US20190003064A1 - Crumpled Transition Metal Dichalcogenide Sheets - Google Patents

Crumpled Transition Metal Dichalcogenide Sheets Download PDF

Info

Publication number
US20190003064A1
US20190003064A1 US16/010,650 US201816010650A US2019003064A1 US 20190003064 A1 US20190003064 A1 US 20190003064A1 US 201816010650 A US201816010650 A US 201816010650A US 2019003064 A1 US2019003064 A1 US 2019003064A1
Authority
US
United States
Prior art keywords
mos
transition metal
sheets
metal dichalcogenide
fission
Prior art date
Legal status (The legal status is an assumption and is not a legal conclusion. Google has not performed a legal analysis and makes no representation as to the accuracy of the status listed.)
Abandoned
Application number
US16/010,650
Inventor
Stanley Shihyao Chou
Vincent Tung
Yen-Chang Chen
Current Assignee (The listed assignees may be inaccurate. Google has not performed a legal analysis and makes no representation or warranty as to the accuracy of the list.)
National Technology and Engineering Solutions of Sandia LLC
Original Assignee
National Technology and Engineering Solutions of Sandia LLC
Priority date (The priority date is an assumption and is not a legal conclusion. Google has not performed a legal analysis and makes no representation as to the accuracy of the date listed.)
Filing date
Publication date
Application filed by National Technology and Engineering Solutions of Sandia LLC filed Critical National Technology and Engineering Solutions of Sandia LLC
Priority to US16/010,650 priority Critical patent/US20190003064A1/en
Assigned to U.S. DEPARTMENT OF ENERGY reassignment U.S. DEPARTMENT OF ENERGY CONFIRMATORY LICENSE (SEE DOCUMENT FOR DETAILS). Assignors: NATIONAL TECHNOLOGY & ENGINEERING SOLUTIONS OF SANDIA, LLC
Assigned to NATIONAL TECHNOLOGY & ENGINEERING SOLUTIONS OF SANDIA, LLC reassignment NATIONAL TECHNOLOGY & ENGINEERING SOLUTIONS OF SANDIA, LLC ASSIGNMENT OF ASSIGNORS INTEREST (SEE DOCUMENT FOR DETAILS). Assignors: CHOU, Stanley Shihyao
Publication of US20190003064A1 publication Critical patent/US20190003064A1/en
Abandoned legal-status Critical Current

Links

Images

Classifications

    • CCHEMISTRY; METALLURGY
    • C25ELECTROLYTIC OR ELECTROPHORETIC PROCESSES; APPARATUS THEREFOR
    • C25BELECTROLYTIC OR ELECTROPHORETIC PROCESSES FOR THE PRODUCTION OF COMPOUNDS OR NON-METALS; APPARATUS THEREFOR
    • C25B11/00Electrodes; Manufacture thereof not otherwise provided for
    • C25B11/04Electrodes; Manufacture thereof not otherwise provided for characterised by the material
    • CCHEMISTRY; METALLURGY
    • C25ELECTROLYTIC OR ELECTROPHORETIC PROCESSES; APPARATUS THEREFOR
    • C25BELECTROLYTIC OR ELECTROPHORETIC PROCESSES FOR THE PRODUCTION OF COMPOUNDS OR NON-METALS; APPARATUS THEREFOR
    • C25B11/00Electrodes; Manufacture thereof not otherwise provided for
    • C25B11/04Electrodes; Manufacture thereof not otherwise provided for characterised by the material
    • C25B11/051Electrodes formed of electrocatalysts on a substrate or carrier
    • C25B11/073Electrodes formed of electrocatalysts on a substrate or carrier characterised by the electrocatalyst material
    • C25B11/075Electrodes formed of electrocatalysts on a substrate or carrier characterised by the electrocatalyst material consisting of a single catalytic element or catalytic compound
    • C25B11/0447
    • CCHEMISTRY; METALLURGY
    • C25ELECTROLYTIC OR ELECTROPHORETIC PROCESSES; APPARATUS THEREFOR
    • C25BELECTROLYTIC OR ELECTROPHORETIC PROCESSES FOR THE PRODUCTION OF COMPOUNDS OR NON-METALS; APPARATUS THEREFOR
    • C25B1/00Electrolytic production of inorganic compounds or non-metals
    • C25B1/01Products
    • C25B1/02Hydrogen or oxygen
    • C25B1/04Hydrogen or oxygen by electrolysis of water
    • C25B11/0405
    • C25B11/0415
    • CCHEMISTRY; METALLURGY
    • C25ELECTROLYTIC OR ELECTROPHORETIC PROCESSES; APPARATUS THEREFOR
    • C25BELECTROLYTIC OR ELECTROPHORETIC PROCESSES FOR THE PRODUCTION OF COMPOUNDS OR NON-METALS; APPARATUS THEREFOR
    • C25B11/00Electrodes; Manufacture thereof not otherwise provided for
    • C25B11/04Electrodes; Manufacture thereof not otherwise provided for characterised by the material
    • C25B11/051Electrodes formed of electrocatalysts on a substrate or carrier
    • CCHEMISTRY; METALLURGY
    • C25ELECTROLYTIC OR ELECTROPHORETIC PROCESSES; APPARATUS THEREFOR
    • C25BELECTROLYTIC OR ELECTROPHORETIC PROCESSES FOR THE PRODUCTION OF COMPOUNDS OR NON-METALS; APPARATUS THEREFOR
    • C25B11/00Electrodes; Manufacture thereof not otherwise provided for
    • C25B11/04Electrodes; Manufacture thereof not otherwise provided for characterised by the material
    • C25B11/051Electrodes formed of electrocatalysts on a substrate or carrier
    • C25B11/055Electrodes formed of electrocatalysts on a substrate or carrier characterised by the substrate or carrier material
    • C25B11/057Electrodes formed of electrocatalysts on a substrate or carrier characterised by the substrate or carrier material consisting of a single element or compound
    • YGENERAL TAGGING OF NEW TECHNOLOGICAL DEVELOPMENTS; GENERAL TAGGING OF CROSS-SECTIONAL TECHNOLOGIES SPANNING OVER SEVERAL SECTIONS OF THE IPC; TECHNICAL SUBJECTS COVERED BY FORMER USPC CROSS-REFERENCE ART COLLECTIONS [XRACs] AND DIGESTS
    • Y02TECHNOLOGIES OR APPLICATIONS FOR MITIGATION OR ADAPTATION AGAINST CLIMATE CHANGE
    • Y02EREDUCTION OF GREENHOUSE GAS [GHG] EMISSIONS, RELATED TO ENERGY GENERATION, TRANSMISSION OR DISTRIBUTION
    • Y02E60/00Enabling technologies; Technologies with a potential or indirect contribution to GHG emissions mitigation
    • Y02E60/30Hydrogen technology
    • Y02E60/36Hydrogen production from non-carbon containing sources, e.g. by water electrolysis

Definitions

  • the present invention relates to transition metal dichalcogenides and, in particular, to crumpled transition metal dichalcogenide sheets, such as crumpled molybdenum disulfide (MoS 2 ), that can have catalytic and other applications.
  • transition metal dichalcogenides and, in particular, to crumpled transition metal dichalcogenide sheets, such as crumpled molybdenum disulfide (MoS 2 ), that can have catalytic and other applications.
  • Molybdenum disulfide MoS 2
  • MoS 2 Molybdenum disulfide
  • MoS 2 edges possess a favorable ⁇ G H , as well as metallic 1D conductivity due to the presence of electronic states in close proximity to the Fermi level.
  • synthetic approaches that prioritize MoS 2 edges sites over the inert basal plane sites have attracted broad interest. See Q. Ding et al., Chem. 1, 699 (2016); T. F. Jaramillo et al., Science 317, 100 (2017); J. Kibsgaard et al., Nat. Mater. 11, 963 (2012); M.-R. Gao et al., Nat. Commun. 6, 7493 (2015); Z. Lu et al., Adv. Mater.
  • basal plane activation via lithiation and ultrasonication has been used to induce phase transformation into the conductive 1T′ MoS 2 polymorph (sometimes termed “metallic,” “chemically exfoliated” MoS 2 , or ce-MoS 2 ), which is catalytically active on the basal plane.
  • conductive 1T′ MoS 2 polymorph sometimes termed “metallic,” “chemically exfoliated” MoS 2 , or ce-MoS 2
  • M. A. Py and R. R. Haering Can. J. Phys. 61, 76 (1983); S. S. Chou et al., Nat. Commun. 6, 8311 (2015); and D. Voiry et al., Nat. Mater. 12, 850 (2013).
  • phase engineering strategy provides a new pathway to drastically enhance the overall HER metrics of 1T′ MoS 2 that has been recently extended to WS 2 and MoSe 2 .
  • M. a Lukowski et al. Energy Environ. Sci. 7, 2608 (2014); and Y. Yin et al., Adv. Mater. 29, 1700311 (2017).
  • this strategy comes at the cost of reduced robustness, since the 1T′ phase is metastable and temperature sensitive. See M. Chhowalla et al., Nat. Chem. 5, 263 (2013).
  • the present invention is directed to a method of crumpling transition metal dichalcogenide sheets, comprising forming an electrostatically charged liquid droplet comprising a plurality of planar sheets of a transition metal dichalcogenide; applying a sufficient electric field to the droplet to cause the droplet to rupture into a plurality of fission droplets by an electrohydrodynamic process, wherein each fission droplet comprises at least one planar sheet of the transition metal dichalcogenide; and evaporating the remaining liquid from the plurality of fission droplets, thereby inducing spontaneous self-crumpling of the at least one planar sheet of the transition metal dichalcogenide in each of the fission droplets.
  • the invention provides a facile and high throughput method for obtaining electrochemically stable, thermally robust, and highly efficient HER catalysts via electrohydrodynamic (EHD) induced dimensional transition of 2D planar ce-MoS 2 sheets into 3D structurally deformed MoS 2 nanostructures, namely crumpled MoS 2 (c-MoS 2 ).
  • EHD electrohydrodynamic
  • the material obtained from EHD-assisted crumpling, electrostatically isolates and then transforms individual ce-MoS 2 sheets into 3D crumpled architectures, provides highly accessible and electrochemically active surface-area (310.8+/ ⁇ 0.5 m 2 g ⁇ 1 and catalytic site density of 5.86 ⁇ 10 17 sites cm ⁇ 2 ) that allows infiltration of electrolytes to the fully exposed active sites on the activated basal plane and edges while allowing efficient charge transfer between electrolyte, activated basal plane and edges, as confirmed by a substantially reduced charge-transfer resistance (R ct ).
  • the hierarchically strained conformational features can substantially reduce the series resistance (R s ), closely resembling the orbital effect induced by regions of atomically localized curvature in nanocarbon allotropes and textured MoS 2 .
  • R s series resistance
  • the net result is much enhanced HER activity with a Tafel slope of 64 mV per dec, exchange current density of 10.4 ⁇ A cm ⁇ 2 , high cathodic current density (>100 mA cm ⁇ 2 at less than ⁇ 300 mV), and an overpotential ( ⁇ 10 ) of ⁇ 191 mV to achieve a current density of ⁇ 10 mA cm ⁇ 2 at an exceedingly low aerial mass loading of only 13 ⁇ g cm ⁇ 2 .
  • the significantly enhanced intrinsic and total electrode HER characteristics compare favorably with state-of-art MoS 2 catalysts while using a fully scalable and easy to integrate material process. See J. Kibsgaard et al., Nat. Mater. 11, 963 (2012); M.-R.
  • the structurally deformed MoS 2 with in situ strain fields propagating on the thermodynamically favored 2H phase, demonstrates consistent HER characteristics under continuous operations of 5000 cycles or annealing at high temperatures (up to 300° C.), enabling effective integration into fuel cell and other applications.
  • FIG. 1( a ) is a schematic illustration the chemical exfoliation of bulk MoS 2 powder.
  • FIG. 1( b ) is a transmission electron microscope (TEM) image of the exfoliated MoS 2 , showing internal edges which increase overall edge density.
  • FIG. 1( c ) is a high-angle annular dark-field scanning transmission microscope (HAADF-STEM) image, showing spatially distributed atomistic point defects, including single and double sulfur vacancies (V S1 , V S2 ), Mo anti-site substitutions (Mo S2 ) and S anti-site substitutions (S2 Mo ).
  • FIG. 2 is a schematic illustration of EHD-assisted crumpling of exfoliated MoS 2 , showing that chemical exfoliation also facilitates stable colloidal dispersions of ce-MoS 2 within the EHD-generated droplets that can be viewed as charged, noncolloidal reactors.
  • the reactant, 2D ce-MoS 2 within these nanoreactors undergoes stages of (i) electrostatically induced separation, (ii) fission, and (iii) capillarity-induced self-crumpling into 3D c-MoS 2 .
  • FIG. 3( a ) is a transmission electron microscope (TEM) image of the 3D c-MoS 2 .
  • FIG. 3( b ) is a high-resolution transmission electron microscope (HRTEM) image showing extremely thin layers because of electrostatic forces and fission.
  • FIG. 3( c ) is an HRTEM image showing sharp vertices and facets with radii of curvatures at nanoscale.
  • FIG. 3( d ) is an HRTEM image showing propagating ridges.
  • FIG. 4 is a graph of Raman spectra showing a progressive red shift of the E 1 2g with increased deposition temperature, indicating a progressively strained morphology.
  • FIG. 5( a ) shows SEM and HRTEM images of MoS 2 deposited at 100° C., revealing rugged island like morphology and relatively flat, featureless basal plane with slightly corrugated folds around the edges when isolated.
  • FIG. 5( b ) shows SEM and HRTEM images of EHD deposition at 200° C. coincides with emergence of wrinkled sheets.
  • FIG. 5( c ) shows SEM and HRTEM images of MoS 2 deposited at 250° C., showing fully realized individual crumples with porous structures, and highly rippled edges were observed.
  • FIG. 6 is an XPS spectra of the Mo 3d (left) and S 2p (right) peaks showing mixed 1T′ and 2H XPs signatures in samples deposited at 100° C. and pure 2H phase at 200° C. and 250° C.
  • FIGS. 7( a ) and 7( b ) are SEM images showing the uniform and full coverage of c-MoS 2 supported on CC using EHD deposition at different scales.
  • FIG. 7( c ) is an HRSEM image providing a close-up view of densely packed crumpled MoS 2 assemblies with porous morphology and highly curved edges.
  • FIG. 8 is a graph of linear sweep voltammogram (LSV) curves for the carbon cloth (CC) scaffolds, a planar 2H MoS 2 sheet, crumpled MoS 2 deposited at 200° C., and crumpled MoS 2 prepared at 250° C. Overpotentials required to reach current densities of ⁇ 10 mA/cm 2 are labeled.
  • LSV linear sweep voltammogram
  • FIG. 9 is a graph of Tafel plots corresponding to the LSV curves. The plots are labeled with Tafel slopes and exchange current densities, showing much improved total electrode HER characteristics at an exceedingly low mass loading of crumpled MoS 2 .
  • FIG. 1( a ) is a schematic illustration of the synthesis of single layer ce-MoS 2 by a modified Li-intercalation followed by ultrasonic exfoliation. See M. B. Dines, Mater. Res. Bull. 10, 287 (1975); and G. Eda et al., ACS Nano 6, 7311 (2012). In the example described below, a prolonged reaction time of 96 h resulted in the conversion of 64% of trigonal prismatic coordination (2H phase) into octahedral coordination (1T′ phase). High resolution transmission electron microscopy (HRTEM) revealed that the basal plane of the resultant ce-MoS 2 can be readily addressed with significant tears in the 10-nm scale that affords edge-like structures, as shown in FIG.
  • HRTEM transmission electron microscopy
  • HAADF high-angle annular dark field
  • STEM scanning transmission electron microscopy
  • the EHD deposition is characterized by a droplet fission process, in which surface charge density drastically increases during solvent-evaporation-induced volumetric shrinkage, ultimately reaching the Rayleigh limit. At this point, electrostatically charged droplets rupture into smaller ones with a diameter distribution in the sub-micrometer range. Thus, each fine droplet contains single- to few-layered ce-MoS 2 sheets before proceeding to a final stage of self-crumpling.
  • these individualized, charged colloidal droplets can be considered soft templates that impose capillarity-induced geometric constraints to induce the crumpling of ce-MoS 2 sheets upon evaporation. Therefore, ce-MoS 2 within these charged droplets will first separate into approximately individual sheets before being compressed into crumpled nanostructures.
  • Transition metal dichalcogenide sheets are atomically thin semiconductors of the type MX 2 , where M is a transition metal atom (e.g., Mo, W) and X is a chalcogenide atom (e.g., S, Se, Te).
  • M is a transition metal atom (e.g., Mo, W)
  • X is a chalcogenide atom (e.g., S, Se, Te).
  • the sheets comprise a layer of M atoms sandwiched between two layers of X atoms, typically in a hexagonal structure.
  • Bulk crystals of transition metal dichalcogenides comprise monolayer sheets that are bound to each other via weak Van-der-Waals attraction.
  • MX 2 monolayers can be obtained by exfoliation of the bulk crystal, either by mechanical abrasion or by liquid-phase exfoliation.
  • monolayer MX 2 can be grown by chemical vapor deposition or molecular beam epitaxy. These MX 2 sheets can be crumpled following the same EHD process described above.
  • Li-intercalation was accomplished by immersing 2 g of MoS 2 powder in 15 mL of 0.8 m n-butyl lithium in hexane. The mixture was stirred vigorously in an Ar-filled glovebox for 96 h. Next, the intercalated compound was transferred to DI-H 2 O and sonicated to yield exfoliated monolayers. Upon contact with the DI-H 2 O, copious gas evolved at the interlayer interfaces, thus yielding high concentrations of monolayers and forming a black, opaque solution. After ultrasonication, the compound was washed over Whatman filter paper with 500 mL of hexane. 200 mL of DI-H 2 O was then added to the mixture.
  • the mixture was then sonicated for 90 min.
  • the resulting solution was then aliquoted into 45 mL centrifuge tubes with solution depths of 2-3 cm and centrifuged at 5200 rcf for 12 h, decanted, and then resuspended in 100 mL of DI-H 2 O. After resuspension, the centrifugation procedure was repeated twice, followed by resuspension in 50 mL and, later, 25 mL of water.
  • c-MoS 2 with the highest strain ( ⁇ 3.7%), a hot plate was pre-annealed at 250° C. for 60 min prior to deposition.
  • aqueous solutions of ce-MoS 2 250 ⁇ g mL ⁇ 1 ) were fed to a spinneret (gauge 18 TW needle) by a programmable syringe pump.
  • An external electric field of 1.25 kV cm ⁇ 1 was generated with a high-power supply.
  • Computerized multi-pass deposition onto pre-cleaned carbon cloth (CC) HER electrodes was achieved through the integration of an x-y translational stage at a linear stage speed of 2.5 mm s ⁇ 1 .
  • the flow rate and surface temperature were carefully maintained at 25 ⁇ L min ⁇ 1 and 250° C., respectively, to achieve the high yield of c-MoS 2 .
  • the amount of deposited c-MoS 2 catalyst was ⁇ 13 ⁇ g cm ⁇ 2 . Deposition yield was found to linearly scale with the concentration of ce-MoS 2 dispersions, flow rate, and duration of EHD process.
  • edges of c-MoS 2 consist of only few layers, as determined by the clear steps in the relatively flat region, whereas propagating ridges along with sharp vertices, and facets can also be clearly identified, thus creating localized conformational strain with radii of curvature of nanometer scale.
  • Raman spectroscopy substantiates the c-MoS 2 structures by tracing the evolution of two distinct peaks at 385 and 407 cm ⁇ 1 , assigned to the E 1 2g and A 1 g phonon modes of the 2H planar ce-MoS 2 , respectively, as a function of processing temperatures, shown in FIG. 4 .
  • FIG. 4 See S. S. Chou et al., J. Am. Chem. Soc. 137, 1742 (2015).
  • a much more pronounced out-of-plane Mo—S A 1 g phonon was observed, along with a significant red shift of the in-plane Mo—S E 1 2g phonon mode.
  • the enhancement of the A 1 g mode is characteristic of exposed edges that preferentially align with the polarization of the exciting laser. See D. Kong et al., Nano Lett. 13, 1341 (2013). Additionally, the relatively small E 1 2g to A 1 g ratio ( ⁇ 0.5) is an indication of both small domain size and textured morphology. This is not surprising since the crumpling process concurrently induces volumetric shrinkage of ce-MoS 2 and surface texturing of the resultant film. The E 1 2g phonon mode, however, did not shift toward a lower wavenumber relative to the planar 2H ce-MoS 2 even though the overall morphology appears to be heavily wrinkled at a deposition temperature of 100° C.
  • c-MoS 2 films maintained mesoporosity, as indicated by Brunauer-Emmett-Teller (BET) isotherms (surface area of 310.8+/ ⁇ 0.5 m 2 g ⁇ 1 ). This is notable because the combination of electrostatic stabilization and subsequent fission effectively suppresses the formation of large clumps, enabling 3D integration with HER accessible surfaces and micropores. Moreover, in a similar fashion to a paper ball, the deformed ridges made of several overlapping folds form structural supports that prevent the c-MoS 2 from unfolding or collapsing when subjected to external stimuli.
  • BET Brunauer-Emmett-Teller
  • c-MoS 2 can sustain common material processing conditions, such as solution processing, high temperature annealing, or the combination of both, and still maintain a porous, nanotextured morphology necessary for preserving efficient transport of H + and H 2 to the edge sites and activated basal plane under continuous electrochemical operation or elevated temperate.
  • these structural features revealed by the Raman, BET, SEM, and TEM characterizations described above, indicate that c-MoS 2 can be used as a catalytically active, electrochemically and thermally stable component for HER.
  • EHD process offers the distinct advantage to directly integrate c-MoS 2 into diverse device architectures, as well as onto a variety of substrates, including thin films of carbon nanotubes (CNTs), silicon, and even nanostructured polystyrene, in both monolithic and a selective fashion.
  • CNTs carbon nanotubes
  • Si silicon
  • nanostructured polystyrene in both monolithic and a selective fashion.
  • electrodes were formed by depositing thin films of c-MoS 2 onto CC via EHD deposition, as described above.
  • the substrate had an area of 1 cm 2 and the total aerial mass loading was ⁇ 13 ⁇ g cm ⁇ 2 .
  • FIG. 7( c ) further suggests that mesoporous c-MoS 2 structures remains intact and retains a high degree of crumpling even upon continuous deposition.
  • EDX energy-dispersive X-ray mapping of relevant elements (not shown), including Mo and S, closely overlap the contours and areas of c-MoS 2 and underlying CC substrates.
  • the HER cathodes made of c-MoS 2 supported on CC were examined using a standard three-electrode electrochemical configuration in 0.5 m H 2 SO 4 electrolyte de-aerated with Ar. Following deposition and prior to measurement, all samples were annealed at 250° C. for 1 h to ensure complete conversion of remnant 1T′ phases to 2H phases. This provides a well-defined system to unambiguously explore the correlation between the impact of 3D nanostructures, strain-induced modulation, and the catalytic activity in phase pure c-MoS 2 .
  • LSVs linear sweep voltammograms
  • CC electrodes did not show any pronounced cathodic current, and remained inactive during LSV scans.
  • the planar 2H ce-MoS 2 standard electrodes showed very poor HER characteristics even with a high density of S vacancies that have shown to activate the basal plane, delivering an exceedingly high ⁇ 10 of ⁇ 364 mV.
  • This poor HER performance can be attributed to the lack of electronic coupling between the substrates and active sites in combination with a heavily aggregated morphology that limits the exposure of activated basal plane toward HER. See H. Li et al., Nat. Mater. 15, 48 (2016). Meanwhile, increasing the deposition temperature to 100° C. led to the formation of a rippled morphology and ⁇ 10 was reduced to ⁇ 340 mV.
  • ⁇ 10 of the c-MoS 2 synthesized at 200° C. was significantly reduced to ⁇ 309 mV due to the presence of wrinkled morphology and moderate strain load ( ⁇ 1.2%).
  • c-MoS 2 synthesized at 250° C. which was characterized by a combination of high strain load ( ⁇ 3.7%) and 3D, fully crumpled nanostructures, displayed substantially improved total electrode metrics, including ⁇ 10 of ⁇ 191 mV, exchange current densities of 10.4 ⁇ A cm ⁇ 2 , and high cathodic current densities.
  • Tafel slopes were found to decrease with increasing magnitude of crumpling, and were correlated with the strain load.
  • the lowest Tafel slope of 64 mV per dec was obtained for the c-MoS 2 synthesized at 250° C., suggesting a Heyrovsky mechanism at the surface of c-MoS 2 with the hydrogen desorption reaction as the limiting step: H ads +H 3 O + +e ⁇ ⁇ H 2 +H 2 O.
  • the active site numbers were further estimated using the underpotential deposition (UPD) method. See A. Y. Lu et al., Small 12, 5530 (2016); and C. L. Green and A. Kucernak, J. Phys. Chem. 2, 11446 (2002).
  • UPD underpotential deposition
  • the density of active sites was determined to be approximately 1 ⁇ 10 14 sites cm ⁇ 2 for planar 2H ce-MoS 2 , increased to 5.7 ⁇ 10 16 sites cm ⁇ 2 for planar 1T′ ce-MoS 2 polymorph, and further increased to 5.86 ⁇ 10 17 sites cm ⁇ 2 for 2H c-MoS 2 , thus corroborating that the advent of the hierarchically structured, 3D morphology of c-MoS 2 from the intrinsically planar morphology of ce-MoS 2 and enabling infiltration of electrolyte for exposure to both activated basal plane and edges. See J. Kibsgaard et al., Nat. Mater. 11, 963 (2012); D. Kong et al., Nano Lett.
  • EIS electrochemical impedance spectroscopy
  • the measured R s was significantly reduced from 38 ⁇ (planar 2H ce-MoS 2 ) to 1.5 ⁇ (2H c-MoS 2 ). This is unexpected, since the 2H c-MoS 2 does not comprise locally transformed phases and atomically sharp interfaces to lower the contact resistance. See D. Voiry et al., Nat. Mater. 15, 1003 (2016); and D.
  • the TOF was calculated to infer the intrinsic catalytic activity per active sites.
  • TOF was determined to be ⁇ 0.130 s ⁇ 1 at ⁇ 200 mV vs. RHE for 2H c-MoS 2 synthesized at 250° C., an order of magnitude higher than the 2H planar counterpart (0.016 s ⁇ 1 ).
  • the high strain load modulates the ⁇ G H and tailors the TOF at each active site.
  • the TOF achieved is comparable to or better than values measured for strain-vacancy engineered MoS 2 , edge sites of MoS 2 , vertical MoS 2 flakes, 1T′ MoS 2 nanoparticles, and [Mo 3 S 4 ] 4+ cluster of amorphous MoS 2 .
  • c-MoS 2 has advantages to synergistically tailor both intrinsic and total electrode activities in tandem with readily scalable synthesis, assembly, and processing routes. See J. Kibsgaard et al., Nat. Chem. 6, 248 (2014). Meanwhile, the loading of c-MoS 2 was increased in an attempt to further improve the HER metrics through increasing numbers of active sites. However, HER metrics drastically decreased such that the thickest film in fact exhibited the poorest performance.
  • the major limiting factors are the high ⁇ 10 and low current densities.
  • the high resistance may arise from the discontinuous pathways where outer layers of c-MoS 2 are loosely attached to each other, as suggested by EDX mapping.
  • the limited contacts at the top of the films serve as constriction points, thus diminishing electron accessibility and adversely affecting the overall HER performance.
  • This limitation can be potentially mediated by adding conductive adducts, such as graphene, CNTs, or even 1T′ ce-MoS 2 , to simultaneously minimize charge transport distance and series resistance, while maintaining the accessible surface area.
  • the electrochemical stability of the c-MoS 2 catalyst was also evaluated by prolonged electrolysis at constant potentials.
  • the current density of c-MoS 2 remained stable in 0.5 m H 2 SO 4 for more than 100 h and showed a negligible change after 110 h.
  • the uniformity and crumpled nanostructure of c-MoS 2 was also examined under SEM and TEM after continuous linear potential sweeps, and appeared to be intact after the degradation measurements. In addition, when subjecting to thermal annealing at 300° C.
  • the invention enables the formation of 3D structurally deformed c-MoS 2 catalysts for electrochemically stable, thermally resistant, and highly efficient HER.
  • the use of electrostatically charged droplets as nanoreactors facilitates colloidal dispersion, fission, and capillarity-induced-self-crumpling of 2D ce-MoS 2 , ultimately transforming it into 3D c-MoS 2 .
  • electrochemical characterization indicates that HER electrodes composed of c-MoS 2 catalysts synergistically integrate advantageous features of morphologically, chemically, electronically, and mechanically engineered MoS 2 catalysts. See J. Kibsgaard et al., Nat. Mater. 11, 963 (2012); D.
  • R ct 13 2H c-MoS 2 (200° C.) ⁇ 309 92 5.2
  • R s 17 2.2 48.6 13 2.0 ⁇ 10 15 0.080 2000 300° C.

Landscapes

  • Chemical & Material Sciences (AREA)
  • Engineering & Computer Science (AREA)
  • Chemical Kinetics & Catalysis (AREA)
  • Electrochemistry (AREA)
  • Materials Engineering (AREA)
  • Metallurgy (AREA)
  • Organic Chemistry (AREA)
  • Inorganic Chemistry (AREA)
  • Catalysts (AREA)

Abstract

An electrohydrodynamic (EHD) method results in charge driven droplet fission and volumetric shrinkage of metal dichalcogenide sheet droplets, thus wrinkling individual exfoliated metal dichalcogenide sheets on the nanoscale. For example, the method can be used to activate the basal plane of solution exfoliated, stable 2H reverted sheets of MoS2. In this manner, the basal plane of 2H MoS2 can be strained at an unprecedented scale (3.7%), resulting in charge transport to basal plane defects. The resulting crumpled MoS2 integrates few layer sheets and atomistic defects with nano- and micro-scale ridges and vertices to enable centimeter-length films that elevate the catalytic site count of MoS2 from 1.0×1014 sites/cm2 to 2.93×1017 sites/cm2 (planar 2H vs. crumpled 2H) and a turn-over-frequency (TOF) from 0.016 s−1 to 0.130 s−1.

Description

    CROSS-REFERENCE TO RELATED APPLICATION
  • This application claims the benefit of U.S. Provisional Application No. 62/526,828, filed Jun. 29, 2017, which is incorporated herein by reference.
  • STATEMENT OF GOVERNMENT INTEREST
  • This invention was made with Government support under Contract No. DE-NA0003525 awarded by the United States Department of Energy/National Nuclear Security Administration. The Government has certain rights in the invention.
  • FIELD OF THE INVENTION
  • The present invention relates to transition metal dichalcogenides and, in particular, to crumpled transition metal dichalcogenide sheets, such as crumpled molybdenum disulfide (MoS2), that can have catalytic and other applications.
  • BACKGROUND OF THE INVENTION
  • The urgent need to transition the global energy infrastructure away from fossil fuels and toward carbon-neutral sources has propelled a resurgence of research in renewable hydrogen generation. Within this context, there is great interest in the development of hydrogen evolution reaction (HER) electrocatalysts that concurrently possess high total electrode activity (e.g., overpotentials, Tafel slope, and exchange current density) at a low catalyst loading, and tailored intrinsic activity (e.g., turnover frequency (TOF) per active site). See J. D. Benck et al., ACS Catal. 4, 3957 (2014). To have widespread impact, such catalytic materials must also be electrochemically stable, thermally robust, and synthetically scalable without the need for specialized processing equipment. While Pt nanoparticles represent the current benchmark for HER electrocatalysis, the elemental scarcity, high cost, and poor stability against nanoparticle coarsening and agglomeration impose practical limitations and motivate the need for a next generation of alternative HER catalysts.
  • Molybdenum disulfide (MoS2), a van der Waals solid composed of adhered-2D monolayers, has emerged as a promising alternative to Pt due to its elemental abundance, high catalytic activity, and electrochemical stability. See J. D. Benck et al., ACS Catal. 4, 3957 (2014); J. Greeley et al., Nat. Mater. 5, 909 (2006); B. Hinnemann et al., J. Am. Chem. Soc. 127, 5308 (2005); and Q. Ding et al., Chem. 1, 699 (2016). Despite the various merits of MoS2, its functional performance remains inferior to that of Pt due to comparatively small active site concentrations as well as its relatively low electrical conductivity. At the root of functional limitations is the inability of MoS2, in its natural 2H semiconducting phase, to catalyze HER on its inert basal plane, where the Gibbs free energy for hydrogen adsorption (ΔGH) is exceedingly large. See J. D. Benck et al., ACS Catal. 4, 3957 (2014); T. F. Jaramillo et al., Science 317, 100 (2017); and J. Kibsgaard et al., Nat. Mater. 11, 963 (2012). In contrast, MoS2 edges possess a favorable ΔGH, as well as metallic 1D conductivity due to the presence of electronic states in close proximity to the Fermi level. Because of the highly catalytic nature of MoS2 edges, synthetic approaches that prioritize MoS2 edges sites over the inert basal plane sites have attracted broad interest. See Q. Ding et al., Chem. 1, 699 (2016); T. F. Jaramillo et al., Science 317, 100 (2017); J. Kibsgaard et al., Nat. Mater. 11, 963 (2012); M.-R. Gao et al., Nat. Commun. 6, 7493 (2015); Z. Lu et al., Adv. Mater. 26, 2683 (2014); and J. Ekspong et al., Adv. Funct. Mater. 26, 6766 (2016). Notable innovations in this area include nanostructured MoS2 particles, vertically aligned flakes, and wires in conjunction with the engineering of defects, interfacing to conductive scaffolds, and mimicry of the edge structures. See Y. Li et al., J. Am. Chem. Soc. 133, 7296 (2011); D. Kong et al., Nano Lett. 13, 1341 (2013); Z. Chen et al., Nano Lett. 11, 4168 (2011); J. Xie et al., Adv. Mater. 25, 5807 (2013); M. A. Lukowski et al., J. Am. Chem. Soc. 135, 10274 (2013); G. Ye et al., Nano Lett. 16, 1097 (2016); A. J. Smith et al., Adv. Energy Mater. 4, 1400398 (2014); Y. H. Chang et al., Adv. Mater. 25, 756 (2013); and H. I. Karunadasa et al., Science 335, 698 (2012). Yet, despite these innovations, the fundamental inactivity of the MoS2 basal plane remains a basic material constraint. To enhance catalytic activity, basal plane activation via lithiation and ultrasonication has been used to induce phase transformation into the conductive 1T′ MoS2 polymorph (sometimes termed “metallic,” “chemically exfoliated” MoS2, or ce-MoS2), which is catalytically active on the basal plane. See M. A. Py and R. R. Haering, Can. J. Phys. 61, 76 (1983); S. S. Chou et al., Nat. Commun. 6, 8311 (2015); and D. Voiry et al., Nat. Mater. 12, 850 (2013). The phase engineering strategy provides a new pathway to drastically enhance the overall HER metrics of 1T′ MoS2 that has been recently extended to WS2 and MoSe2. See M. a Lukowski et al., Energy Environ. Sci. 7, 2608 (2014); and Y. Yin et al., Adv. Mater. 29, 1700311 (2017). However, this strategy comes at the cost of reduced robustness, since the 1T′ phase is metastable and temperature sensitive. See M. Chhowalla et al., Nat. Chem. 5, 263 (2013).
  • Beyond phase transformation, basal plane activation via strain engineering has been explored. This approach takes advantage of the shift of the d-band electronic structure with strain, which facilitates a semiconducting-to-metallic phase transformation. See L. Hromadova et al., Phys. Rev. B: Condens. Matter Mater. Phys. 87, 1 (2013). However, retention of the metallic phase requires continual application of pressure (≈35 GPa), which is impractical for technological use. See Y.-C. Lin et al., Nat. Nanotechnol. 9, 391 (2014); A. P. Nayak et al., Nat. Commun. 5, 1 (2014); and Z. H. Chi et al., Phys. Rev. Lett. 113, 1 (2014). A hybrid approach using strain on plasma-treated MoS2 has also been reported, with localized substrate strain plus S vacancies from the plasma treatment shifting its ΔGH towards thermal neutral (≈0 eV), thus activating the basal plane and enabling HER. See H. Li et al., Nat. Mater. 15, 48 (2016); H. Li et al., Nat. Commun. 6, 7381 (2015); C. C. Cheng et al., Nano Energy 30, 846 (2016); and G. Li et al., J. Am. Chem. Soc. 138, 16632 (2016). However, as strain here is derived from lithographically defined substrate topography, it is potentially too costly for production. See H. Li et al., Nat. Commun. 6, 7381 (2015). Furthermore, because the product is inherently planar, it has limited prospects for integration in multilayer systems, which is a practical need for many applications and constrains device architectures. Most recently, the basal plane of MoS2 has also been “activated” by improving substrate/electron transport to defects within MoS2 sheets. See D. Voiry et al., Nat. Mater. 15, 1003 (2016). However, this method similarly requires lithography and patterning, rendering scalability a challenge. It should also be noted that the 1T′, the strained S-vacancy, and the electronic coupling approaches to basal plane activation are all limited to extremely thin specimens to 2 layers), thus limiting the number of active sites and making scalable integration difficult.
  • SUMMARY OF THE INVENTION
  • The present invention is directed to a method of crumpling transition metal dichalcogenide sheets, comprising forming an electrostatically charged liquid droplet comprising a plurality of planar sheets of a transition metal dichalcogenide; applying a sufficient electric field to the droplet to cause the droplet to rupture into a plurality of fission droplets by an electrohydrodynamic process, wherein each fission droplet comprises at least one planar sheet of the transition metal dichalcogenide; and evaporating the remaining liquid from the plurality of fission droplets, thereby inducing spontaneous self-crumpling of the at least one planar sheet of the transition metal dichalcogenide in each of the fission droplets.
  • The invention provides a facile and high throughput method for obtaining electrochemically stable, thermally robust, and highly efficient HER catalysts via electrohydrodynamic (EHD) induced dimensional transition of 2D planar ce-MoS2 sheets into 3D structurally deformed MoS2 nanostructures, namely crumpled MoS2 (c-MoS2). The material, obtained from EHD-assisted crumpling, electrostatically isolates and then transforms individual ce-MoS2 sheets into 3D crumpled architectures, provides highly accessible and electrochemically active surface-area (310.8+/−0.5 m2 g−1 and catalytic site density of 5.86×1017 sites cm−2) that allows infiltration of electrolytes to the fully exposed active sites on the activated basal plane and edges while allowing efficient charge transfer between electrolyte, activated basal plane and edges, as confirmed by a substantially reduced charge-transfer resistance (Rct). Meanwhile, the hierarchically strained conformational features, including facets, folds, ridges, vertices, and wrinkles, can substantially reduce the series resistance (Rs), closely resembling the orbital effect induced by regions of atomically localized curvature in nanocarbon allotropes and textured MoS2. See S. Luo et al., Nanotechnology 26, 105705 (2015); and D. Srivastava et al., J. Phys. Chem. B 103, 4330 (1999). In addition, the synergistic combination of high strain load emanating from structurally deformed morphology and spatially distributed S vacancies intrinsic to the chemical exfoliation process collectively tailors the 2H basal plane with a catalytically and thermodynamically favorable environment, as confirmed by the improved TOF of 0.130 H2 molecules per second. See D. Srivastava et al., J. Phys. Chem. B 103, 4330 (1999). The net result is much enhanced HER activity with a Tafel slope of 64 mV per dec, exchange current density of 10.4 μA cm−2, high cathodic current density (>100 mA cm−2 at less than −300 mV), and an overpotential (η10) of −191 mV to achieve a current density of −10 mA cm−2 at an exceedingly low aerial mass loading of only 13 μg cm−2. The significantly enhanced intrinsic and total electrode HER characteristics compare favorably with state-of-art MoS2 catalysts while using a fully scalable and easy to integrate material process. See J. Kibsgaard et al., Nat. Mater. 11, 963 (2012); M.-R. Gao et al., Nat. Commun. 6, 7493 (2015); Y. Li et al., J. Am. Chem. Soc. 133, 7296 (2011); D. Kong et al., Nano Lett. 13, 1341 (2013); M. A. Lukowski et al., J. Am. Chem. Soc. 135, 10274 (2013); G. Ye et al., Nano Lett. 16, 1097 (2016); H. Li et al., Nat. Mater. 15, 48 (2016); and D. Voiry et al., Nano Lett. 13, 6222 (2013). More importantly, the structurally deformed MoS2, with in situ strain fields propagating on the thermodynamically favored 2H phase, demonstrates consistent HER characteristics under continuous operations of 5000 cycles or annealing at high temperatures (up to 300° C.), enabling effective integration into fuel cell and other applications.
  • BRIEF DESCRIPTION OF THE DRAWINGS
  • The detailed description will refer to the following drawings, wherein like elements are referred to by like numbers.
  • FIG. 1(a) is a schematic illustration the chemical exfoliation of bulk MoS2 powder. FIG. 1(b) is a transmission electron microscope (TEM) image of the exfoliated MoS2, showing internal edges which increase overall edge density. FIG. 1(c) is a high-angle annular dark-field scanning transmission microscope (HAADF-STEM) image, showing spatially distributed atomistic point defects, including single and double sulfur vacancies (VS1, VS2), Mo anti-site substitutions (MoS2) and S anti-site substitutions (S2Mo).
  • FIG. 2 is a schematic illustration of EHD-assisted crumpling of exfoliated MoS2, showing that chemical exfoliation also facilitates stable colloidal dispersions of ce-MoS2 within the EHD-generated droplets that can be viewed as charged, noncolloidal reactors. Under a high electric field, the reactant, 2D ce-MoS2, within these nanoreactors undergoes stages of (i) electrostatically induced separation, (ii) fission, and (iii) capillarity-induced self-crumpling into 3D c-MoS2.
  • FIG. 3(a) is a transmission electron microscope (TEM) image of the 3D c-MoS2. FIG. 3(b) is a high-resolution transmission electron microscope (HRTEM) image showing extremely thin layers because of electrostatic forces and fission. FIG. 3(c) is an HRTEM image showing sharp vertices and facets with radii of curvatures at nanoscale. FIG. 3(d) is an HRTEM image showing propagating ridges.
  • FIG. 4 is a graph of Raman spectra showing a progressive red shift of the E1 2g with increased deposition temperature, indicating a progressively strained morphology.
  • FIG. 5(a) shows SEM and HRTEM images of MoS2 deposited at 100° C., revealing rugged island like morphology and relatively flat, featureless basal plane with slightly corrugated folds around the edges when isolated. FIG. 5(b) shows SEM and HRTEM images of EHD deposition at 200° C. coincides with emergence of wrinkled sheets. FIG. 5(c) shows SEM and HRTEM images of MoS2 deposited at 250° C., showing fully realized individual crumples with porous structures, and highly rippled edges were observed.
  • FIG. 6 is an XPS spectra of the Mo 3d (left) and S 2p (right) peaks showing mixed 1T′ and 2H XPs signatures in samples deposited at 100° C. and pure 2H phase at 200° C. and 250° C.
  • FIGS. 7(a) and 7(b) are SEM images showing the uniform and full coverage of c-MoS2 supported on CC using EHD deposition at different scales. FIG. 7(c) is an HRSEM image providing a close-up view of densely packed crumpled MoS2 assemblies with porous morphology and highly curved edges.
  • FIG. 8 is a graph of linear sweep voltammogram (LSV) curves for the carbon cloth (CC) scaffolds, a planar 2H MoS2 sheet, crumpled MoS2 deposited at 200° C., and crumpled MoS2 prepared at 250° C. Overpotentials required to reach current densities of −10 mA/cm2 are labeled.
  • FIG. 9 is a graph of Tafel plots corresponding to the LSV curves. The plots are labeled with Tafel slopes and exchange current densities, showing much improved total electrode HER characteristics at an exceedingly low mass loading of crumpled MoS2.
  • DETAILED DESCRIPTION OF THE INVENTION
  • FIG. 1(a) is a schematic illustration of the synthesis of single layer ce-MoS2 by a modified Li-intercalation followed by ultrasonic exfoliation. See M. B. Dines, Mater. Res. Bull. 10, 287 (1975); and G. Eda et al., ACS Nano 6, 7311 (2012). In the example described below, a prolonged reaction time of 96 h resulted in the conversion of 64% of trigonal prismatic coordination (2H phase) into octahedral coordination (1T′ phase). High resolution transmission electron microscopy (HRTEM) revealed that the basal plane of the resultant ce-MoS2 can be readily addressed with significant tears in the 10-nm scale that affords edge-like structures, as shown in FIG. 1(b). Additionally, as shown in FIG. 1(c), high-angle annular dark field (HAADF) and aberration-corrected scanning transmission electron microscopy (STEM) images reveal atomically resolved point defects including S vacancies (VS), Mo anti-site substitution (Mos), and Mo vacancies (VMo), mono in tandem with bi-sulfur vacancies (VS1, VS2) and antisite substitutions within MoS2. See J. Xie et al., Adv. Mater. 25, 5807 (2013). Detailed analysis of the MoS2 basal plane with lateral dimensions of 6.75 nm×6.75 nm (estimated from the lattice distance of 3.16 Å) revealed that the large majority of the defects consist of VS with a total number about 79 (out of 440 Mo atoms, ≈9% S-vacancies). Theoretical calculations indicate that S vacancies as low as ≈3% lowers ΔGH from more than 2 eV down to nearly 0.2 eV. See H. Li et al., Nat. Mater. 15, 48 (2016). The average S-vacancy percentage over 8 such areas was calculated to be 9.3+/−1.7%, thus creating a catalytically active basal plane. Although continuous modulation of ΔGH toward thermal neutral can be made possible by mounting up the number of S-vacancies, the overall structural integrity of ce-MoS2 sheets is at the mercy of fragmentation.
  • To further modulate the ΔGH of ce-MoS2 sheets while preserving the structural integrity, an EHD process was adapted, which is typically employed for generating ultrafine droplets in high throughput thin film production. See H. Ishihara et al., Sci. Rep. 6, 38701 (2016); H. Ishihara et al., Adv. Mater. Interfaces 3, 1 (2016); and H. Ishihara et al., J. Mater. Chem. A 4, 6989 (2016). As shown in FIG. 2, under a high electric field, the liquid medium deforms and then disintegrates into self-dispersing, charged droplets, providing electrostatically stabilized colloidal systems that inhibit ce-MoS2 sheets from restacking. The EHD deposition is characterized by a droplet fission process, in which surface charge density drastically increases during solvent-evaporation-induced volumetric shrinkage, ultimately reaching the Rayleigh limit. At this point, electrostatically charged droplets rupture into smaller ones with a diameter distribution in the sub-micrometer range. Thus, each fine droplet contains single- to few-layered ce-MoS2 sheets before proceeding to a final stage of self-crumpling. In addition to being a dispersing medium, these individualized, charged colloidal droplets can be considered soft templates that impose capillarity-induced geometric constraints to induce the crumpling of ce-MoS2 sheets upon evaporation. Therefore, ce-MoS2 within these charged droplets will first separate into approximately individual sheets before being compressed into crumpled nanostructures.
  • The method of EHD-assisted crumpling can generally be applied to the crumpling of any transition metal dichalcogenide sheet. Transition metal dichalcogenide sheets are atomically thin semiconductors of the type MX2, where M is a transition metal atom (e.g., Mo, W) and X is a chalcogenide atom (e.g., S, Se, Te). The sheets comprise a layer of M atoms sandwiched between two layers of X atoms, typically in a hexagonal structure. Bulk crystals of transition metal dichalcogenides comprise monolayer sheets that are bound to each other via weak Van-der-Waals attraction. Therefore, MX2 monolayers can be obtained by exfoliation of the bulk crystal, either by mechanical abrasion or by liquid-phase exfoliation. Alternatively, monolayer MX2 can be grown by chemical vapor deposition or molecular beam epitaxy. These MX2 sheets can be crumpled following the same EHD process described above.
  • EXAMPLE Lithium Intercalation and Exfoliation and MoS2 Crumpling by EHD
  • As an example of the invention, Li-intercalation was accomplished by immersing 2 g of MoS2 powder in 15 mL of 0.8 m n-butyl lithium in hexane. The mixture was stirred vigorously in an Ar-filled glovebox for 96 h. Next, the intercalated compound was transferred to DI-H2O and sonicated to yield exfoliated monolayers. Upon contact with the DI-H2O, copious gas evolved at the interlayer interfaces, thus yielding high concentrations of monolayers and forming a black, opaque solution. After ultrasonication, the compound was washed over Whatman filter paper with 500 mL of hexane. 200 mL of DI-H2O was then added to the mixture. With the filter paper removed, the mixture was then sonicated for 90 min. The resulting solution was then aliquoted into 45 mL centrifuge tubes with solution depths of 2-3 cm and centrifuged at 5200 rcf for 12 h, decanted, and then resuspended in 100 mL of DI-H2O. After resuspension, the centrifugation procedure was repeated twice, followed by resuspension in 50 mL and, later, 25 mL of water. The resulting ce-MoS2 was further diluted to 250 μg mL−1 in a mixture of DI water and isopropyl alcohol (DI-H2O: IPA=7:3, v/v) for subsequent EHD crumpling.
  • To prepare c-MoS2 with the highest strain (≈3.7%), a hot plate was pre-annealed at 250° C. for 60 min prior to deposition. Next, aqueous solutions of ce-MoS2 (250 μg mL−1) were fed to a spinneret (gauge 18 TW needle) by a programmable syringe pump. An external electric field of 1.25 kV cm−1 was generated with a high-power supply. Computerized multi-pass deposition onto pre-cleaned carbon cloth (CC) HER electrodes was achieved through the integration of an x-y translational stage at a linear stage speed of 2.5 mm s−1. The flow rate and surface temperature were carefully maintained at 25 μL min−1 and 250° C., respectively, to achieve the high yield of c-MoS2. The amount of deposited c-MoS2 catalyst was ≈13 μg cm−2. Deposition yield was found to linearly scale with the concentration of ce-MoS2 dispersions, flow rate, and duration of EHD process.
  • Scanning electron microscopy (SEM) and atomic force microscopy (AFM) images revealed largely separated and spatially distributed ce-MoS2 sheets with average thickness in the range of nm when the threshold electric field reached 1.25 kV cm−1 for the inception of the fission process. This result confirms that electrostatic stabilization prevents unwanted aggregation. Furthermore, ce-MoS2 was found to spontaneously crumple into 3D porous, structurally deformed nanostructures with many wrinkles, ridges, and corrugated edges upon reaching both the threshold electric field of 1.25 kV cm−1 and a deposition temperature of 250° C., as suggested in the TEM image shown in FIG. 3(a). HAADF and STEM images collectively showed that the resulting c-MoS2 deposited at 250° C. comprises a predominant 2H phase because the deposition temperature exceeded the temperature for transformation of the metastable 1T′ phase to the 2H phase. See G. Eda et al., ACS Nano 6, 7311 (2012); and S. S. Chou et al., J. Am. Chem. Soc. 137, 1742 (2015). The impact of the EHD-assisted crumpling process on the structure of c-MoS2 at the nanoscale was further probed using HRTEM. The HRTEM image in FIG. 3(b) pinpoints areas near edges. FIGS. 3(c) and 3(d) show areas on the basal plane. The edges of c-MoS2 consist of only few layers, as determined by the clear steps in the relatively flat region, whereas propagating ridges along with sharp vertices, and facets can also be clearly identified, thus creating localized conformational strain with radii of curvature of nanometer scale.
  • Raman spectroscopy substantiates the c-MoS2 structures by tracing the evolution of two distinct peaks at 385 and 407 cm−1, assigned to the E1 2g and A1 g phonon modes of the 2H planar ce-MoS2, respectively, as a function of processing temperatures, shown in FIG. 4. See S. S. Chou et al., J. Am. Chem. Soc. 137, 1742 (2015). At higher deposition temperatures, a much more pronounced out-of-plane Mo—S A1 g phonon was observed, along with a significant red shift of the in-plane Mo—S E1 2g phonon mode. In the former, the enhancement of the A1 g mode is characteristic of exposed edges that preferentially align with the polarization of the exciting laser. See D. Kong et al., Nano Lett. 13, 1341 (2013). Additionally, the relatively small E1 2g to A1 g ratio (≈0.5) is an indication of both small domain size and textured morphology. This is not surprising since the crumpling process concurrently induces volumetric shrinkage of ce-MoS2 and surface texturing of the resultant film. The E1 2g phonon mode, however, did not shift toward a lower wavenumber relative to the planar 2H ce-MoS2 even though the overall morphology appears to be heavily wrinkled at a deposition temperature of 100° C. Instead, a pronounced red shift of the E1 2g mode toward lower wavenumbers only took place in the specimens synthesized at temperatures higher than 200° C. The shift of the E1 2g phonon mode provides an index for quantifying surface strain load and has been shown to shift by ≈1.7 cm−1 per % strain on the MoS2. See A. Castellanos-Gomez et al., Nano Lett. 13, 5361 (2013). Here, c-MoS2 synthesized at 100° C. did not yield a significant E1 2g shift, and thus strain, whereas the E1 2g peak of c-MoS2 synthesized at 200° C. was red shifted by 2 cm−1 (≈1.2% of strain). The E1 2g phonon mode further redshifted for c-MoS2 formed at 250° C., generating an average strain load of ≈3.7%. For reference, this strain is comparable to that reported by Zheng and co-workers for modulating S-deficient MoS2 basal planes. See H. Li et al., Nat. Mater. 15, 48 (2016).
  • To better correlate morphology with the observed strain induced E1 2g shift, the morphological evolution of c-MoS2 deposited on carbon cloth (CC) was analyzed as a function of deposition temperature. As described above, ce-MoS2 sheet precursors aggregated into discrete, rugged islands at the low deposition temperature of 100° C. Nevertheless, TEM image reveals that individual ce-MoS2 sheets display a relatively flat and featureless morphology with slightly corrugated folds around the edges, as shown in FIG. 5(a). Comparatively, aggregated clumps begin to disintegrate into largely buckled structures with the emergence of thin, vertically protruding walls, meandering ridges, curled edges, and folds when processed at 200° C., as shown in FIG. 5(b), matching the onset of the red shift in the E1 2g phonon mode and the corresponding moderate strain load. At 250° C., almost all the ce-MoS2 sheets were transformed and compressed into 3D structurally deformed nanostructures with petal-like, curved morphology, as shown in FIG. 5(c), and thus the high magnitude of strain, an important feature necessary for further modulating the ΔGH. See J. D. Benck et al., ACS Catal. 4, 3957 (2014); and H. Li et al., Nat. Mater. 15, 48 (2016). In particular, the hierarchically crumpled MoS2 at 250° C. reveals a highly strained c-MoS2 that comprises corrugated edges and extremely thin layers, narrow folds converged to form propagating ridges, and sharp vertices and facets with radii of curvatures at nanoscale. Rich and spatially distributed defects are throughout.
  • In parallel, the chemical states of Mo and S were investigated by X-ray photoelectron spectroscopy (XPS) on c-MoS2 synthesized at different temperatures, as shown in FIG. 6. Deconvolution of the Mo 3d core level region from c-MoS2 synthesized at 100° C. showed additional peaks at 228.1 eV (Mo 3d5/2) and 231.3 eV (Mo 3d3/2), relative to pure 2H MoS2, which exhibited peaks at 229.1 eV (Mo 3d5/2) and 232.2 eV (Mo 3d3/2). The presence of these additional core level photoemission peaks indicates the presence of the 1T′ phase and integrated peak area ratios reveal a 2H phase fraction of 76% for c-MoS2 synthesized at 100° C. See S. S. Chou et al., J. Am. Chem. Soc. 137, 1742 (2015); and J. Heising and M. G. Kanatzidis, J. Am. Chem. Soc. 121, 11720 (1999). Images obtained through HRTEM, along with false-colored HAADF, confirm the coexistence of both phases at a deposition temperature lower than that of phase transformation (100° C.), which is consistent with the spectroscopic interpretation. In contrast, a phase transition to pure 2H material was observed when c-MoS2 was processed at 200° C. The predominant 2H to 1T′ ratio remained constant at 250° C., demonstrating stable 2H crumpling. As shown in the right-hand panels in FIG. 6, complementary analysis of S 2p core-level spectra corroborate these results.
  • These c-MoS2 films maintained mesoporosity, as indicated by Brunauer-Emmett-Teller (BET) isotherms (surface area of 310.8+/−0.5 m2 g−1). This is notable because the combination of electrostatic stabilization and subsequent fission effectively suppresses the formation of large clumps, enabling 3D integration with HER accessible surfaces and micropores. Moreover, in a similar fashion to a paper ball, the deformed ridges made of several overlapping folds form structural supports that prevent the c-MoS2 from unfolding or collapsing when subjected to external stimuli. Indeed, structurally robust c-MoS2 can sustain common material processing conditions, such as solution processing, high temperature annealing, or the combination of both, and still maintain a porous, nanotextured morphology necessary for preserving efficient transport of H+ and H2 to the edge sites and activated basal plane under continuous electrochemical operation or elevated temperate. Together with the strained S vacancies, these structural features, revealed by the Raman, BET, SEM, and TEM characterizations described above, indicate that c-MoS2 can be used as a catalytically active, electrochemically and thermally stable component for HER.
  • The use of an EHD process offers the distinct advantage to directly integrate c-MoS2 into diverse device architectures, as well as onto a variety of substrates, including thin films of carbon nanotubes (CNTs), silicon, and even nanostructured polystyrene, in both monolithic and a selective fashion. To evaluate the catalytic performance of c-MoS2 for HER, electrodes were formed by depositing thin films of c-MoS2 onto CC via EHD deposition, as described above. The substrate had an area of 1 cm2 and the total aerial mass loading was ≈13 μg cm−2. In FIGS. 7(a)-(b), it can be seen that densely packed c-MoS2 assemblies are uniformly distributed throughout the entire carbon clothes (CC). FIG. 7(c) further suggests that mesoporous c-MoS2 structures remains intact and retains a high degree of crumpling even upon continuous deposition. In addition, corresponding energy-dispersive X-ray (EDX) mapping of relevant elements (not shown), including Mo and S, closely overlap the contours and areas of c-MoS2 and underlying CC substrates.
  • The HER cathodes made of c-MoS2 supported on CC were examined using a standard three-electrode electrochemical configuration in 0.5 m H2SO4 electrolyte de-aerated with Ar. Following deposition and prior to measurement, all samples were annealed at 250° C. for 1 h to ensure complete conversion of remnant 1T′ phases to 2H phases. This provides a well-defined system to unambiguously explore the correlation between the impact of 3D nanostructures, strain-induced modulation, and the catalytic activity in phase pure c-MoS2. To this end, linear sweep voltammograms (LSVs) were measured from thin films of planar 2H ce-MoS2 standard (S vacancies, strain: 0%, and heavily aggregated morphology), c-MoS2 synthesized at 100° C. (S vacancies, strain: 0%, and rippled morphology), c-MoS2 synthesized at 200° C. (S vacancies, strain: 1.2%, and partially crumpled morphology), and c-MoS2 synthesized at 250° C. (S vacancies, strain: 3.7%, and 3D fully crumpled morphology), along with the bare CC support as a reference, as shown in FIG. 8. CC electrodes did not show any pronounced cathodic current, and remained inactive during LSV scans. The planar 2H ce-MoS2 standard electrodes showed very poor HER characteristics even with a high density of S vacancies that have shown to activate the basal plane, delivering an exceedingly high η10 of ≈−364 mV. This poor HER performance can be attributed to the lack of electronic coupling between the substrates and active sites in combination with a heavily aggregated morphology that limits the exposure of activated basal plane toward HER. See H. Li et al., Nat. Mater. 15, 48 (2016). Meanwhile, increasing the deposition temperature to 100° C. led to the formation of a rippled morphology and η10 was reduced to −340 mV. The η10 of the c-MoS2 synthesized at 200° C. was significantly reduced to −309 mV due to the presence of wrinkled morphology and moderate strain load (≈1.2%). Importantly, c-MoS2 synthesized at 250° C., which was characterized by a combination of high strain load (≈3.7%) and 3D, fully crumpled nanostructures, displayed substantially improved total electrode metrics, including η10 of −191 mV, exchange current densities of 10.4 μA cm−2, and high cathodic current densities.
  • Further insights into function were obtained by extracting slopes from the Tafel plots shown in FIG. 9. Tafel slopes were found to decrease with increasing magnitude of crumpling, and were correlated with the strain load. The lowest Tafel slope of 64 mV per dec was obtained for the c-MoS2 synthesized at 250° C., suggesting a Heyrovsky mechanism at the surface of c-MoS2 with the hydrogen desorption reaction as the limiting step: Hads+H3O++e→H2+H2O. See J. D. Benck et al., ACS Catal. 4, 3957 (2014); M. A. Lukowski et al., J. Am. Chem. Soc. 135, 10274 (2013); B. E. Conway and B. V. Tilak, Electrochim. Acta 47, 3571 (2002); and C. Tsai et al., Surf. Sci. 640, 133 (2015). Note that total electrode activities are typically commensurate with the catalyst loading, as shown in Table 1. See J. D. Benck et al., ACS Catal. 4, 3957 (2014). Thus, from a practical perspective, the ability to deliver compelling total electrode characteristics on a HER electrode with very low loadings of c-MoS2 catalysts per geometric area (≈13 μg cm−2) may be transformative and economically viable.
  • The active site numbers were further estimated using the underpotential deposition (UPD) method. See A. Y. Lu et al., Small 12, 5530 (2016); and C. L. Green and A. Kucernak, J. Phys. Chem. 2, 11446 (2002). The density of active sites was determined to be approximately 1×1014 sites cm−2 for planar 2H ce-MoS2, increased to 5.7×1016 sites cm−2 for planar 1T′ ce-MoS2 polymorph, and further increased to 5.86×1017 sites cm−2 for 2H c-MoS2, thus corroborating that the advent of the hierarchically structured, 3D morphology of c-MoS2 from the intrinsically planar morphology of ce-MoS2 and enabling infiltration of electrolyte for exposure to both activated basal plane and edges. See J. Kibsgaard et al., Nat. Mater. 11, 963 (2012); D. Kong et al., Nano Lett. 13, 1341 (2013); and H. Wang et al., ACS Nano 8, 4940 (2014). The effective electrochemically active surface area, deduced from the electric double-layer capacitance (EDLC), increased by 12× when the c-MoS2 was deposited at 100° C., 24× when the c-MoS2 was deposited at 200° C., and by 125x when deposited at 250° C. relative to that of planar 2H ce-MoS2 annealed at 250° C. Meanwhile, electrochemical impedance spectroscopy (EIS) also suggested that the 3D porous morphology facilitates efficient transport across the electrode/electrolyte interfaces as indicated from the substantially reduced Rct (13Ω of 2H c-MoS2) relative to that of the planar 2H ce-MoS2 (Rct=275Ω, Table 1). Additionally, the measured Rs was significantly reduced from 38Ω (planar 2H ce-MoS2) to 1.5Ω (2H c-MoS2). This is unexpected, since the 2H c-MoS2 does not comprise locally transformed phases and atomically sharp interfaces to lower the contact resistance. See D. Voiry et al., Nat. Mater. 15, 1003 (2016); and D. Voiry et al., Nano Lett. 13, 6222 (2013). Even with the much enhanced electrochemically active surface area, the electronic coupling between the substrate and active sites will be sluggish unless selectively forming the electrically addressable interfaces through heterogeneous doping. Alternatively, both experimental demonstration and theoretical prediction have confirmed that regions of atomically localized curvature lead to a loss of spatial overlap of the d orbitals with neighboring Mo atoms and a shift in hybridization of Mo—S bonding from mixed ionic-covalent bonding between Mo and S atoms. See L. Yu et al., Nano Lett. 16, 2444 (2016); and M. Mavrikakis et al., Phys. Rev. Lett. 81, 2819 (1998). Notably, under applied strain, depletion of charges residing on Mo atoms is increased, thus leading to excess charges redistributed around S atoms. The net result of these orbital effects is a drastic increase in energy locally and reduced bonding-antibonding splitting. Consequently, the bandgap decreases, with the valence band maximum (VBM) and conductive band minimum (CBM) crossing the Fermi level, leading to a reduction of overall Rs similar to the semiconductor-to-metal transition and thus improved electronic coupling. See A. P. Nayak et al., Nat. Commun. 5, 1 (2014); and H. Li et al., J. Am. Chem. Soc. 138, 5123 (2016).
  • In parallel, using the measured active sites and exchange current densities, the TOF was calculated to infer the intrinsic catalytic activity per active sites. TOF was determined to be ≈0.130 s−1 at −200 mV vs. RHE for 2H c-MoS2 synthesized at 250° C., an order of magnitude higher than the 2H planar counterpart (0.016 s−1). These results suggest that the improvement of the HER performance comes from the combined effects of dimensional transition. First, the 3D porous morphology not only effectively exposes the initially buried active sites on both edges and basal plane but also facilitates effective transport across various interfaces between 2H c-MoS2, electrolyte and current collecting CC substrate. Second, the high strain load modulates the ΔGH and tailors the TOF at each active site. Specifically, the TOF achieved is comparable to or better than values measured for strain-vacancy engineered MoS2, edge sites of MoS2, vertical MoS2 flakes, 1T′ MoS2 nanoparticles, and [Mo3S4]4+ cluster of amorphous MoS2. See H. Li et al., Nat. Mater. 15, 48 (2016); T. F. Jaramillo et al., Science 317, 100 (2017); D. Kong et al., Nano Lett. 13, 1341 (2013); H. Wang et al., ACS Nano 8, 4940 (2014); and T. F. Jaramillo et al., J. Phys. Chem. C 112, 17492 (2008). Although not quite as active as the molecular thiomolybdate [Mo3S13]2−, c-MoS2 has advantages to synergistically tailor both intrinsic and total electrode activities in tandem with readily scalable synthesis, assembly, and processing routes. See J. Kibsgaard et al., Nat. Chem. 6, 248 (2014). Meanwhile, the loading of c-MoS2 was increased in an attempt to further improve the HER metrics through increasing numbers of active sites. However, HER metrics drastically decreased such that the thickest film in fact exhibited the poorest performance. From the comparison of all three specimens of different thicknesses, it is clear that the major limiting factors are the high η10 and low current densities. The high resistance may arise from the discontinuous pathways where outer layers of c-MoS2 are loosely attached to each other, as suggested by EDX mapping. The limited contacts at the top of the films serve as constriction points, thus diminishing electron accessibility and adversely affecting the overall HER performance. This limitation can be potentially mediated by adding conductive adducts, such as graphene, CNTs, or even 1T′ ce-MoS2, to simultaneously minimize charge transport distance and series resistance, while maintaining the accessible surface area.
  • Another important aspect for practical implementation of MoS2-based HER catalyst is the long-term stability under continuous operation and elevated temperatures. A durability study on the c-MoS2-based electrodes was first carried out by employing continuous LSV cycles at a scan rate of 5 mV s−1. The HER catalytic activity remains almost unchanged in the first 2500 cycles and displays a negligible decay after 5000 cycles. See M. A. Lukowski et al., J. Am. Chem. Soc. 135, 10274 (2013); and Y. Yin et al., J. Am. Chem. Soc. 138, 7965 (2016). In parallel, the electrochemical stability of the c-MoS2 catalyst was also evaluated by prolonged electrolysis at constant potentials. In accordance with the durability tests, the current density of c-MoS2 remained stable in 0.5 m H2SO4 for more than 100 h and showed a negligible change after 110 h. The uniformity and crumpled nanostructure of c-MoS2 was also examined under SEM and TEM after continuous linear potential sweeps, and appeared to be intact after the degradation measurements. In addition, when subjecting to thermal annealing at 300° C. under ambient conditions, unlike the temperature sensitive 1T′ counterparts, 2H c-MoS2, with in situ strain fields propagating on the thermodynamically favored 2H phase, remained catalytically active, revealing an unparalleled operational window up to 300° C. When the thermal annealing temperature was further increased to 350° C., a transition from MoS2 to MoO3 was observed as confirmed by XPS spectra. Nevertheless, the unprecedented combination of electrochemical and thermal stabilities not only suggests the preservation of both catalytic activity and structural integrity of c-MoS2, consistent with the structural robustness of c-MoS2, but also renders c-MoS2 more versatile for long-term utilization in energy conversion devices. Meanwhile, the chemically and mechanically robust nature of c-MoS2 provides flexibility in using chemical and mechanical means to further improve catalytic properties for HER. See Z. W. She et al., Science 355, 146 (2017).
  • In summary, the invention enables the formation of 3D structurally deformed c-MoS2 catalysts for electrochemically stable, thermally resistant, and highly efficient HER. The use of electrostatically charged droplets as nanoreactors facilitates colloidal dispersion, fission, and capillarity-induced-self-crumpling of 2D ce-MoS2, ultimately transforming it into 3D c-MoS2. As summarized in Table 1, electrochemical characterization indicates that HER electrodes composed of c-MoS2 catalysts synergistically integrate advantageous features of morphologically, chemically, electronically, and mechanically engineered MoS2 catalysts. See J. Kibsgaard et al., Nat. Mater. 11, 963 (2012); D. Kong et al., Nano Lett. 13, 1341 (2013); H. Li et al., Nat. Mater. 15, 48 (2016); D. Voiry et al., Nat. Mater. 15, 1003 (2016); H. Wang et al., ACS Nano 8, 4940 (2014); and Y. Yin et al., J. Am. Chem. Soc. 138, 7965 (2016). These include high surface area, efficient transport at interfaces, reduced series resistance, exposed basal plane with increased density of active sites, improved TOF per active site, chemical and mechanical robustness. The invention simultaneously achieves state-of-art metrics while addressing long-term stability under continuous operation as well as exposure to high temperature treatment while using a fully scalable and easy to integrate process. Given the wide variety and availability of chemically exfoliated 2D functional materials, including graphene, transition metal dichalcogenide (TMD), and the emerging MXenes, structurally robust, thermally stable, electronically heterogeneous, and catalytically active, multifunctional hybrid nanocomposites can now be readily, rapidly, and rationally assembled through this template free and scalable nanomanufacturing route. See J. Luo et al., ACS Nano 9, 9432 (2015). This enables applications beyond HER, including electrochemical capacitors, battery, catalysis, reactors and separation, drug delivery, biocompatible scaffolds, sensing and high complexity composites.
  • The present invention has been described as method to form crumpled transition metal dichalcogenide sheets. It will be understood that the above description is merely illustrative of the applications of the principles of the present invention, the scope of which is to be determined by the claims viewed in light of the specification. Other variants and modifications of the invention will be apparent to those of skill in the art.
  • TABLE 1
    Summary of HER characteristics of MoS2 catalysts with various nanostructures, strain S vacancy, edge and 1T′ phase engineering.
    η10, j0, Rs and Rct stand for overpotentials, exchange current densities, series and charge-transfer resistances, respectively.
    Total electrode properties Intrinsic properties
    Tafel Surface Mass Active
    η10 slope J0 Rs + Rct Capacitance area loading sites TOF Stability
    (mV) (mV/dec) (μA/cm2) (Ω) (mF/cm2) (m2/g) (μg/cm2) (sites/cm2) (s−1) Cycles Temp
    2H c-MoS2 (300° C.) −193 66 10.4 Rs = 1.7 11.0 310.5 13 5.7 × 1017 0.128 5000 300° C.
    Rct = 13
    2H c-MoS2 (250° C.) −191 64 10.4 Rs = 1.5 11.4 310.8 13 5.8 × 1017 0.130 5000 300° C.
    Rct = 13
    2H c-MoS2 (200° C.) −309 92 5.2 Rs = 17 2.2 48.6 13 2.0 × 1015 0.080 2000 300° C.
    Rct = 50
    2H c-MoS2 (100° C.) −340 107 4.5 Rs = 29 1.1 37.2 13 4.7 × 1014 0.052 1000 N/A
    Rct = 150
    2H planar MoS2 −364 108 0.7 Rs = 38 0.091 19.0 15 1.0 × 1014 0.016 <1000 N/A
    Rct = 275
    *2H porous MoS2 −218 62 10.5 Rct = 280 8.2 N/A 140 2.3 × 1014 N/A 1000 N/A
    **Vertically aligned 2H −400 105-120 2.2 Rs = 2.3 N/A N/A 8.5 5.4 × 1014 0.013-0.016 1000 N/A
    MoS2 Rct > 20
    ***Double gyroid MoS2 −206 50 0.69 N/A 4.8 N/A 60 N/A N/A N/A N/A
    ****Strained 2H-MoS2 −170 60 50-60 N/A N/A N/A N/A 3.1 × 1014 0.08-0.31 N/A N/A
    *1T′ Porous MoS2 −153 43 15.8 Rct = 16 63.1 N/A 140 5.8 × 1016 0.5 1000 N/A
    nanosheets (η = 15
    3 mV)
    *****1T′ MoS2 −118 66 50 Rs = 18 345 N/A 3,400 6.4 × 1018 0.10 >5000 N/A
    nanoparticle Rct > 5
    ******1T′ MoSe2 −152 52 N/A Rct = 16 25.2 N/A 140 N/A N/A 1000 N/A
    *******1T′ WS2 −142 70 N/A Rct = 5 48 N/A 1000 N/A N/A <500 N/A
    *Y. Yin et al., J. Am. Chem. Soc. 138, 7965 (2016);
    **D. S. Kong et al., Nano Lett. 13, 1341 (2013);
    ***J. Kibsgaard et al., Nat. Mater. 11, 963 (2012);
    ****H. Li et al., Nat. Mater. 15, 48 (2016);
    *****H. Wang et al., Acs Nano 8, 4940 (2014);
    ******Y. Yin et al., Adv. Mater. 29, 1700311 (2017);
    *******M. a Lukowski et al., Energy Environ. Sci. 7, 2608 (2014).

Claims (14)

We claim:
1. A method of crumpling transition metal dichalcogenide sheets, comprising:
forming an electrostatically charged liquid droplet comprising a plurality of planar sheets of a transition metal dichalcogenide;
applying a sufficient electric field to the droplet to cause the droplet to rupture into a plurality of fission droplets by an electrohydrodynamic process, wherein each fission droplet comprises at least one planar sheet of the transition metal dichalcogenide; and
evaporating the remaining liquid from the plurality of fission droplets, thereby inducing spontaneous self-crumpling of the at least one planar sheet of the transition metal dichalcogenide in each of the fission droplets.
2. The method of claim 1, wherein the transition metal comprises molybdenum or tungsten.
3. The method of claim 1, wherein the chalcogenide comprises sulfur, selenium, or tellurium.
4. The method of claim 1, wherein the transition metal dichalcogenide comprises molybdenum disulfide.
5. The method of claim 4, wherein the planar sheets of molybdenum disulfide comprise 2H phase sheets having basal plane defects and whereby the self-crumpling activates the basal plane defects.
6. The method of claim 1, wherein the planar sheets of the transition metal dichalcogenide comprise exfoliated sheets.
7. The method of claim 1, wherein the forming an electrostatically charged liquid droplet comprises feeding a solution of the planar sheets of the transition metal dichalcogenide through a spinneret.
8. The method of claim 1, wherein the sufficient electric field is greater than approximately 1.25 kV/cm.
9. The method of claim 1, wherein the evaporating comprises depositing the plurality of fission droplets on a substrate above a deposition temperature.
10. The method of claim 9, wherein the substrate comprises a hydrogen evolution reaction substrate.
11. The method of claim 9, wherein the substrate comprises carbon or silicon.
12. The method of claim 9, wherein the deposition temperature is greater than 200° C.
13. The method of claim 9, wherein the deposition temperature is greater than 250° C.
14. A method for electrochemical hydrogen evolution, comprising exposing an aqueous electrolyte to a crumpled transition metal dichalcogenide in an electrochemical cell at a sufficient cell voltage to catalyze the electrolysis of water and evolve hydrogen.
US16/010,650 2017-06-29 2018-06-18 Crumpled Transition Metal Dichalcogenide Sheets Abandoned US20190003064A1 (en)

Priority Applications (1)

Application Number Priority Date Filing Date Title
US16/010,650 US20190003064A1 (en) 2017-06-29 2018-06-18 Crumpled Transition Metal Dichalcogenide Sheets

Applications Claiming Priority (2)

Application Number Priority Date Filing Date Title
US201762526828P 2017-06-29 2017-06-29
US16/010,650 US20190003064A1 (en) 2017-06-29 2018-06-18 Crumpled Transition Metal Dichalcogenide Sheets

Publications (1)

Publication Number Publication Date
US20190003064A1 true US20190003064A1 (en) 2019-01-03

Family

ID=64737934

Family Applications (1)

Application Number Title Priority Date Filing Date
US16/010,650 Abandoned US20190003064A1 (en) 2017-06-29 2018-06-18 Crumpled Transition Metal Dichalcogenide Sheets

Country Status (1)

Country Link
US (1) US20190003064A1 (en)

Cited By (7)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
CN110219017A (en) * 2019-06-06 2019-09-10 燕山大学 Molybdenum disulfide/expanded graphite hydrogen-precipitating electrode preparation method
CN112194185A (en) * 2020-10-13 2021-01-08 西南科技大学 Defect-rich ultrathin two-dimensional nano MoS2Preparation method
CN113174580A (en) * 2021-04-15 2021-07-27 清华大学 Humidity-based film wrinkle shape regulation and control method
CN115124079A (en) * 2022-06-28 2022-09-30 陕西科技大学 VS rich in sulfur vacancy defect 2-x Material, preparation method and application thereof
CN115231616A (en) * 2022-07-19 2022-10-25 南京大学 Method for preparing molybdenum disulfide micropore pattern without mask
US11484867B2 (en) 2020-11-10 2022-11-01 National Technology & Engineering Solutions Of Sandia, Llc Electrocatalyst comprising a crumpled transition metal dichalcogenide support loaded with monodispersed metal nanoparticles
CN116425199A (en) * 2023-05-04 2023-07-14 昆明理工大学 Preparation method of molybdenum disulfide nano-coil electrocatalyst

Cited By (7)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
CN110219017A (en) * 2019-06-06 2019-09-10 燕山大学 Molybdenum disulfide/expanded graphite hydrogen-precipitating electrode preparation method
CN112194185A (en) * 2020-10-13 2021-01-08 西南科技大学 Defect-rich ultrathin two-dimensional nano MoS2Preparation method
US11484867B2 (en) 2020-11-10 2022-11-01 National Technology & Engineering Solutions Of Sandia, Llc Electrocatalyst comprising a crumpled transition metal dichalcogenide support loaded with monodispersed metal nanoparticles
CN113174580A (en) * 2021-04-15 2021-07-27 清华大学 Humidity-based film wrinkle shape regulation and control method
CN115124079A (en) * 2022-06-28 2022-09-30 陕西科技大学 VS rich in sulfur vacancy defect 2-x Material, preparation method and application thereof
CN115231616A (en) * 2022-07-19 2022-10-25 南京大学 Method for preparing molybdenum disulfide micropore pattern without mask
CN116425199A (en) * 2023-05-04 2023-07-14 昆明理工大学 Preparation method of molybdenum disulfide nano-coil electrocatalyst

Similar Documents

Publication Publication Date Title
Chen et al. Structurally deformed MoS2 for electrochemically stable, thermally resistant, and highly efficient hydrogen evolution reaction
US20190003064A1 (en) Crumpled Transition Metal Dichalcogenide Sheets
Yang et al. Self-templated growth of vertically aligned 2H-1T MoS2 for efficient electrocatalytic hydrogen evolution
Huang et al. Synergistically modulating electronic structure of NiS2 hierarchical architectures by phosphorus doping and sulfur-vacancies defect engineering enables efficient electrocatalytic water splitting
Chen et al. Superior electrocatalysis for hydrogen evolution with crumpled graphene/tungsten disulfide/tungsten trioxide ternary nanohybrids
He et al. Molybdenum disulfide nanomaterials: Structures, properties, synthesis and recent progress on hydrogen evolution reaction
Ejigu et al. A simple electrochemical route to metallic phase trilayer MoS 2: evaluation as electrocatalysts and supercapacitors
Jiao et al. Defect-rich one-dimensional MoS2 hierarchical architecture for efficient hydrogen evolution: Coupling of multiple advantages into one catalyst
Sun et al. Oxygen-incorporated MoS2 microspheres with tunable interiors as novel electrode materials for supercapacitors
Prasannachandran et al. Functionalized phosphorene quantum dots as efficient electrocatalyst for oxygen evolution reaction
Zhao et al. Controllable nanoscale engineering of vertically aligned MoS2 ultrathin nanosheets by nitrogen doping of 3D graphene hydrogel for improved electrocatalytic hydrogen evolution
Zheng et al. Space-confined growth of MoS2 nanosheets within graphite: the layered hybrid of MoS2 and graphene as an active catalyst for hydrogen evolution reaction
Zhang et al. MoSe 2 porous microspheres comprising monolayer flakes with high electrocatalytic activity
Zhang et al. Water-soluble defect-rich MoS2 ultrathin nanosheets for enhanced hydrogen evolution
Luo et al. Defect engineering of carbons for energy conversion and storage applications
Le et al. Controllably doping nitrogen into 1T/2H MoS2 heterostructure nanosheets for enhanced supercapacitive and electrocatalytic performance by low-power N2 plasma
Sun et al. Size-controlled MoS 2 nanodots supported on reduced graphene oxide for hydrogen evolution reaction and sodium-ion batteries
Sun et al. Tungsten disulfide-based nanomaterials for energy conversion and storage
Tian et al. Improving the activity of Co x P nanoparticles for the electrochemical hydrogen evolution by hydrogenation
Cheng et al. In situ hydrothermal synthesis MoS2/guar gum carbon nanoflowers as advanced electrocatalysts for electrocatalytic hydrogen evolution
Lv et al. One-pot hydrothermal approach towards 2D/2D heterostructure based on 1 T MoS2 chemically bonding with GO for extremely high electrocatalytic performance
Wang et al. Addressable surface engineering for N-doped WS 2 nanosheet arrays with abundant active sites and the optimal local electronic structure for enhanced hydrogen evolution reaction
Qiao et al. Ultrathin MoSSe alloy nanosheets anchored on carbon nanotubes as advanced catalysts for hydrogen evolution
Adarakatti et al. MoS2-graphene-CuNi2S4 nanocomposite an efficient electrocatalyst for the hydrogen evolution reaction
Vikraman et al. Facile synthesis of molybdenum diselenide layers for high-performance hydrogen evolution electrocatalysts

Legal Events

Date Code Title Description
AS Assignment

Owner name: U.S. DEPARTMENT OF ENERGY, DISTRICT OF COLUMBIA

Free format text: CONFIRMATORY LICENSE;ASSIGNOR:NATIONAL TECHNOLOGY & ENGINEERING SOLUTIONS OF SANDIA, LLC;REEL/FRAME:046824/0815

Effective date: 20180822

AS Assignment

Owner name: NATIONAL TECHNOLOGY & ENGINEERING SOLUTIONS OF SAN

Free format text: ASSIGNMENT OF ASSIGNORS INTEREST;ASSIGNOR:CHOU, STANLEY SHIHYAO;REEL/FRAME:047065/0837

Effective date: 20180910

STPP Information on status: patent application and granting procedure in general

Free format text: DOCKETED NEW CASE - READY FOR EXAMINATION

STPP Information on status: patent application and granting procedure in general

Free format text: NON FINAL ACTION MAILED

STCB Information on status: application discontinuation

Free format text: ABANDONED -- FAILURE TO RESPOND TO AN OFFICE ACTION