Abstract
Understanding the role of macroscopic and atomic defects in the interfacial electron transfer properties of layered transition metal dichalcogenides is important in optimizing their performance in energy conversion and electronic devices. Means of determining the heterogeneous electron transfer rate constant, k, have relied on deliberate exposure of specific electrode regions or additional surface characterization to correlate proposed active sites to voltammetric features. Few studies have investigated the electrochemical activity of surface features of layered dichalcogenides under the same experimental conditions. Herein, MoS2 flakes with well-defined features were mapped using scanning electrochemical microscopy (SECM). At visually flat areas of MoS2, k of hexacyanoferrate(III) ([Fe(CN)6]3−) and hexacyanoferrate(II) ([Fe(CN)6]4−) was typically smaller and spanned a larger range than that of hexaammineruthenium(III) ([Ru(NH3)6]3+), congruent with current literature. However, in contrast to previous studies, reduction of [Fe(CN)6]3− and oxidation of [Fe(CN)6]4− exhibited similar rate constants, attributed to the dominance of charge transfer through surface states. Comparison of SECM with optical and atomic force microscopy images revealed that while most of the flake was electroactive, edge sites associated with freshly exposed areas that include both macrosteps consisting of several monolayers and recessed areas exhibited the highest reactivity, consistent with reported results.
Graphical Abstract

Introduction
Effective use of layered transition metal dichalcogenides, e.g., MoS2, WSe2, and their related heterostructures in energy storage and conversion systems and electronic devices requires understanding charge transfer at the dichalcogenide-electrolyte interface1,2–,3,4,5 and how transfer rates correlate to the number of layers,6,7 atomic defects,8,9 and edge sites.4,10,11 Similar to sp2 carbon materials, dichalcogenides contain covalently bonded atoms in monolayers held together by Van der Waals forces.1,2,12 The identity of the active site13,14 and the role of edge sites and basal planes in catalysis, where electron transfer occurs in-plane and perpendicular to each layer, respectively, have been of much interest. Similar to sp2 carbons, the generally accepted view is that electron transfer is more facile at edge sites than basal planes due to higher electron density and the presence of dangling bonds at edges1,5,12,15,16,17–18,19 although non-zero rate constants have been reported at the basal plane.11,16,20,21
Electrochemical methods such as cyclic voltammetry have provided valuable information about the heterogeneous electron transfer rate constant, k, of electrodes. However, in order to investigate spatial differences in the reactivity of layered dichalcogenides, special sample preparation is required to expose specific regions of the electrode,10,15,20,22 or an ex situ characterization method is needed to correlate the number of proposed active sites to the voltammetric signal.6,9,12 The electrode’s surface may change upon release of potential control and emersion from electrolyte during such measurements, leading to ambiguous results about the true behavior of the electrode surface.
Electrochemical scanning probe methods and droplet cells have been used to overcome these challenges.23 Although several studies have spatially probed sp2 carbon using electron transfer mediators,16,24–,25,26 similar investigations of MoS2 are fewer.7,10,11,22 Recently, a scanning droplet cell was used to measure k at various regions of MoS2.7,11,20 Scanning electrochemical microscopy (SECM)27 showed that strain enhances the hydrogen evolution reaction (HER) at MoS228 and that the conductivity of MoS2 flakes varies widely depending on the nature of the contact.29 Other applications of SECM include measuring the HER kinetics and hydrogen adsorption at amorphous MoS2 on gold30 and mapping the HER at MoS2/graphene oxide composites31 although spatial differences in the active MoS2 areas is unclear due to the high probe currents in the SECM images. Likewise, scanning electrochemical cell microscopy (SECCM) demonstrated that the HER at MoS2 proceeds at the basal planes albeit at much lower rates than edge sites.21
Herein, to demonstrate spatial inhomogeneity in the electrochemical behavior of MoS2 under the same experimental conditions, exfoliated MoS2 flakes supported on an insulator, viz., Si/SiO2, were mapped with sub-micrometer resolution SECM. The electron transfer kinetics of three electron transfer mediators, hexaammineruthenium(III) chloride ([Ru(NH3)6]3+), potassium hexacyanoferrate(III) (ferricyanide, [Fe(CN)6]3−), and potassium hexacyanoferrate(II) (ferrocyanide, [Fe(CN)6]4−) were examined. Au electrical contacts served as fiducial marks for comparison of SECM, optical, atomic force microscopy (AFM), and X-ray photoelectron spectroscopy (XPS) images and facilitated positioning the SECM tip probe using the contrast in electrochemical reactivity between Si/SiO2 and Au.
Experimental Section
Materials.
Chemicals were used as received without purification. Sigma-Aldrich (St. Louis, MO) supplied [Ru(NH3)6]3+ (98 %) and [Fe(CN)6]4− (ACS Reagent, 95.0 % to 102.0 %). [Fe(CN)6]3− (ACS grade) was from Mallinckrodt (Paris, KY). Supporting electrolyte was potassium nitrate (KNO3, 99.0 % min, Alfa Aesar, Ward Hill, MA) or sodium chloride (NaCl, ACS grade, Mallinckrodt). Electrolyte solutions were passed through filters with 0.3 μm pore size (Advantec MFS, Pleasanton, CA). Water (18.3 Ω cm) from an EASYpure UV system (Barnstead-Thermolyne, Dubuque, IA) was used to clean glassware and prepare solutions. Commercial materials and equipment are identified to adequately specify the experimental procedure. In no case does such identification imply recommendation or endorsement by the National Institute of Standards and Technology nor does it imply necessarily that the product is the best available for its purpose.
Au contact fabrication.
Au contacts were made to exfoliated MoS2 flakes using a typical lift-off procedure. Blue cleanroom tape of medium tack was used to cleave flakes from a MoS2 crystal (Lot 1211005, Structure Probe, West Chester, PA) which were gently pressed onto (1 to 2) cm × (1 to 2) cm pieces of Si(100) wafer (n-doped, 0.05 Ω/cm2 to 0.5 Ω/cm2, WRS Materials, San Jose, CA) with 300 nm of thermally grown (1150 °C) SiO2, measured with an ellipsometer (Model M-2000, J. A. Woollam). Wafer pieces were first cleaned in a Model 1020 oxygen plasma cleaner (Fischione Instruments, Export, PA) for 6 min or Piranha solution (4 mL of 18 mol/L H2SO4 + 1 mL of 9.8 mol/L (30 % (w/w) H2O2) for 1 h and rinsed with H2O. Caution: Piranha solutions can be explosive if they contact organic materials.
MoS2 on Si/SiO2, i.e., the substrate, was rinsed with isopropanol, dried with N2, and then dried for 5 min on a hotplate at 115 °C before spin coating with LOR 5A resist (MicroChem, Newton, MA) at 370 rad/s (3500 rpm) for 45 s and then baking at 175 °C for 5 min. To avoid silylation of the MoS2, LOR 5A was used in place of hexamethyldisilazane (HMDS) primer for adhering photoresist to the wafer pieces. AZ5214E image-reversal photoresist (Clarient, Somerville, NJ) was spin coated at 310 rad/s (3000 rpm) for 45 s before prebaking for 1 min at 115 °C. After aligning the chrome photomask (Photo Sciences, Torrance, CA) with the flake in a SussMA6 contact aligner (Suss MicroTech, Garching, Germany), the photoresist was exposed using a dose of 150 mJ/cm2 at 405 nm. Postbake was performed for 2 min at 115 °C before a flood exposure with a dose of 200 mJ/cm2 at 405 nm. Samples were immersed in AZ300 MIF developer (Clarient, Somerville, NJ) for 2 min and then in H2O for 1 min before rinsing with H2O and drying with N2. Then, 10 nm of Ti and 100 nm of Au were deposited using a Denton Vacuum (Moorestown, NJ) Infinity 22 electron beam evaporator. To remove the photoresist, samples were immersed in Remover 1165 (Rohm and Haas Electronics Materials, Marlborough, MA) for 30 min at 60 °C followed by a second remover bath at 30 min at 60 °C or overnight at room temperature. Finally, samples were rinsed with isopropanol and dried with N2. Resistance between the two contact pads was determined from the slope of voltage vs. current measured using the potentiostat.
Electrochemical measurements.
All electrochemical measurements were performed with a CH Instruments (Austin, TX) Model 920D SECM on a marble table inside an aluminum Faraday cage with solid sides to minimize any photochemistry induced from the room lights. At least three separate flakes were used for each measurement. Intervals represent the 95 % confidence level.32 A PTFE electrochemical cell with a volume of 2 mL was used with a Pt or Au wire auxiliary electrode and an Ag|AgCl|1 mol/L KCl (+0.22 V vs. SHE)33 reference electrode in a salt bridge containing 30 g/L agar (ash 2.5 % to 4.5 %, Sigma-Aldrich) + 0.2 mol/L KNO3. For mounting the sample in the cell, Cu tape with In contacts to the Au were covered with electroplater’s tape (3M, St. Paul, MN) with a hole of area ≈1 mm × ≈1 mm to expose the flake to electrolyte. Electrochemical impedance measurements were done in 0.1 mol/L KNO3 on a cleaved crystal piece contacted to Cu tape with a mass fraction of 75/25 Ga/In eutectic on the back. Electroplater’s tape was used to cover the electrode contacts and to expose a crystal area of ≈1 mm × ≈1 mm to solution.
SECM tip probe fabrication.
Pt SECM tips with radius, a, of 5 μm were from CH Instruments and polished with 1.0 μm and then 0.3 μm alumina before use. To make sub-micrometer-sized tips (Figure S1), borosilicate glass capillary tubes (1.0 mm OD, 0.3 mm ID, 7.5 cm long, Sutter Instrument, Novato, CA) and Pt wire (ϕ = 25 μm tempered wire (99.99 %, Goodfellow, Oakdale, PA) were cleaned by soaking overnight in isopropanol, drying at 120 °C for at least 90 min, and then cooling to room temperature immediately before sealing. A 1.5 cm length of Pt wire was inserted into the tube and then sealed using a three-step pulling procedure34,35 (Table S1) in a Model P-2000 pipette puller (Sutter). A 400–14 Two Stage Rotary Vane Vacuum Pump (Vacuum Research, Pittsburgh, PA) was used to evacuate the capillary tubing to <8.0 Pa (60 mtorr) by connecting the ends of the capillary tube and the pump with PTFE tubing.
A Helios NanoLab 660 Dual Beam Scanning Electron Microscope (SEM) and Focused Ion Beam (FIB) (FEI, Hillsboro, OR) system was used to polish the tips34 after final sealing in a MFG-5AP Microforge-Grinding Center (MicroData Instrument, Plainfield, NJ) and coating with 10 nm of Pt using an EM ACE600 Sputter Coater (Leica Microsystems, Buffalo Grove, IL). The ion beam (30 keV, 790 pA) was scanned perpendicular to the end of the tip using a cleaning cross section pattern with 1 ms dwell time. After polishing, back contact was made with Cu wire and a mass fraction of 75/25 Ga/In eutectic. Immediately before use, tips were immersed for 30 s in Piranha solution and then rinsed with H2O.
Equipment.
Optical images were recorded using an Epiphot 300 Metallurgy Inverted Microscope (Nikon, Tokyo, Japan) with a Model OCS-5.0 OptixCam Summit Series CMOS Camera (Microscope Store, Roanoke, VA) and then processed using ImageJ (1.49).36 AFM images were obtained using an Asylum Research (Santa Barbara, CA) Cypher instrument in tapping mode with model AC160T-R3 micro-cantilevers (frequency, 300 Hz; force constant, 26 N/m; radius, 7 nm) from Olympus (Center Valley, PA). Processing with Gwyddion (2.49) software37 included performing a plane fit, fixing the minimum height value to zero, and constraining the height scale bar. Raman spectra were obtained using a LabRAM HR800 (Horiba Scientific) microscope and 50X long working distance objective with an excitation wavelength of 532 nm, laser power of 34 mW, and an average of 10 spectra integrated for 2 s.
Following SECM and AFM measurements, XPS was performed using a Kratos AXIS Ultra DLD with a monochromatic Al Kα source. The spectrometer was calibrated to Au 4f7/2 84.0 eV, Ag 3d5/2 368.2 eV, and Ag MNN 1128.9 eV signals, and all spectra were collected at a 90° take-off angle. Spectra were fit using Casa XPS software (2.3.15) with the peaks modeled using a Voigt function (70 % Gaussian-30 % Lorentizian) and a Shirley background correction.
Results and Discussion
Characterization of substrate.
Lateral dimensions of the flakes (Figure 1) were typically between 30 μm and 100 μm with thickness between 80 nm and 250 nm (>100 monolayers).12 Resistance between the two Au contacts among the samples varied between 1 kΩ and 30 kΩ. In Raman spectra of the flakes (Figure 1b), and A1g peaks at 380 cm−1 and 404 cm−1, respectively, were similar to the bulk crystal and literature values.38
Figure 1.
(a) Optical image of MoS2 with Au contacts. (b) Raman spectra of MoS2 flake and bulk crystal.
Electrochemical impedance data of the bulk crystal were plotted (Figure 2) according to the Mott-Schottky equation at 25 °C:
| (1) |
where C (μF/cm2) is series capacitance, ɛ is dielectric constant, ND (cm−3) is dopant concentration, E (V) is applied potential, and EFB (V) is flatband potential.39 Values of C were calculated from the measured impedance, Zim (Ω), at varying frequency, f (Hz):
| (2) |
Figure 2.
Mott-Schottky plot of MoS2 in 0.1 mol/L KNO3 at varying frequency. Potential step, 0.05 V; amplitude, 0.005 V.
The slope of C–2 vs. E provides ND, and EFB is derived from where the line crosses the potential axis, i.e., the y-intercept, using EFB = E + 0.0257. Using ε = 4.304,40 the MoS2 crystal was n-doped with a moderate carrier concentration39 of (4.1 ± 1.0) × 1016 cm−3, about half the reported value for synthetic n-type MoS2 in 0.5 M KNO3.41 The EFB, ≈ +0.1 V, was similar to that of synthetic MoS2.41
Characterization of SECM tips.
Cyclic voltammetry (Figures S1b, S2a, and S3a) was used to determine the radius, a, of the Pt electrode using the equation iT,inf = 4nFaDC, where iT,inf is tip current under semi-infinite conditions, n is number of electrons, F is Faraday’s constant (96,485 C/mol e−), D is the diffusion coefficient, and C is the concentration of the mediator.39 Reported values of D, 7.2 × 10−6 cm2/s,42,43 7.2 × 10−6 cm2/s,42,43 7.6 × 10−6 cm2/s,39 and 6.5 × 10−6 cm2/s,39 for [Ru(NH3)6]2+, [Ru(NH3)6]3+, [Fe(CN)6]3−, and [Fe(CN)6]4−, respectively, were used. SECM tips were characterized by fitting experimental approach curves, i.e., plot of tip current, iT, vs. tip-substrate separation, d, at Si/SiO2 in ca. 4 mmol/L [Ru(NH3)6]3+ + 0.1 mol/L KNO3 to theory44 to determine the shortest distance between the tip and substrate, d0, and the ratio, RG, of the radius of the glass sheath to the Pt disk (Figures S2b, S2c, S3b, and S3c). Error in d was less than ±1 μm. For SECM mapping, the substrate was first leveled by monitoring iT of the a = 5 μm tip as it was scanned laterally so that the variation in height was less than 1 μm over 1 mm length of the substrate.
Characterization of substrate electron transfer kinetics.
Although the exposed Au contact dominates the voltammetry of the composite MoS2/Au working electrode, electron transfer at MoS2 could be probed directly using the SECM tip, as shown in Figure 3 for sample MoS2 1 listed in Table 1. The tip was poised at a potential, ET, to collect any species generated at the substrate as the substrate potential, ES, was changed to modulate its electrochemical reactivity. This configuration is similar to the substrate generation-tip collection (SG-TC) mode of SECM that also involves feedback between the tip and substrate,27,45 as suggested in the inset of Figure 3. The formal potential, E0’, estimated as E1/2 of the substrate, was −0.17 V and + 0.23 V for [Ru(NH3)6]3+/2+ and [Fe(CN)6]3−/4−, respectively. As expected, iT increased with increasing substrate overpotential, i.e., the difference between E0’ and ES.
Figure 3.
SECM measurements of MoS2 flake with Au contacts using tip (a = 5 μm) as a collector electrode in (blue) 3.6 mmol/L [Ru(NH3)6]3+, (black) 2.9 mmol/L [Fe(CN)6]3−, and (red) 2.9 mmol/L [Fe(CN)6]4−. Supporting electrolyte, 0.1 mol/L KNO3. ET = +0.10 V, +0.40 V, and −0.10 V for [Ru(NH3)6]3+, [Fe(CN)6]3−, and [Fe(CN)6]4−, respectively. d = 6 μm; sweep rate, 10 mV/s. Inset is diagram showing collection for [Fe(CN)6]3−. iT,inf = 4.8 nA, 3.9 nA, and 3.7 nA for [Ru(NH3)6]3+, [Fe(CN)6]3−, and [Fe(CN)6]4−, respectively.
Table 1.
Normalized tip current, I = iT/iT,inf, for hexaammineruthenium(III) ([Ru(NH3)6]3+), hexacyanoferrate(III) ([Fe(CN)6]3−), and hexacyanoferrate(II) ([Fe(CN)6]4−) at separate locations on three MoS2 samples and the Au contacts of MoS2 1. Number in parentheses for MoS2 is the range of values. The intervals at Au are reported at the 95 % confidence level.
| Sample | Location | I, [Ru(NH3)6]3+ | I, [Fe(CN)6]3− | I, [Fe(CN)6]4− |
|---|---|---|---|---|
| MoS2 1 | a | 0.92 | 0.62 | 0.70 |
| MoS2 1 | b | 0.90 | 0.72 | 0.68 |
| MoS2 1 | c | 0.90 | 0.77 | 0.59 |
| Average of 1 | 0.90 (0.02) | 0.70 (0.15) | 0.66 (0.11) | |
| MoS2 2 | a | 1.06 | 0.49 | 0.46 |
| MoS2 2 | b | 0.81 | 0.59 | 0.54 |
| MoS2 2 | c | 0.77 | 0.59 | 0.57 |
| Average of 2 | 0.88 (0.29) | 0.56 (0.10) | 0.52 (0.11) | |
| MoS2 3 | a | 1.04 | 1.05 | 1.15 |
| MoS2 3 | b | Not measured | 0.74 | 0.66 |
| MoS2 3 | c | Not measured | 0.73 | 0.68 |
| MoS2 3 | d | 1.00 | 0.89 | 0.89 |
| Average of 3 | 1.02 (0.04) | 0.85 (0.33) | 0.84 (0.48) | |
| Average of all | 0.92 (0.29) | 0.72 (0.56) | 0.69 (0.69) | |
| Au | 1.74 ± 0.08 | 0.87 ± 0.06 | 0.88 ± 0.01 |
The electron transfer kinetics of the substrate can be inferred from iT,27,44 where an increase in iT is due to increasing k at the substrate electrode as well as positive feedback between the tip and substrate. Although many studies use the feedback mode to determine k, SECM approach curves were shown to be similar for both feedback and SG-TC involving feedback if the ratio of diffusion coefficients of the reduced and oxidized species is near unity.45 Table 1 lists the normalized tip current, I = iT/iT,inf, at d = 6 μm for each mediator at separate, visually flat locations on three different MoS2 flakes and the Au contacts of sample MoS21 as well as average values for each flake and all flakes. For MoS2, the number in parentheses is the range of values. Values of iT were obtained from the steady-state tip current, i.e., where iT is independent of ES (Figures 3 and S4). The slope in the rising tip current as well as the steady-state tip current at Au were higher than those at MoS2 (Figure 3) due to faster electron transfer that gives rise to greater positive feedback at Au. At Au, the lower I of [Fe(CN)6]3− and [Fe(CN)6]4− compared to that of [Ru(NH3)6]3+ is likely due to the greater sensitivity of [Fe(CN)6]3−/4− to surface contamination of Au46,47 (vide infra). At MoS2, I was typically 0.8 to 1.0 for [Ru(NH3)6]3+ and 0.5 to 0.9 for [Fe(CN)6]3−, and [Fe(CN)6]4−. Thus, [Ru(NH3)6]3+ typically exhibited faster kinetics while a wider variation in k across samples and within the same sample were observed for [Fe(CN)6]3− and [Fe(CN)6]4−, similar to trends observed at defective basal planes on bulk MoS2 in the literature.7,10,11 These observations are attributed to outer- versus inner-sphere interactions of [Ru(NH3)6]3+/2+ and [Fe(CN)6]3−/4−, respectively, at the substrate11,15,39,48 as well as the electronic structure of n-MoS2.11,17,49 For [Ru(NH3)6]3+/2+, E0’ is negative of the flat band potential, leading to metallic behavior in contrast to the more positive E0’ of [Fe(CN)6]3−/4−, which may decrease the number of available carriers. Accordingly, oxidation of [Fe(CN)6]4− is expected to be slower than reduction of [Fe(CN)6]3− based on available states17,49 although I was not noticeably different for [Fe(CN)6]3− and [Fe(CN)6]4−. However, because a = 5 μm and the conduction band of MoS2 is near E0’ of [Fe(CN)6]3−/4−,49 iT likely represents the average reactivity of multiple regions containing local differences in doping concentration50 and surface states that may obscure such nuances.
SECM mapping of MoS2.
The magnitude of iT was measured as a function of lateral tip position using the SECM tip as a generator (Figure 4a) or collector (Figures 4b and 4c) electrode (Figure 4d) in the electrolyte with [Ru(NH3)6]3+. When using tips with radii less than 1 μm, relatively large tip heights, e.g., d was 4.5 times the size of a in Figure 4, were used to avoid damaging the substrate or the delicate end of the SECM tip. SECM and optical images from additional MoS2 samples are given in Figures S5, S6, and S7. In addition, Figure S8 shows Figures 4a and 4b plotted using a diverging color scheme. Larger iT values correspond to higher reactivity. Only effects from electronic structure or surface contamination that could change the electron tunneling length should be observed in the case of an outer-sphere mediator, e.g., [Ru(NH3)6]3+.15,39,48 Before the SECM images in Figure 4, MoS2 was mechanically cleaved in the electrolyte and then stored in water overnight (Figure S9). Spatial differences in reactivity were more apparent with the smaller tip (Figure S10), demonstrating the necessity of matching probe size with substrate features. Geometric features in optical (Figures 4e, S5e, and S6f) and AFM (Figures 4f and S11) images corresponded well to variations in iT. Most of the MoS2 was electroactive, similar to previous results,11,16,21 indicated by increasing iT with increasing substrate overpotential (Figures 4b and 4c, Figures S5c and S5d, Figure S6c, S6d, and S6e). Due to decreased overlap between the tip and MoS2 at the flake perimeter, iT was lower in these areas. At macrosteps consisting of several monolayers (region A, Figure 4), iT was larger, suggesting that electron transfer was more facile at these edges, similar to SECM of graphene16,24 as well as scanning photoluminescence51 and STM18 of MoS2. The lowest reactivity was observed on basal plane areas (region E, Figure 4), particularly those that may not have been disturbed during specimen cleavage (Figure S9). Changes in topography may affect iT, e.g., a taller feature would increase iT due to a smaller tip-substrate distance52 (region B, Figures 4 and S11d). However, a 2 μm-wide, 1 μm-deep trench (region C, Figures 4 and S11b) was more active than the surrounding region, reflecting a high density of edge sites in the sidewalls. Observed reactivity at the nominally flat region D in Figures 4 and S11b corresponded to an unambiguously freshly cleaved basal surface (Figure S9) although the effect of defects8,9,20 smaller than the resolution of the current system must be considered. For comparison, SECM images with [Fe(CN)6]3− (Figure S12) were obtained for the same substrate overpotentials, i.e., 80 mV and 180 mV, as with [Ru(NH3)6]3+ (Figure 3b, 3c). Like [Ru(NH3)6]3+, most of the MoS2 was electroactive, and reactivity across the surface became more homogeneous with larger substrate overpotentials.
Figure 4.
SECM images of MoS2 using the tip (a = 0.22 μm) as (a) generator and (b,c) collector electrode in 4.4 mmol/L [Ru(NH3)6]3+ + 0.1 mol/L KNO3. Current scale bar for (b) and (c) is the same. ET = (a) –0.45 V, (b,c) +0.10 V. ES = (a) +0.10 V, (b) –0.25 V, (c) –0.35 V. Scan rate, 2.5 μm/s; step size, 0.5 μm; d = 1 μm. (d) Diagram showing feedback between tip and substrate, where Ru(III) and Ru(II) represent [Ru(NH3)6]3+ and [Ru(NH3)6]2+, respectively. (e) Optical and (f) AFM images (scan rate, 1 Hz) after electrochemical measurements. Height difference between Au and MoS2 is 80 nm.
XPS characterization of substrate after SECM and AFM measurements.
The distribution of Au, Mo, S, C, and O on the sample presented in Figures 4, S9, S11, and S12 was mapped using the spherical mirror analyzer for parallel XPS imaging and analysis (Figures 5 and S13). The time delay between SECM and AFM measurements was ≈60 days while the specimen was transferred to the XPS immediately after the AFM study. The Au contacts on the SiO2 surface were clearly revealed in Figure S13a, and the gap between the contacts was filled in the Mo 3d and S 2p maps although the intensity of the latter is much weaker due to the photoelectron sensitivity factors, 0.668 for S 2p versus 3.321 for Mo 3d. The Au contacts were evident in several of the elemental maps due to contribution of strong inelastic scattering from Au 4f and Si 2p and Si 2s. Nevertheless, a color scaled overlay of the respective spectral maps clearly revealed MoS2 between the Au contact pads in Figure S14. The O 1s map (Figures 5a and S13e) showed negligible oxygen on MoS2 while the Au contact was extensively covered by oxygen-containing species, making them indistinguishable from the surrounding SiO2. Brief Ar+ sputtering removed this adventitious layer from Au (Figures 5b and S13f). Despite the surface contamination, [Ru(NH3)6]3+/2+ at Au was still electrochemically reversible (Figure 3). In contrast to oxygen, some carbon was evident on the MoS2 before sputtering. Comparison of MoS2 before and after cleavage (Figure S9) along with the XPS oxygen and carbon maps (Figure S15) indicated that the freshly exposed surface regions were the most active, namely region D, the recessed trench C, and the macrostep edge A. As evident from Figure S9, the cleanliness of certain areas of the MoS2 relative to Au may be ascribed to the creation of a fresh MoS2 surface caused by tip collision with the MoS2 that resulted in formation of the recessed trench and substantial delamination of the surrounding MoS2 surface. The state of the most active sections of the MoS2 surface in Figure 4 is thus analogous to a freshly cleaved surface like that studied by Dryfe and others11 while the Au electrode contact remained covered by an adventitious layer related to extended exposure to laboratory ambient conditions and possibly lithographic processing.11,53
Figure 5.
O 1s map of the sample in Figure 4 (a) before and (b) after sputtering with 500 eV Ar+ for 8 minutes. (c) Corresponding optical image.
Conclusions
Although our current measurement system is unable to resolve atomic scale features, SECM mapping of MoS2 flakes with sub-micrometer resolution using electron transfer mediators demonstrated that electrochemical reactivity was highest at visible physical defects and freshly cleaved surfaces, congruent with reported voltammetric measurements. Attention to surface topology is important when interpreting SECM images of MoS2. Combination of scanning probe functionalities such as AFM-SECM with advances in fabrication of layered materials and related heterostructures with specified defects, e.g., step edges and atomic vacancies, will provide an excellent platform to unravel the connection between charge transfer kinetics and surface structure.
Supplementary Material
Acknowledgments
We thank Jey Velmurugan, Joshua Schumacher, and Amit Agrawal for advice on tip fabrication, Debnath Ratan, Asha Rani, Liya Yu, and Rob Ilic for discussions on contact fabrication, Robert Newby for preparing Si/SiO2 wafers, Andrei Kolmakov for use of the sputter coater, Joey Robertson for access to the microforge, and Heather Hill for advice on transferring MoS2. N.L.R. was supported by the U.S. Department of Commerce, National Institute of Standards and Technology (NIST), under financial assistance award 70NANB16H294 and the Cooperative Research Agreement between the University of Maryland and the NIST Center for Nanoscale Science and Technology, Award 70NANB10H193, through the University of Maryland.
Footnotes
Supporting Information
Sealing parameters for SECM tips. SEM image and cyclic voltammetry of SECM tip. Cyclic voltammetry and approach curves for SECM tips. SECM images of additional MoS2 flakes in [Ru(NH3)6]3+. Comparison of SECM images using rainbow and diverging color maps. Optical image of cleaved MoS2 flake. SECM images with different tip sizes. AFM images of MoS2 with line profiles. SECM images with [Fe(CN)6]3−. XPS images of substrate. Overlaid, colored XPS images of substrate. Comparison of optical and XPS images of MoS2. This material is available free of charge via the Internet at http://pubs.acs.org.
References
- (1).Tributsch H and Bennett JC. Electrochemistry and Photochemistry of MoS2 Layer Crystals. I. J. Electroanal. Chem 1977, 81, 97–111. [Google Scholar]
- (2).Tributsch H Photoelectrochemical Behaviour of Layer-Type Transition Metal Dichalcogenides. Faraday Discuss. Chem. Soc 1980, 70, 189–205. [Google Scholar]
- (3).Laurson AB; Kegnaes S; Dahl S; Chorkendorff I Molybdenum Sulfides-Efficient and Viable Materials for Electro- and Photoelectrocatalytic Hydrogen Evolution. Energy Environ. Sci 2012, 5, 5577–5591. [Google Scholar]
- (4).Benck JD; Hellstern TR; Kibsgaard J; Chakthranont P; Jaramillo TF Catalyzing the Hydrogen Evolution Reaction (HER) with Molybdenum Sulfide Nanomaterials. ACS Catal. 2014, 4, 3957–3971. [Google Scholar]
- (5).Velický M and Toth PS. From Two-Dimensional Materials to Their Heterostructures: An Electrochemist’s Perspective. Appl. Mater. Today 2017, 8, 68–103. [Google Scholar]
- (6).Yu Y; Huang S-Y; Li Y; Steinmann SN; Yang W; Cao L Layer-Dependent Electrocatalysis of MoS2 for Hydrogen Evolution. Nano Lett. 2014, 14, 553–558. [DOI] [PubMed] [Google Scholar]
- (7).Velický M; Bissett MA; Woods CR; Toth PS; Georgiou T; Kinlock IA; Novoselov KS; Dryfe RAW Photoelectrochemistry of Pristine Mono- and Few-Layer MoS2. Nano Lett. 2016, 16, 2023–2032. [DOI] [PubMed] [Google Scholar]
- (8).Li H; Tsai S; Koh AL; Cai L; Contryman AW; Fragapane AH; Zhao J; Han HS; Manoharan HC; Abild-Pederson F; Norskov JK; Zheng X Activating and Optimizing MoS2 Basal Planes for Hydrogen Evolution through the Formation of Strained Sulphur Vacancies. Nat. Mater 2016, 15, 48–53. [DOI] [PubMed] [Google Scholar]
- (9).Li G; Zhang D; Qiao Q; Yu Y; Peterson D; Zafar A; Kumar R; Curtarolo S; Hunte F; Shannon S; Zhu Y; Yang W; Cao L All the Catalytic Active Sites of MoS2 for Hydrogen Evolution. J. Am. Chem. Soc 2016, 138, 16632–16638. [DOI] [PubMed] [Google Scholar]
- (10).Tan SM; Ambrosi A; Sofer Z; Huber S; Sedmidubský D; Pumera M Pristine Basal- and Edge-Plane-Oriented Molybdenite MoS2 Exhibiting Highly Anisotropic Properties. Chem. Eur. J 2015, 21, 7170–7178. [DOI] [PubMed] [Google Scholar]
- (11).Velický M; Bissett MA; Toth PS; Patten HV; Worrall SD; Rodger ANJ; Hill EW; Kinloch IA; Novoselov KS; Georgiou T; Britnell L; Dryfe RAW Electron Transfer Kinetics on Natural Crystals of MoS2 and Graphite. Phys. Chem. Chem. Phys 2015, 17, 17844–17853. [DOI] [PubMed] [Google Scholar]
- (12).Jaegermann W and Tributsch H. Interfacial Properties of Semiconducting Transition Metal Chalcogenides. Prog. Surf. Sci 1988, 29, 1–167. [Google Scholar]
- (13).Hinnemann B; Moses PG; Bonde J; Jorgensen KP; Nielson JH; Horch S; Chorkendorff I; Norskov JK Biomimetic Hydrogen Evolution: MoS2 Nanoparticles as Catalyst for Hydrogen Evolution. J. Am. Chem. Soc 2005, 127, 5308–5309. [DOI] [PubMed] [Google Scholar]
- (14).Huang Y; Nielson RJ; Goddard WA; Soriaga MP The Reaction Mechanism with Free Energy Barriers for Electrochemical Dihydrogen Evolution on MoS2. J. Am. Chem. Soc 2015, 137, 6692–6698. [DOI] [PubMed] [Google Scholar]
- (15).McCreery RL Advanced Carbon Electrode Materials for Molecular Electrochemistry. Chem. Rev 2008, 108, 2646–2687. [DOI] [PubMed] [Google Scholar]
- (16).Unwin PR; Güell AG; Zhang G Nanoscale Electrochemistry of sp2 Carbon Materials: from Graphite and Graphene to Carbon Nanotubes. Acc. Chem. Res 2016, 49, 2041–2048. [DOI] [PubMed] [Google Scholar]
- (17).Ahmed SM and Gerischer H. Influence of Crystal Surface Orientation on Redox Reactions at Semiconducting MoS2. Electrochim. Acta 1979, 24, 705–711. [Google Scholar]
- (18).Jaramillo TF; Jorgensen KP; Bonde J; Nielson JH; Horch S; Chorkendorff I Identification of Active Edge Sites for Electrochemical H2 Evolution from MoS2 Nanocatalysts. Science 2007, 317, 100–102. [DOI] [PubMed] [Google Scholar]
- (19).Chen Z; Forman AJ; Jaramillo TF Bridging the Gap Between Bulk and Nanostructured Photoelectrodes: The Impact of Surface States on the Electrocatalytic and Photoelectrochemical Properties of MoS2. J. Phys. Chem. C 2013, 117, 9713–9722. [Google Scholar]
- (20).Voiry D; Fullon R; Yang J; de Carvalho Castro e Silva C; Kappera R; Bozkurt I; Kaplan D; Lagos MJ; Baston PE; Gupta G; Mohite AD; Dong L; Er D; Shenoy VB; Asefa T; Chhowalla M Role of Electronic Coupling between Substrate and 2D MoS2 Nanosheets in Electrocatalytic Production of Hydrogen. Nat. Mater 2016, 15, 1003–1009. [DOI] [PubMed] [Google Scholar]
- (21).Bentley CL; Kang M; Maddar FM; Li F; Walker M; Zhang J; Unwin PR Electrochemical Maps and Movies of the Hydrogen Evolution Reaction on Natural Crystals of Molybdenite (MoS2): Basal vs. Edge Plane Activity. Chem. Sci 2017, 8, 6583–6593. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).Schneemeyer LF and Wrighton MS. Flat-Band Potential of n-Type Semiconducting Molybdenum Disulfide by Cyclic Voltammetry of Two-Electron Reductants: Interface Energetics and the Sustained Photooxidation of Chloride. J. Am. Chem. Soc 1979, 101, 6496–6500. [Google Scholar]
- (23).Jaouen K; Henrotte O; Campidelli S; Jousselme B; Derycke V; Cornut R Localized Electrochemistry for the Investigation and the Modification of 2D Materials. Appl. Mater. Today 2017, 8, 116–124. [Google Scholar]
- (24).Ritzert NL; Rodríguez-López J; Tan C; Abruña HD Kinetics of Interfacial Electron Transfer at Single-Layer Graphene Electrodes in Aqueous and Nonaqueous Solutions Langmuir 2013, 29, 1683–1695. [DOI] [PubMed] [Google Scholar]
- (25).Tan C; Rodríguez-López J; Parks JJ; Ritzert NL; Abruña HD Reactivity of Monolayer Chemical Vapor Deposited Graphene Imperfections Studied Using Scanning Electrochemical Microscopy. ACS Nano 2012, 6, 3070–3079. [DOI] [PubMed] [Google Scholar]
- (26).Zhong J-H; Zhang J; Jin X; Liu J-X; Li Q; Li M-H; Cai W; Wu D-Y; Zhan D; Ren B Quantitative Correlation between Defect Density and Heterogeneous Electron Transfer Rate of Single Layer Graphene. J. Am. Chem. Soc 2014, 136, 16609–16617. [DOI] [PubMed] [Google Scholar]
- (27).Amemiya S; Bard AJ; Fan F-RF; Mirkin MV; Unwin PR Scanning Electrochemical Microscopy. Annu. Rev. Anal. Chem 2008, 1, 95–131. [DOI] [PubMed] [Google Scholar]
- (28).Li H; Du M; Mleczko MJ; Koh AL; Nishi Y; Pop E; Bard AJ; Zheng X Kinetic Study of Hydrogen Evolution Reaction over Strained MoS2 with Sulfur Vacancies using Scanning Electrochemical Microscopy. J. Am. Chem. Soc 2016, 138, 5123–5129. [DOI] [PubMed] [Google Scholar]
- (29).Henrotte O; Bottein T; Casademont H; Jaouen K; Bourgeteau T; Campidelli S; Derycke V; Jousselme B; Cornut R Electronic Transport of MoS2 Monolayered Flakes Investigated by Scanning Electrochemical Microscopy. ChemPhysChem 2017, 18, 2777–2781. [DOI] [PubMed] [Google Scholar]
- (30).Ahn HS and Bard AJ. Electrochemical Surface Interrogation of a MoS-2 Hydrogen-Evolving Catalyst: In Situ Determination of the Surface Hydride Coverage and the Hydrogen Evolution Kinetics. J. Phys. Chem. Lett 2016, 7, 2748–2752. [DOI] [PubMed] [Google Scholar]
- (31).Kumar S; Sahoo PK; Satpati AK Electrochemical and SECM Investigation of MoS2/GO and MoS2/rGO Nanocomposite Materials for HER Electrocatalysis. ACS Omega 2017, 2, 7532–7545. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (32).Skoog DA; West DM; Holler FJ; Crouch SR Fundamentals of Analytical Chemistry, 8th ed.; Brooks/Cole-Thomson Learning: Belmont, CA, 2004. [Google Scholar]
- (33).Bertocci U; Wagman DD Copper, Silver, and Gold. In Standard Potentials in Aqueous Solution, Bard AJ, Parsons R, Jordan J, Eds.; Marcel Dekker: New York, 1985; pp 287–320. [Google Scholar]
- (34).Velmurugan J; Agrawal A; An S; Choudhary E; Szalai VA Fabrication of Scanning Electrochemical Microscopy-Atomic Force Microscopy Probes to Image Surface Topography and Reactivity at the Nanoscale. Anal. Chem 2017, 89 2687–2691. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (35).Mezour MA; Morin M; Mauzeroll J Fabrication and Characterization of Laser Pulled Platinum Microelectrodes with Controlled Geometry. Anal. Chem 2011, 83, 2378–2382. [DOI] [PubMed] [Google Scholar]
- (36).Rasband WS Image J. http://imagej.nih.gov/ij/ (accessed September 7, 2017).
- (37).Nečas D and Klapetek P. Gwyddion. http://gwyddion.net (accessed November 1, 2017).
- (38).Lee C; Yan H; Brus LE; Heinz TF; Hone J; Ryu S Anomalous Lattice Vibrations of Single- and Few-Layer MoS2. ACS Nano 2010, 4, 2695–2700. [DOI] [PubMed] [Google Scholar]
- (39).Bard AJ and Faulkner LR. Electrochemical Methods: Fundamentals and Applications, 2nd ed.; Wiley: New York, 2001. [Google Scholar]
- (40).Evans BL and Young DA. Delocalized Excitons in Thin Anisotropic Crystals. Phys. Stat. Sol 1968, 25, 417–425. [Google Scholar]
- (41).Phillips ML; and Spitler MT. Role of Electron-Hole Recombination at MoS2 and MoSe2 Photoelectrodes. J. Electrochem. Soc 1981, 10, 2137–2141. [Google Scholar]
- (42).Molina A; Labora E; Rogers EI; Martínez-Ortiz F; Serna C; Limon-Peterson JG; Rees NV; Compton RG Theoretical and Experimental Study of Differential Pulse Voltammetry at Spherical Electrodes: Measuring Diffusion Coefficients and Formal Potentials. J. Electroanal. Chem 2009, 634, 73–81. [Google Scholar]
- (43).Licht S; Cammarata V; Wrighton MR Direct Measurements of the Physical Diffusion of Redox Active Species: Microelectrochemical Experiments and Their Simulation. J. Phys. Chem 1990, 94, 6133–3140. [Google Scholar]
- (44).Lefrou C and Cornut R. Analytical Expressions for Quantitative Scanning Electrochemical Microscopy (SECM). Chem. Phys. Chem 2010, 11, 547–556. [DOI] [PubMed] [Google Scholar]
- (45).Martin RD; Unwin PR Theory and Experiment for the Substrate Generation/Tip Collection Mode of the Scanning Electrochemical Microscope: Application as an Approach for Measuring the Diffusion Coefficient Ratio of a Redox Couple. Anal. Chem 1998, 70, 276–284. [Google Scholar]
- (46).Daum PH and Enke CG. Electrochemical Kinetics of the Ferri-Ferrocyanide Couple on Platinum. Anal. Chem 1969, 41, 653–656. [Google Scholar]
- (47).Macpherson JV; Jones CE; Unwin PR Radial Flow Microring Electrode: Investigation of Fast Heterogeneous Electron-Transfer Processes. J. Phys. Chem. B 1998, 102, 9891–9897. [Google Scholar]
- (48).Gerischer H Impact of Semiconductors on the Concepts of Electrochemistry. Electrochim. Acta 1990, 35, 1677–1699. [Google Scholar]
- (49).Ahmed SM Surface and Photoelectrochemical Studies of Semiconducting MoS2. Electrochim. Acta 1982, 27, 707–712. [Google Scholar]
- (50).Patten HV; Meadow KE; Hutton LA; Iacobini JG; Battistel D; McKelvey K; Colburn AW; Newton ME; Macpherson JV; Unwin PR Electrochemical Mapping Reveals Direct Correlation between Heterogeneous Electron-Transfer Kinetics and Local Density of State in Diamond Electrodes. Angew. Chem. Int. Ed 2012, 51, 7002–7006. [DOI] [PubMed] [Google Scholar]
- (51).Oh HM; Han GH; Kim H; Bae JJ; Jeong MS; Lee YH Photochemical Reaction in Monolayer MoS2 via Correlated Photoluminescence, Raman Spectroscopy, and Atomic Force Microscopy. ACS Nano 2016, 10, 5230–5236. [DOI] [PubMed] [Google Scholar]
- (52).Bard AJ; Denault G; Lee C; Mandler D; Wipf DO Scanning Electrochemical Microscopy-a New Technique for the Characterization and Modification of Surfaces. Acc. Chem. Res 1990, 23, 357–363. [Google Scholar]
- (53).Patten HV; Velicky M; Clark N; Muryn CA; Kinloch IA; Dryfe RAW Electrochemistry of Well-defined Graphene Samples: Role of Contaminants. Faraday Disc. 2014, 172, 261–272. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





