With global water use intensifying and the effects of climate change increasingly salient, water scarcity has become a primary international concern. In response, communities have turned to sea and brackish water to provide a steady flow of potable water to their constituents. In its current form, however, desalination requires an order of magnitude more energy than freshwater treatment. Membranes occupy a central role in reverse osmosis (RO) filtration of salt water; increasing membrane water permeability and salt selectivity would result in lower energy requirements, fewer processing steps, and an overall reduction in the capital cost of desalination facilities.
Restacked two dimensional (2D) materials, which are assemblies of individual atomically thin sheets with their basal planes lying parallel to each other, comprise a new class of nanofiltration membranes that show great promise as efficient separators of ions and small molecules from water. The high aspect ratios of 2D materials determine the macroscopic geometries of these membranes: laminar structures with nanometer-scale spaces between layers. In a filtration device, water flows through the channels between layers of material with little obstruction, while ions and other small molecules are excluded. Graphene and graphene oxide form the basis for the first molecular sieves but the breadth of constituent materials has grown in recent years to include, among others, boron nitride, MXenes and transition metal dichalcogenides (TMDs) such as WS2 and MoS2.
Chemically exfoliated MoS2 (ce-MoS2) is a particularly viable candidate for RO desalination. The channel width of restacked ce-MoS2 is on the appropriate length scale for size-based exclusion of ions while facilitating high water flux. A result of its mild hydrophilicity, water molecules interact weakly with the ce-MoS2 surface. This weak interaction leads to a higher water flux relative to its strongly hydrophilic graphene oxide counterpart. Furthermore, ce-MoS2 has greater structural stability relative to graphene oxide; when soaked in water, the interlayer spacing of ce-MoS2 membranes remains stable at ˜1 nm, whereas graphene oxide membranes swell and easily disintegrate. Recent efforts have demonstrated the viability of horizontally aligned ce-MoS2 as an RO membrane. (Sun, L., et al., Chem. Commun. 49, 10718-10720 (2013); Hirunpinyopas, W. et al., ACS Nano 11, 11082-11090 (2017); Zheng, S., et al., ACS Nano 11, 6440-6450 (2017); Deng, M., Kwac, K., Li, M., Jung, Y. & Park, H. G. Stability, molecular sieving, and ion diffusion selectivity of a lamellar membrane from two-dimensional molybdenum disulfide. Nano Lett. 17, 2342-2348 (2017).) However, the association between the membrane's hydration dependent structure and filtration performance has remained underexplored.
Illustrative embodiments of the invention will hereafter be described with reference to the accompanying drawings, wherein like numerals denote like elements.
Molybdenum disulfide membranes for ionic and/or molecular filtration applications are provided. The membranes have high separation performance, including high water flux and high molecule/ion rejection, and do not need to be stored in a hydrated condition in order to enable their reuse.
The membranes are based on stacked MoS2 sheets having small hydrophilic organic functional groups covalently bound thereto. The functional groups serve to reduce or eliminate mesoporous voids between MoS2 sheets, to provide a more uniform and optimal interlayer spacing, and to render the membranes reusable after drying.
The covalently functionalized MoS2 sheets can be made from exfoliated MoS2 (e.g., chemically exfoliated MoS2), whereby the exfoliated MoS2 sheets are covalently functionalized and then restacked to form a membrane comprising vertically stacked, horizontally aligned sheets of MoS2 with covalently bound organic functional groups intercalated between the sheets. A schematic illustration of a membrane is shown in
Without intending to be bound to any particular theory of the invention, the inventors observe that the organic functional groups can prevent regions of local impermeability from forming in the membranes upon drying, due to irreversible restacking of the MoS2 sheets into a bulk stacked morphology. This is illustrated in
In contrast, the organic functional groups used in the membranes described herein allow for the closing of mesoscale pores, while preventing the MoS2 sheets in the membrane from irreversibly collapsing into the bulk state during drying, where a bulk state can be detected using XRD, as illustrated in the Example. Thus, the present membranes can be dried to remove the mesoscale voids without sacrificing water permeability (
In addition, the hydrophilic organic functional groups promote the swelling of the membranes in water and may tune the structure of water within the MoS2 channel to provide an interlayer spacing between the MoS2 sheets that promotes both high water flux and high ion/molecule rejection when the membranes are in a hydrated state. For example, interlayer spacings between the MoS2 sheets in neighboring layers in the range from 11 Å to 12 Å can be achieved.
For the purposes of this disclosure, an organic functional group may be considered hydrophilic if a MoS2 membrane that is functionalized with the organic functional groups has a hydrophilicity that is the same as, or greater than, that of the unfunctionalized MoS2. Hydrophilicity can be measured via water contact angle measurements, as described in the Example, where a greater hydrophilicity corresponds to a lower water contact angle. By way of illustration, membranes functionalized with hydrophilic organic functional groups may have water contact angles of 65° or lower, including 60° or lower. Organic functional groups that impart a net negative charge at neutral pH (pH=7) to the membranes can be used as hydrophilic functionalities.
Acetic acid and nitrile groups are examples of hydrophilic organic functional groups that can be used to covalently functionalize the MoS2 membranes.
The degree of membrane functionalization should be sufficient to provide adequate water flux and molecule and/or ion rejection properties for the intended filtration application. By way of illustration, some embodiments of the MoS2 membranes have a degree of organic (e.g., acetic acid and/or nitrile) functionalization of at least 10%, including embodiments having a degree of functionalization of at least 20%. For example, some of the MoS2 membranes have a degree of organic (e.g., acetic acid and/or nitrile) functionalization in the range from about 20% to about 40%. However, degrees of functionalization outside of this range can be used. As used herein, the degree of functionalization refers to the ratio of functional groups to molybdenum atoms.
The membranes can be incorporated into filtration devices for the separation and removal of a variety of ions and/or small molecules from liquid samples, particularly aqueous liquid samples. In a filtration device, the sample is flowed through the channels defined between the layers of aligned MoS2 sheets with little obstruction, while ions and other small molecules having sizes greater than the channel widths are excluded. The filtration devices may further include various components that are common to such devices, such as a housing, a membrane support, an input chamber, an output chamber, and valves and/or pumps to control liquid flow, wherein the membrane is disposed between the input and output chambers. Once the filtration is complete, the excluded ions and/or molecules may be collected and removed from the filtration device.
The removal of salt ions, such as Na+ and/or Cl−, from salinized water is one application for which the membranes are well-suited. The salinized water may be from a natural body of water (e.g., seawater) or wastewater from an industrial or municipal plant. Other ions that can be filtered using the membranes include, but are not limited to, inorganic anions and/or cations, such as K+, Ca+, Mg2+, and SO42+ and small organic molecules.
The molybdenum disulfide membranes can be made by forming an aqueous suspension of exfoliated MoS2 sheets and adding an organohalide compound, such as an acetate halide (e.g., iodoacetic acid) or a nitrile halide (e.g., iodonitrile) to the aqueous suspension, whereby the organohalide compound reacts with the MoS2 sheets to form MoS2 sheets that are covalently functionalized with organic groups from the organohalide compound. The suspension can then be filtered on a porous polymer substrate, such as cellulose ether, to form a supported membrane comprising stacked sheets of the covalently functionalized MoS2 sheets. The membrane is then dried and delaminated from the porous polymer substrate to form a free-standing membrane. The MoS2 particles may be exfoliated using, for example, lithium intercalation. The suspension can be filtering using, for example, vacuum filtration. The membrane can be delaminated from the porous polymer substrate by, for example, submerging the membrane and the substrate in water.
This Example describes the results of an array of tests that probed the structure of MoS2 membranes on a wide range of length scales. From powder x-ray diffraction (XRD) and scanning electron microscopy (SEM) studies, paired with reverse osmosis (RO) tests, it was found that the physical structure of MoS2 membranes evolved with the hydration level (determined by the membrane drying time) at both ˜1 nm (microporous) and ˜100 nm (mesoporous) length scales.
This Example describes the water structure that determines the microporous scale membrane morphology, as well as the diffusion of water molecules, at a range of interlayer spacings and surface chemistries using molecular dynamics (MD) simulations. The MD simulations predicted that a bilayer of water occupies the membrane channels at the interlayer spacing that were measured experimentally for ce-MoS2 and acetate-MoS2, but that a smaller interlayer spacing with only a single layer of water was more stable for amide-MoS2. It was also found that water diffusion in the membrane was tuned by surface functionalization, but that this had a modest effect on overall water flux relative to the changes in the membrane structure on the micro and mesoporous scales that come with functionalization.
An aqueous suspension of ce-MoS2 flakes were synthesized following the standard lithium-intercalation and exfoliation procedure outlined in the Methods section of this disclosure. (Eda, G. et al., Nano Lett 11, 5111-5116 (2011); Joensen, P., et al., Materials Research Bulletin 21, 457-461 (1986); and Zheng, J. et al., Nat Commun 5, 2995 (2014).) The detailed physical characteristics of individual flakes are provided in Section 1 of Supplementary Information. The flakes were generally 100-500 nm in lateral size and only a few nm in total thickness. Membranes were assembled via vacuum filtration on a porous polymer substrate (mixed cellulose ester, 25 nm average pore diameter).
ce-MoS2 sheets were covalently functionalized with two small organic molecules: iodoacetic acid and iodoacetamide. The procedure provided in recent studies and outlined in the Methods section of this disclosure was followed to graft these organic molecules on MoS2 flakes. (Paredes, J. I. et al., ACS Appl. Mater. Interfaces 8, 27974-27986 (2016); Voiry, D. et al. Nature Chem 7, 45-49 (2015).) The mechanisms underlying the covalent functionalization are outlined in Section 2 of Supplementary Information. Importantly, the net negative surface charge of ce-MoS2 (˜0.25 electrons per Mo atom) was neutralized during functionalization. Acetic acid, however, deprotonated in neutral pH and induced a net negative charge on the sheet of equal magnitude to that of ce-MoS2; the sheets remained neutral for amide-MoS2.
The fraction of organic fragments decorating the MoS2 surface was determined via X-ray photoelectron spectroscopy (XPS) of samples drop cast on Si wafers. The degree of functionalization could be obtained by deconvolving the S2p region (
The structure of MoS2 membranes on the microporous scale was determined via a combination of XRD and TEM. The interlayer spacing, the parameter most directly affecting ion separation performance, was determined via XRD to be 6.2 Å (bulk-like) for dried ce-MoS2; as shown in
Flakes of acetate-MoS2, characterized by TEM in
Direct visualization of the hydrated MoS2 membrane structure was accomplished using a standard freeze drying procedure, outlined in Section 4 of Supplementary Information. This procedure allows direct visual comparison between dried ce-MoS2 (
The evolving mesoporous scale morphology is accompanied by evolving structure on the microporous scale. As shown in the XRD spectra in
The evolving structure of ce-MoS2 membranes dramatically affects separation performance, as indicated by the dependence of water flux and ion rejection on membrane drying time. To detail this association, membranes were dried for a defined period in room temperature conditions (humidity 20-30%) after fabrication, then loaded in a stirred, pressure-assisted dead-end filtration cell for tests. All samples were tested in brackish water conditions with 17 mM (˜2500 ppm) Na2SO4 under pressures of 150 psi (10.3 bar). The performance metrics were measured until the ion rejection stabilized.
The flux through ce-MoS2 membranes decreased by two orders of magnitude as a function of drying time, a result of the closing of voids and restacking to bulk (
The drying-dependent performance of acetate-MoS2 differed qualitatively from that of ce-MoS2. The ion rejection plateaued more slowly for acetate-MoS2 membranes (
The dependence of water flux on MoS2 surface treatment was a result of differing water diffusion constants within membrane channels and varying membrane porosities. Diffusion constants were calculated via MD simulations discussed in the subsequent section; porosity was partly determined by the widths of the MoS2 channel, which can be measured by XRD. As depicted in
To elucidate the factors contributing to water diffusion in ce- and functionalized MoS2, MD simulations were conducted on single channels formed by parallel MoS2 sheets. By characterizing the water structure and dynamics at a sequence of interlayer spacings for ce-, acetate- and amide-MoS2, the entire parameter space attainable was covered in the experiments. The setup shown in
To compare water flux between samples, the diffusion constant (derived from the mean squared displacement of water molecules in the MoS2 channel) was computed for interlayer spacings from 10 to 17 Å. It was found that for a 12 Å interlayer spacing, the diffusion constant varied with surface chemistry as ce-MoS2>amide-MoS2>acetate-MoS2 (
The equilibrium spacing between MoS2 sheets in an aqueous medium governed the ion rejection and water flux through the membranes assembled from those sheets. The equilibrium spacing was determined by the complex interplay of electrostatic and Van der Waals interactions, in addition to hydration forces arising from the confinement of water in the MoS2 channel. To quantify the sum of these interactions, umbrella sampling was used to calculate the potential of mean force (PMF) as a function of interlayer spacing (
The PMF profiles depicted in
Given that hydrated ce- and acetate-MoS2 membranes have experimental interlayer spacings of 12 Å, the density profiles in
It was shown that MoS2 membranes demonstrate a range of separation performances that depend on the degree of hydration: at short drying times, ce-MoS2 membranes showed very low rejection but high water permeance; at intermediate drying times, ce-MoS2 membranes demonstrated high rejection with a moderate water flux; when completely dried they were impermeable. Membrane functionalization is a route towards consistently achieving high salt rejection and water flux, as was demonstrated with acetate functionalized MoS2 membranes. The more hydrophobic acetamide functionalized MoS2 membranes demonstrated lower separation performance, perhaps a result of the absence of surface charge. MD simulations revealed the structure of water within the membrane that forms the basis for its hydration dependent structure. In addition, the trends derived from these MD simulations corroborate the dominance of the rehydration behavior of MoS2 membranes in separation tests and emphasize the importance of water structure on both the microporous and mesoporous scales. This Example provides structural and chemical information of MoS2 membranes, providing insight into their behavior as separation membranes or when employed in nanofluidic platforms or even as electrode materials.
The degree of MoS2 functionalization was computed by deconvolving XPS spectra with a sum of Gaussian/Lorentzian lineshapes in the S2p, N1s, and C1s regions. First, however, the Mo3d region was analyzed to determine the approximate fraction of 2H and 1T phases (
The S2p regions were deconvolved in a similar fashion to that of the Mo3d region (
Finally, results were confirmed by analyzing the C1s region (
8.6
4.3
8.8
4.4
21.0
The experimental measurement of water flow through the membrane is water flux, defined by the flow rate per unit area of membrane. The simplest model of a membrane is an array of straight, unconnected pores extending through the entirety of its bulk. Using this model as a starting point, the flux J through the membrane is determined by the flux through each pore Jp times the porosity of the membrane:
where Np is the number of pores in the membrane, Ap is the area of each pore, Am is the area of the full membrane and the factor
is the membrane porosity. The geometry of a MoS2 membrane is more complicated than the simple picture described above, however. To build a model, a single MoS2 flake was considered at the surface of a membrane; here, the area available for water permeation was proportional to the space between the flake and the membrane times the circumference of the flake. This area was taken to represent a pore in this system, so
A
p˜(d−dMoS
where d is the interlayer spacing, dMoS
Putting Supplementary Eqs. 1-3 together gave the following result:
The flux through a single pore Jp has contributions from the entrance and exit of the pore, as well as the transport through the length of the pore. In the limit where entrance and exit rates are fast relative to intra-pore transport, the flux is related solely to the pressure drop and diffusion constant. This can be seen using Fick's law and the Van't Hoff equation. Fick's law states that the one-dimensional osmotic flux (in this specific case Jp) due to a concentration gradient
J
p
=−D
where D is the diffusion constant and C is a number concentration. The inventors were interested in flow driven by pressure rather than an osmotic gradient. The osmotic pressure drop generated by a concentration difference is given by the Van't Hoff equation in the low concentration limit
ΔP=kBTΔC (6)
where kB is Boltzmann's constant, T is the temperature and P is the applied pressure. The concentration difference ΔC is measured between two reservoirs on opposite sides of the membrane. The average concentration gradient is given by
where L is the distance between the two reservoirs or the thickness of the membrane. Combining Supplementary Eqs. 5-7,
Plugging Supplementary Eq. 8 into Supplementary Eq. 4 gave the following relationship between the experimentally measured flux and the diffusion constant computed through MD simulations:
In a real membrane, the pores are interconnected in complicated ways. Two assumptions can remove this difficulty, however. First, it was assumed that the porosity of each layer is the same; second, it was assumed that the connections between layers permit fast transport, just like the entrances and exits of the channels. Finally, the distance a water molecule must traverse between two reservoirs is much further than the thickness of the membrane, due to the tortuosity of the path. This can be accounted for by simply replacing L with an effective length Leff.
The 1T MoS2 lattice structure was taken from Py and Haering. (Py, M. A. et al., Can: J. Phys. 61, 76-84 (1983).) Exclusively 1T MoS2 were used in the simulations instead of 2H MoS2 because the simulations were completed before the experiments had confirmed the presence of 2H MoS2. The MoS2 atoms interacted with the water molecules via Lennard-Jones and Coulomb interactions. The MoS2 partial charges came from Varshney et al. and the Lennard-Jones parameters came from Luan and Zhou with one exception: the Luan and Zhou force field was parameterized based on the water contact angle on 2H MoS2, which ranged from about 70° to 90° depending on the condition of the surface, whereas the simulations used 1T MoS2, which has a much smaller contact angle of about 28°. (Varshney, V. et al. Computational Materials Science 48, 101-108 (2010); Luan, B. Appl. Phys. Lett. 108, 131601 (2016); Kozbial, A., et al., Langmuir 31, 8429-8435 (2015); Acerce, M., et al., Nature Nanotechnology 10, 313-318 (2015).) Crucially, Zhang, Luan, and Zhou showed that the sulfur Lennard-Jones c parameter tunes the water contact angle linearly over a wide range, although not all the way down to 28°. (Zhang, L., et al., J. Phys. Chem. B 123, 7243-7252 (2019).) Their relationship was extrapolated to estimate the value of the sulfur Lennard-Jones ε parameter on 1T MoS2. Note that the contact angle of 1T MoS2 used for this extrapolation was different than the contact angle of ce-MoS2 measured here (
In simulations with acetate- and amide-functionalized MoS2 sheets, the acetate and acetamide groups were modeled using the DREIDING force field. (Mayo, S. L. et al., J. Phys. Chem. 94, 8897-8909 (1990).) The acetate partial charges were taken from Minofar et al., which used the restrained electrostatic potential (RESP) method to compute the partial charges. (Minofar, B. et al. J. Phys. Chem. B 110, 15939-15944 (2006).) The acetamide partial charges were computed with the same approach, using Hartree-Fock theory with a 6-31G* basis set and the antechamber program to compute the RESP charges. (Frisch, M. J. et al. Gaussian 16 Revision A. 03. (2016); Wang, J., et al., Journal of Computational Chemistry 25, 1157-1174 (2004).) Since the acetate groups have a −1 charge, each one was accompanied by a sodium ion to keep the system charge neutral. The sodium ions were modelled with the Joung and Cheatham potential. (Joung, I. S. et al., J. Phys. Chem. B 112, 9020-9041 (2008).)
In the charged MoS2 simulations, the excess charge was modelled by augmenting the partial charges of the sulfur atoms on the inner surfaces of the MoS2 sheets. Since 1T MoS2 is metallic, the excess charge was spread evenly over all the inner sulfur atoms. Charge neutrality was maintained by adding sodium ions, as in the acetate simulations. The partial charges on the sulfur atoms on the outside of the sheets were unchanged since there was only water on the inside of the channel, even though in the experiment the outer sulfur atoms would be equivalently charged due to the metallic nature of 1T MoS2. Likewise, the acetate and acetamide groups were only placed on the inner side of the MoS2 sheets, even though both sides were functionalized in the experimental system.
The simulation was periodic in all three dimensions. The simulation box was about 13×3×4 nm3 (
The system was held at a temperature of 298 K and a pressure of 1 atm using the Nosé-Hoover style algorithm of Shinoda, Shiga, and Mikami. (Nosê, S. Molecular Physics 52, 255-268 (1984); Hoover, W. G. Phys. Rev. A 31, 1695-1697 (1985); Shinoda, W., et al., Phys. Rev. B 69, 134103 (2004); Martyna, G. J., et al., J. Chem. Phys. 101, 4177-4189 (1994); Parrinello, M. et al., Journal of Applied Physics 52, 7182-7190 (1981).) The thermostat damping time was 0.1 ps and the barostat damping time was 1 ps. The system was only barostatted in the x-direction, and it was not barostatted based on the total pressure; instead, it was based on the pressure of the “bulk-like” water in the reservoirs. This avoided artifacts due to the exposed edges of the MoS2 sheets in the x-direction. The region of “bulk-like” water was defined as the region further than 1.2 nm from the edge of the MoS2 sheets. The pressure was computed in that region using the zeroth-order Irving-Kirkwood (local virial) approximation, which is valid in isotropic fluids. (Irving, J. H. et al., J. Chem. Phys. 18, 817-829 (1950).)
The dynamics were integrated using the velocity-Verlet algorithm. (Swope, W. C., et al., J. Chem. Phys. 76, 637-649 (1982).) The systems with unfunctionalized sheets used a 2 fs timestep and the acetate and acetamide functionalized systems used a 1 fs timestep, to accommodate the high frequency intramolecular bond and angle vibrations. The NH2 moieties of the acetamide groups were held rigid using the SHAKE algorithm. (Ryckaert, J.-P., et al., Journal of Computational Physics 23, 327-341 (1977).) This avoided the need for an even shorter timestep or a multi-timescale integrator to handle the fast vibrations of the light hydrogen atoms.
All long-range Coulomb interactions were evaluated using the particle-particle particle-mesh algorithm of Hockney and Eastwood. (Hockney, R. W. et al., SIAM Review 25, (1966).) The simulations were performed using the LAMMPS simulation package modified so that the barostat acted on the pressure in the bulk region. (Plimpton, S. Fast parallel algorithms for short-range molecular dynamics. Journal of Computational Physics 117, 1-19 (1995).)
The systems with unfunctionalized sheets were equilibrated for 0.2 ns and the functionalized sheets were equilibrated for 0.6 ns because it took longer for the spaces between the functional groups to fill in with water. The equilibration of the systems could be evaluated by checking that the volume of the system had stabilized. Data was then collected for 1 ns to compute the mean squared displacement.
To compute the PMFs, biased simulations were performed over the range of relevant interlayer spacings. The difference in interlayer spacing between adjacent simulations ranged from 0.01 Å to 0.2 Å. The simulations were initialized from the unbiased simulation (used to compute the diffusion constant) with the closest interlayer spacing. These unbiased simulations were performed at 1 Å intervals of interlayer spacing, so each biased simulation was initialized no more than 0.5 Å from its bias center. After initialization, the WCA wall particles were removed, the z-dimension of the box was extended, and the empty space was filled with water molecules. This reduced interactions between periodic images of the MoS2 sheets, which would be quite large at the initial z-dimension of about 4 nm. After this process, the final number of water molecules ranged from about 7100 to 8400, depending on the interlayer separation. After equilibration, this yielded simulation boxes with z-dimensions of about 7 Å. This means that for interlayer spacings of about 1-2 nm, the spacing between neighboring periodic images was 5-6 nm. A snapshot of the full periodic simulation box can be seen in
A harmonic bias force was then applied to the interlayer spacing using the COLVARS package. (Fiorin, G., et al., Molecular Physics 111, 3345-3362 (2013).) The harmonic force constants ranged from 250 kcal mol−1 Å−2 to 2500 kcal mol−1 Å−2. Both the window widths and biasing potentials were adjusted iteratively and by hand to fully sample the relevant range of interlayer separations. The system was equilibrated with the bias force for 100 ps, and then the histogram of the interlayer separation was collected for 100 ps. The histograms were reweighted to compute the PMF using the Weighted Histogram Analysis Method. (Grossfield, A. WHAM the weighted histogram analysis method, version 2.0.9; Kumar, S., et al., Journal of Computational Chemistry 13, 1011-1021 (1992); Roux, B., Computer Physics Communications 91, 275-282 (1995).) The PMFs for the acetate- and acetamide-functionalized membranes were averaged over only two instances of the random placement of the functional groups, due to computational cost.
The internal degrees of freedom of the MoS2 sheets were held rigid during the simulations. The body forces on the sheets in the x- and y-directions and the body torques on the sheets were all fixed to zero to prevent the sheets from sliding and rotating. Finally, the nonstandard barostat was the same as described above, except that the simulation was barostatted in both the x- and z-dimensions. All other simulation details were the same as in the unbiased simulations.
The MD force field was parameterized to faithfully reproduce water-water and water-MoS2 interactions, but it was not designed to accurately model MoS2—MoS2 interactions. Force fields did exist that were parameterized based on bulk properties of MoS2, but these were not well tested for applications in water-MoS2 systems. The inventors' choice was appropriate for modelling the structure and dynamics of water in MoS2 channels with fixed interlayer spacings, where the MoS2—MoS2 interactions were irrelevant. It is unclear, however, what effect this choice of force field may have on the PMFs presented in
Even with the MoS2—MoS2 van der Waals interactions entirely absent, the conclusions were qualitatively unchanged: ce-MoS2 and acetate-MoS2 both had global minima at 12.5 Å. In amide-MoS2, the two minima were only separated by about 0.5 kcal mol−1 (<1 kBT), while the experimentally observed interlayer separation was at 9.9 Å. Adding any van der Waals interactions will stabilize the 10 Å minimum relative to the 12.5 Å minimum, breaking this bi-stability and making the 10 Å minimum the globally stable one. This demonstrates that these results are qualitatively robust to the choice of MoS2 potential.
First, 300 mg of MoS2 powder (Sigma-Aldrich) was stirred with 3 ml of n-butyllithium for over 48 hours under an Ar atmosphere. The LixMoS2 product was washed five times with hexane, transferred out of the Ar atmosphere and bath sonicated in 300 ml of water for 60 minutes. The resulting solution was dialyzed until the solution had a pH of ˜5 (several days). Finally, the solution was centrifuged for 15 min at 1500 rpm to remove any unexfoliated material. The final concentration of suspended flakes was on average −0.3 mg/ml.
Functionalization of the ce-MoS2 was conducted in the liquid phase. A 20×molar excess of 2-iodoacetamide (Fisher Chemicals) or iodoacetic acid (Sigma-Aldrich) was stirred with ce-MoS2 for five days. The solution was then washed 5 times with water via centrifugation and re-suspension, then centrifuged at 1500 rpm to remove aggregates and was sonicated at low power for about 20 minutes.
Membranes were fabricated by vacuum filtering a known quantity of MoS2 solution through a mixed cellulose ester substrate with 25 nm pores (Millipore-Sigma), cut to shape and dried for a defined period before testing.
The word “illustrative” is used herein to mean serving as an example, instance, or illustration. Any aspect or design described herein as “illustrative” is not necessarily to be construed as preferred or advantageous over other aspects or designs. Further, for the purposes of this disclosure and unless otherwise specified, “a” or “an” can be mean only one or can mean “one or more.” Both embodiments are covered.
The foregoing description of illustrative embodiments of the invention has been presented for purposes of illustration and of description. It is not intended to be exhaustive or to limit the invention to the precise form disclosed, and modifications and variations are possible in light of the above teachings or may be acquired from practice of the invention. The embodiments were chosen and described in order to explain the principles of the invention and as practical applications of the invention to enable one skilled in the art to utilize the invention in various embodiments and with various modifications as suited to the particular use contemplated. It is intended that the scope of the invention be defined by the claims appended hereto and their equivalents.
The present application claims priority to U.S. provisional patent application No. 63/050,566 that was filed Jul. 10, 2020, the entire contents of which are incorporated herein by reference.
This invention was made with government support under 5J-30161-0064A awarded by the Department of Energy. The government has certain rights in the invention.
Filing Document | Filing Date | Country | Kind |
---|---|---|---|
PCT/US21/40606 | 7/7/2021 | WO |
Number | Date | Country | |
---|---|---|---|
63050566 | Jul 2020 | US |