engineering of Mos2 for Room-Temperature Terahertz Photodetection. ACS Appl. Mater. Interfaces 2020, 12 (6), 73517357,  DOI: 10.1021/acsami.9b21671
    18. 18
      Saha, D.; Kruse, P. Editors’ Choice─Review─Conductive Forms of MoS2 and Their Applications in Energy Storage and Conversion. J. Electrochem. Soc. 2020, 167 (12), 126517,  DOI: 10.1149/1945-7111/abb34b
    19. 19
      Kc, S.; Longo, R. C.; Addou, R.; Wallace, R. M.; Cho, K. Impact of Intrinsic Atomic Defects on the Electronic Structure of MoS2 Monolayers. Nanotechnology 2014, 25 (37), 375703,  DOI: 10.1088/0957-4484/25/37/375703
    20. 20
      McDonnell, S.; Addou, R.; Buie, C.; Wallace, R. M.; Hinkle, C. L. Defect-Dominated Doping and Contact Resistance in MoS2. ACS Nano 2014, 8 (3), 28802888,  DOI: 10.1021/nn500044q
    21. 21
      Sim, D. M.; Kim, M.; Yim, S.; Choi, M.-J.; Choi, J.; Yoo, S.; Jung, Y. S. Controlled Doping of Vacancy-Containing Few-Layer MoS2 Via Highly Stable Thiol-Based Molecular Chemisorption. ACS Nano 2015, 9 (12), 1211512123,  DOI: 10.1021/acsnano.5b05173
    22. 22
      Förster, A.; Gemming, S.; Seifert, G.; Tománek, D. Chemical and Electronic Repair Mechanism of Defects in MoS2 Monolayers. ACS Nano 2017, 11 (10), 99899996,  DOI: 10.1021/acsnano.7b04162
    23. 23
      Kc, S.; Longo, R. C.; Wallace, R. M.; Cho, K. Surface Oxidation Energetics and Kinetics on Mos2 Monolayer. J. Appl. Phys. 2015, 117 (13), 135301,  DOI: 10.1063/1.4916536
    24. 24
      Lu, H.; Kummel, A.; Robertson, J. Passivating the Sulfur Vacancy in Monolayer Mos2. APL Mater. 2018, 6 (6), 066104,  DOI: 10.1063/1.5030737
    25. 25
      Nan, H.; Wang, Z.; Wang, W.; Liang, Z.; Lu, Y.; Chen, Q.; He, D.; Tan, P.; Miao, F.; Wang, X.; Wang, J.; Ni, Z. Strong Photoluminescence Enhancement of MoS2 through Defect Engineering and Oxygen Bonding. ACS Nano 2014, 8 (6), 57385745,  DOI: 10.1021/nn500532f
    26. 26
      Verhagen, T.; Guerra, V. L. P.; Haider, G.; Kalbac, M.; Vejpravova, J. Towards the Evaluation of Defects in MoS2 Using Cryogenic Photoluminescence Spectroscopy. Nanoscale 2020, 12 (5), 30193028,  DOI: 10.1039/C9NR07246B
    27. 27
      Mohtasebi, A.; Broomfield, A. D.; Chowdhury, T.; Selvaganapathy, P. R.; Kruse, P. Reagent-Free Quantification of Aqueous Free Chlorine Via Electrical Readout of Colorimetrically Functionalized Pencil Lines. ACS Appl. Mater. Interfaces 2017, 9 (24), 2074820761,  DOI: 10.1021/acsami.7b03968
    28. 28
      Hoque, E.; Hsu, L. H. H.; Aryasomayajula, A.; Selvaganapathy, P. R.; Kruse, P. Pencil-Drawn Chemiresistive Sensor for Free Chlorine in Water. IEEE Sensors Letters 2017, 1 (4), 14,  DOI: 10.1109/LSENS.2017.2722958
    29. 29
      Dalmieda, J.; Zubiarrain-Laserna, A.; Ganepola, D.; Selvaganapathy, P. R.; Kruse, P. Chemiresistive Detection of Silver Ions in Aqueous Media. Sens. Actuators, B 2021, 328, 129023,  DOI: 10.1016/j.snb.2020.129023
    30. 30
      Gou, P.; Kraut, N. D.; Feigel, I. M.; Bai, H.; Morgan, G. J.; Chen, Y.; Tang, Y.; Bocan, K.; Stachel, J.; Berger, L.; Mickle, M.; Sejdić, E.; Star, A. Carbon Nanotube Chemiresistor for Wireless pH Sensing. Sci. Rep. 2015, 4 (1), 4468,  DOI: 10.1038/srep04468
    31. 31
      Zhou, K.-G.; Mao, N.-N.; Wang, H.-X.; Peng, Y.; Zhang, H.-L. A Mixed-Solvent Strategy for Efficient Exfoliation of Inorganic Graphene Analogues. Angew. Chem., Int. Ed. 2011, 50 (46), 1083910842,  DOI: 10.1002/anie.201105364
    32. 32
      Su, W.; Dou, H.; Li, J.; Huo, D.; Dai, N.; Yang, L. Tuning Photoluminescence of Single-Layer MoS2 Using H2O2. RSC Adv. 2015, 5 (101), 8292482929,  DOI: 10.1039/C5RA12450F
    33. 33
      Dong, L.; Lin, S.; Yang, L.; Zhang, J.; Yang, C.; Yang, D.; Lu, H. Spontaneous Exfoliation and Tailoring of MoS2 in Mixed Solvents. Chem. Commun. 2014, 50 (100), 1593615939,  DOI: 10.1039/C4CC07238C
    34. 34
      Pradhan, N. R.; Rhodes, D.; Zhang, Q.; Talapatra, S.; Terrones, M.; Ajayan, P. M.; Balicas, L. Intrinsic Carrier Mobility of Multi-Layered Mos2 Field-Effect Transistors on SiO2. Appl. Phys. Lett. 2013, 102 (12), 123105,  DOI: 10.1063/1.4799172
    35. 35
      Laskar, M. R.; Nath, D. N.; Ma, L.; Lee, E. W.; Lee, C. H.; Kent, T.; Yang, Z.; Mishra, R.; Roldan, M. A.; Idrobo, J.-C.; Pantelides, S. T.; Pennycook, S. J.; Myers, R. C.; Wu, Y.; Rajan, S. P-Type Doping of MoS2 Thin Films Using Nb. Appl. Phys. Lett. 2014, 104 (9), 092104,  DOI: 10.1063/1.4867197
    36. 36
      Werner, F. Hall Measurements on Low-Mobility Thin Films. J. Appl. Phys. 2017, 122 (13), 135306,  DOI: 10.1063/1.4990470
    37. 37
      Rai, A.; Movva, H. C. P.; Roy, A.; Taneja, D.; Chowdhury, S.; Banerjee, S. K. Progress in Contact, Doping and Mobility Engineering of MoS2: An Atomically Thin 2D Semiconductor. Crystals 2018, 8 (8), 316,  DOI: 10.3390/cryst8080316
    38. 38
      Acerce, M.; Voiry, D.; Chhowalla, M. Metallic 1T Phase MoS2 Nanosheets as Supercapacitor Electrode Materials. Nat. Nanotechnol. 2015, 10 (4), 313318,  DOI: 10.1038/nnano.2015.40
    39. 39
      Attanayake, N. H.; Thenuwara, A. C.; Patra, A.; Aulin, Y. V.; Tran, T. M.; Chakraborty, H.; Borguet, E.; Klein, M. L.; Perdew, J. P.; Strongin, D. R. Effect of Intercalated Metals on the Electrocatalytic Activity of 1T-MoS2 for the Hydrogen Evolution Reaction. ACS Energy Letters 2018, 3 (1), 713,  DOI: 10.1021/acsenergylett.7b00865
    40. 40
      Yin, Y.; Han, J.; Zhang, Y.; Zhang, X.; Xu, P.; Yuan, Q.; Samad, L.; Wang, X.; Wang, Y.; Zhang, Z.; Zhang, P.; Cao, X.; Song, B.; Jin, S. Contributions of Phase, Sulfur Vacancies, and Edges to the Hydrogen Evolution Reaction Catalytic Activity of Porous Molybdenum Disulfide Nanosheets. J. Am. Chem. Soc. 2016, 138 (25), 79657972,  DOI: 10.1021/jacs.6b03714
    41. 41
      Scanlon, D. O.; Watson, G. W.; Payne, D. J.; Atkinson, G. R.; Egdell, R. G.; Law, D. S. L. Theoretical and Experimental Study of the Electronic Structures of MoO3 and MoO2. J. Phys. Chem. C 2010, 114 (10), 46364645,  DOI: 10.1021/jp9093172
    42. 42
      Afanasiev, P.; Lorentz, C. Oxidation of Nanodispersed Mos2 in Ambient Air: The Products and the Mechanistic Steps. J. Phys. Chem. C 2019, 123 (12), 74867494,  DOI: 10.1021/acs.jpcc.9b01682
    43. 43
      Ziembowicz, S.; Kida, M.; Koszelnik, P. Sonochemical Formation of Hydrogen Peroxide. Proceedings 2018, 2 (5), 188,  DOI: 10.3390/ecws-2-04957
    44. 44
      Riesz, P.; Kondo, T. Free Radical Formation Induced by Ultrasound and Its Biological Implications. Free Radical Biol. Med. 1992, 13 (3), 247270,  DOI: 10.1016/0891-5849(92)90021-8
    45. 45
      Hu, X. K.; Qian, Y. T.; Song, Z. T.; Huang, J. R.; Cao, R.; Xiao, J. Q. Comparative Study on MoO3 and HxMoO3 Nanobelts: Structure and Electric Transport. Chem. Mater. 2008, 20 (4), 15271533,  DOI: 10.1021/cm702942y
    46. 46
      Ou, J. Z.; Campbell, J. L.; Yao, D.; Wlodarski, W.; Kalantar-Zadeh, K. In Situ Raman Spectroscopy of H2 Gas Interaction with Layered MoO3. J. Phys. Chem. C 2011, 115 (21), 1075710763,  DOI: 10.1021/jp202123a
    47. 47
      Balendhran, S.; Deng, J.; Ou, J. Z.; Walia, S.; Scott, J.; Tang, J.; Wang, K. L.; Field, M. R.; Russo, S.; Zhuiykov, S.; Strano, M. S.; Medhekar, N.; Sriram, S.; Bhaskaran, M.; Kalantar-Zadeh, K. Enhanced Charge Carrier Mobility in Two-Dimensional High Dielectric Molybdenum Oxide. Adv. Mater. 2013, 25 (1), 109114,  DOI: 10.1002/adma.201203346
    48. 48
      Zhang, H.; Wang, H.; Yang, Y.; Hu, C.; Bai, Y.; Zhang, T.; Chen, W.; Yang, S. Hx moo3-Y Nanobelts: An Excellent Alternative to Carbon Electrodes for High Performance Mesoscopic Perovskite Solar Cells. J. Mater. Chem. A 2019, 7 (4), 14991508,  DOI: 10.1039/C8TA10892G
    49. 49
      Borgschulte, A.; Sambalova, O.; Delmelle, R.; Jenatsch, S.; Hany, R.; Nüesch, F. Hydrogen Reduction of Molybdenum Oxide at Room Temperature. Sci. Rep. 2017, 7 (1), 40761,  DOI: 10.1038/srep40761
    50. 50
      Yang, L.; Zhou, W.; Hou, D.; Zhou, K.; Li, G.; Tang, Z.; Li, L.; Chen, S. Porous Metallic MoO2-Supported MoS2 Nanosheets for Enhanced Electrocatalytic Activity in the Hydrogen Evolution Reaction. Nanoscale 2015, 7 (12), 52035208,  DOI: 10.1039/C4NR06754A
    51. 51
      Lee, C.; Yan, H.; Brus, L. E.; Heinz, T. F.; Hone, J.; Ryu, S. Anomalous Lattice Vibrations of Single- and Few-Layer MoS2. ACS Nano 2010, 4 (5), 26952700,  DOI: 10.1021/nn1003937
    52. 52
      Li, H.; Zhang, Q.; Yap, C. C. R.; Tay, B. K.; Edwin, T. H. T.; Olivier, A.; Baillargeat, D. From Bulk to Monolayer MoS2: Evolution of Raman Scattering. Adv. Funct. Mater. 2012, 22 (7), 13851390,  DOI: 10.1002/adfm.201102111
    53. 53
      Dieterle, M.; Mestl, G. Raman Spectroscopy of Molybdenum Oxides Part II. Resonance Raman Spectroscopic Characterization of the Molybdenum Oxides Mo4O11 and MoO2. Phys. Chem. Chem. Phys. 2002, 4 (5), 822826,  DOI: 10.1039/b107046k
    54. 54
      Zhang, Q.; Li, X.; Ma, Q.; Zhang, Q.; Bai, H.; Yi, W.; Liu, J.; Han, J.; Xi, G. A Metallic Molybdenum Dioxide with High Stability for Surface Enhanced Raman Spectroscopy. Nat. Commun. 2017, 8 (1), 14903,  DOI: 10.1038/ncomms14903
    55. 55
      Castner, D. G.; Hinds, K.; Grainger, D. W. X-Ray Photoelectron Spectroscopy Sulfur 2p Study of Organic Thiol and Disulfide Binding Interactions with Gold Surfaces. Langmuir 1996, 12 (21), 50835086,  DOI: 10.1021/la960465w
    56. 56
      Moonoosawmy, K. R.; Kruse, P. To Dope or Not to Dope: The Effect of Sonicating Single-Wall Carbon Nanotubes in Common Laboratory Solvents on Their Electronic Structure. J. Am. Chem. Soc. 2008, 130 (40), 1341713424,  DOI: 10.1021/ja8036788
    57. 57
      Graf, N.; Yegen, E.; Gross, T.; Lippitz, A.; Weigel, W.; Krakert, S.; Terfort, A.; Unger, W. E. S. XPS and NEXAFS Studies of Aliphatic and Aromatic Amine Species on Functionalized Surfaces. Surf. Sci. 2009, 603 (18), 28492860,  DOI: 10.1016/j.susc.2009.07.029
    58. 58
      Kruse, P. Review on Water Quality Sensors. J. Phys. D: Appl. Phys. 2018, 51 (20), 203002,  DOI: 10.1088/1361-6463/aabb93
    59. 59
      Zubiarrain-Laserna, A.; Kruse, P. Review─Graphene-Based Water Quality Sensors. J. Electrochem. Soc. 2020, 167 (3), 037539,  DOI: 10.1149/1945-7111/ab67a5
    60. 60
      Grahame, D. C. The Electrical Double Layer and the Theory of Electrocapillarity. Chem. Rev. 1947, 41 (3), 441501,  DOI: 10.1021/cr60130a002
    61. 61
      Zafir Mohamad Nasir, M.; Sofer, Z.; Pumera, M. Effect of Electrolyte Ph on the Inherent Electrochemistry of Layered Transition-Metal Dichalcogenides (MoS2, MoSe2, WS2, WSe2). ChemElectroChem 2015, 2 (11), 17131718,  DOI: 10.1002/celc.201500259
    62. 62
      Nishimoto, M.; Muto, I.; Sugawara, Y.; Hara, N. Morphological Characteristics of Trenching around Mns Inclusions in Type 316 stainless Steel: The Role of Molybdenum in Pitting Corrosion Resistance. J. Electrochem. Soc. 2019, 166 (11), C3081C3089,  DOI: 10.1149/2.0131911jes
    63. 63
      Schulman, D. S.; May-Rawding, D.; Zhang, F.; Buzzell, D.; Alem, N.; Das, S. Superior Electro-Oxidation and Corrosion Resistance of Monolayer Transition Metal Disulfides. ACS Appl. Mater. Interfaces 2018, 10 (4), 42854294,  DOI: 10.1021/acsami.7b17660
    64. 64
      Mohtasebi, A.; Kruse, P. Chemical Sensors Based on Surface Charge Transfer. Phys. Sci. Rev. 2018, 3 (2), 20170133,  DOI: 10.1515/psr-2017-0133
    65. 65
      Wan, H.; Xu, L.; Huang, W.-Q.; Zhou, J.-H.; He, C.-N.; Li, X.; Huang, G.-F.; Peng, P.; Zhou, Z.-G. Band Structure Engineering of Monolayer MoS2: A Charge Compensated Codoping Strategy. RSC Adv. 2015, 5 (11), 79447952,  DOI: 10.1039/C4RA12498G
    66. 66
      Donarelli, M.; Ottaviano, L. 2d Materials for Gas Sensing Applications: A Review on Graphene Oxide, MoS2, WS2 and Phosphorene. Sensors 2018, 18 (11), 3638,  DOI: 10.3390/s18113638
    67. 67
      Bazylewski, P.; Van Middelkoop, S.; Divigalpitiya, R.; Fanchini, G. Solid-State Chemiresistors from Two-Dimensional Mos2 Nanosheets Functionalized with L-Cysteine for in-Line Sensing of Part-Per-Billion Cd2+ Ions in Drinking Water. ACS Omega 2020, 5 (1), 643649,  DOI: 10.1021/acsomega.9b03246
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsanm.0c02135.

    • Additional SEM, additional TEM, sheet resistance, Hall measurement, XRD, XPS, Raman, and pH sensor data (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.