CATALYSTS

Information

  • Patent Application
  • 20240100509
  • Publication Number
    20240100509
  • Date Filed
    September 25, 2023
    7 months ago
  • Date Published
    March 28, 2024
    a month ago
Abstract
The invention provides a catalyst comprising 1) Ru0 and 2) Pd0, Pt0, Rh0, or Ir0, on a solid support. The catalyst is useful for water purification applications. In particular, the catalyst is useful for reducing ClO3− and ClO2− from water. The invention also provides catalysts that are useful for reducing nitrate in a water supply and methods for their use.
Description
PRIORITY

This application claims priority to U.S. Provisional Patent Application No. 63/410,083, filed 26 Sep. 2022. The entire content of U.S. Provisional Patent Application No. 63/410,083 is hereby incorporated herein by reference.


BACKGROUND

More than four million tons of sodium chlorate (NaClO3) are manufactured annually worldwide for pulp bleaching, weed control, and pyrotechnics, et al., (researchandmarkets.com/reports-/5323457 (2020)). Water disinfection using hypochlorite or chlorine dioxide and various electrochemical processes (e.g., chloralkali, water splitting, seawater valorization, and wastewater treatment) also generate ClO3 as a byproduct (Gorzalski, A. S. & Spiesman, A. L. J. Am. Water Works Assoc. 107, E613-E626 (2015); Karlsson, R. K. & Cornell, A. Chem. Rev. 116, 2982-3028 (2016); Lakshmanan, S. & Murugesan, T., Clean Technol. Environ. Policy 16, 225-234 (2014); Park, H., et al., J. Phys. Chem. C 113, 370, 7935-7945 (2009); Kumar, A., et al., Nat. Catal. 373, 2, 106 (2019); Cho, K. et al., Environ. Sci. 376, Technol. 48, 2377-2384 (2014); Ibl, N. & Landolt, D. J., Electrochem. Soc. 115, 713-720 (1968); and Landolt, D. & Ibl, N., Electrochim. Acta 15, 1165-1183 (1970)). Not surprisingly, ClO3 enters the water environment, dairy supply chain, and agricultural products.


When ingested, ClO3 can cause red blood cell rupture and thyroid gland malfunction (Mastrocicco, M., et al., Environ. Pollut. 231, 383, 1453-1462 (2017); Rosemarin, A., et al., Environ. Pollut. 85, 3-13 (1994); Stauber, J. L. Aquat. Toxicol. 41, 213-227 387 (1998); McCarthy, W. P. et al. Compr. Rev. Food Sci. Food Saf. 17, 1561-1575 (2018); Kettlitz, B. et al. Food Addit. Contam. Part A 33, 392, 968-982 (2016); Li, M. et al. Environ. Int. 158, 106939 (2022); Liu, Q., et al., Food Addit. Contam. 397, Part A 38, 2045-2054 (2021); EFSA CONTAM Panel. EFSA J. 13, 4135 (2015); Smith, D. J. & Taylor, J. B. J. Agric. Food Chem. 59, 1598-1606 (2011); and Panseri, S. et al. Food Chem. 330, 127205 (2020)). The World Health Organization, European Union, and China have set the limit of ClO3 concentration in drinking water at 0.7 mg L−1. The United States has set the health reference level at 0.21 mg L−1 and the minimum reporting level at 0.02 mg L−1. Hence, a highly efficient ClO3 reduction method would be of significant value.


Although the ClO3 challenge for water systems has been recently recognized (Gorzalski, A. S. & Spiesman, A. L. J. Am. Water Works Assoc. 107, E613-E626 (2015); and Alfredo, K., et al., J. Am. Water Works Assoc. 107, 187 (2015)), research efforts for ClO3 reduction are limited. Platinum group metal (PGM) catalyzed hydrogenation provides a clean degradation route:





ClO3+3H2→Cl+3H2O.


Besides, the ubiquitous use of PGM in automotive catalytic converters (Saguru, C., Ndlovu, S. & Moropeng Hydrometallurgy 182, 44-56 (2018)) and the negligible PGM leaching under the H2 atmosphere (De Corte, S., et al., Microb. Biotechnol. 5, 5-17 (2012)) rationalize the application of PGM for water treatment (Gao, J. et al., ACS ES&T Eng. 1, 562-570 (2021); Chaplin, B. P. et al., Environ. Sci. Technol. 46, 3655-3670 (2012); and Ren, C., et al., ACS ES&T Eng., 2, 181-188 (2022)). Ren, C., et al., ACS ES&T Eng., 2, 181-188 (2022)). However, most reported ClO3 reduction catalysts (e.g., Rh (Van Santen, et al., U.S. Pat. No. 6,270,682 (2001)), Ir (Kuznetsova, L. I. et al., Appl. Catal. A: Gen. 427, 8-15 (2012)), Pd (Ye, T. et al., ACS Applied Nano Materials 1, 6580-6586 (2018)), Mo—Pd (Ren, C. et al., ACS Catal. 10, 8201-8211 (2020)) exhibit maximum activity in acidic conditions. The proton-assisted mechanisms severely restrict the catalytic performance around neutral pH. If acidification is not feasible, a 10-80× dose of PGM catalyst is necessary to compensate the activity loss and maintain the same reaction rate as at pH≤4 (FIG. 1).


Currently there is a need for catalysts that are useful for water purification applications. There is also a need for facile and commercially feasible methods for preparing such catalysts. In particular, there is a need for catalysts that can be used for water purification applications at neutral or alkaline pH.


SUMMARY

To address this major challenge and advance reductive catalysis for water treatment, the invention (i) achieves the unprecedented high activity of ClO3 reduction at pH 7 by harnessing the unique functions of Ru and Pd, Pt, Rh, or Ir, (ii) develops a rapid and convenient preparation method for Ru0 catalyst with metal contents as low as 0.1 wt %, (iii) elucidates the structure and synergy of the metals, and (iv) showcases the catalyst robustness under practical and challenging scenarios.


The invention provides a facile method to prepare functional Ru catalysts on a porous support. The catalysts are useful to reduce aqueous chlorate for water and waste treatment at or near neutral pH, and to reduce aqueous chlorate byproduct in chloralkali brine.


In one aspect, the invention provides a catalyst comprising 1) Ru0 and 2) Pd0, Pt0, Rh0, or Ir0, on a solid support.


In another aspect the present invention provides a method comprising contacting a water sample that comprises ClO3 with a catalyst of the invention under conditions such that at least some of the ClO3 is reduced to Cl. In one embodiment, at least 50% of the ClO3 is reduced to Cl. In another embodiment, at least 70% of the ClO3 is reduced to Cl. In another embodiment, at least 80% of the ClO3 is reduced to Cl. In another embodiment, at least 90% of the ClO3 is reduced to Cl. In another embodiment, at least 90% of the ClO3 is reduced to Cl. In another embodiment, at least 95% of the ClO3 is reduced to Cl. In another embodiment, at least 99% of the ClO3 is reduced to Cl.


In another aspect the present invention provides a method comprising sequentially reducing 1) PdII, PtII, RhIII, or IRIII and 2) RuIII on a support to provide a catalyst.


In another aspect the present invention provides a method comprising sequentially reducing in situ, PdII, PtII, RhIII, or IRIII and RuIII on a support to provide an Ru—Pd/C catalyst.


In another aspect the present invention provides a method comprising sequentially reducing PdII and RuIII in situ on a support to provide a catalyst.


In another aspect the present invention provides a method comprising reducing PdII to Pd0 on a support that comprises carbon and reducing RuIII to Ru0 on the support to provide a Ru—Pd/C catalyst.


In another aspect the present invention provides a catalyst prepared according to a method of the invention.


In another aspect the present invention provides a method for preparing a catalyst as described herein.


In another aspect the present invention provides a method comprising contacting a water sample that comprises NO3 with a catalyst comprising 1) Ru0 and 2) Pd0, Pt0, Rh0, or Ir0, on a solid support, under conditions such that at least some of the NO3 is reduced to NH4+.


In another aspect the present invention provides a method comprising contacting a water sample that comprises NO3 with a catalyst that comprises in situ reduced Ru particles and another metal, under conditions such that at least some of the NO3 is reduced to NH4+. In one embodiment, the catalyst comprises Os—Ru/C, Ru—Ru/C, Ru—Rh/C, Ru—Ir/C, Ru—Pt/C or Ru—Pd/C.


In another aspect the present invention provides a method comprising contacting a water sample that comprises NO3 with a catalyst that comprises Ru—Pd/C or In—Pd/C, under conditions such that at least some of the NO3 is reduced to NH4+.


In another aspect the present invention provides processes and intermediates disclosed herein.





BRIEF DESCRIPTION OF THE DRAWINGS


FIG. 1 shows performance data for ClO3 reduction by a Ru—Pd/C catalyst of the invention and other catalysts.



FIGS. 2A-2I: The reduction of (A) aqueous PdII (360 mg L−1 Pd from Na2 PdCl4), (B) aqueous RuIII (40 mg L−1 Ru from RuCl3), and (C) first PdII (40 mg L−1) then RuIII (40 mg L−1) by 1 atm H2 (blown from the needle tip 5 mm above the liquid) at 20° C. (D) The all-in situ procedure for Ru—Pd/C preparation. Profiles of ClO3 reduction by (E) Ru—Pd/C and (F) Ru/C after different durations of RuIII reduction before adding ClO3. Reaction conditions: 1 mM C ClO3, 0.1 g L−1 of 1 wt % Ru-1 wt % Pd/C or 1 wt % Ru/C, pH 7, 1 atm H2, 20° C. XPS spectra (empty dots) and fits (solid lines) of Pd 3d and Ru 3p of (G+H) Ru—Pd/C and (I) Ru/C.



FIGS. 3A-3B show the ratio of dissolved Ru during the preparation of Ru/C (no Pd) with the all-in situ method: (A) adsorption followed by (B) H2 sparging. Data in panel B started from the 4.9% Ru remained in solution after the 1 h adsorption shown in panel A. Reaction conditions: 0.1 g L−1 carbon support, 1 mg L−1 Ru from RuCl3 (to prepare 1 wt % Ru/C), 1 atm H2, 20° C.



FIG. 4 shows the catalytic reduction of 1 mM ClO3 by three 5 wt % Ru/C catalysts. The two catalysts prepared in the lab used the same carbon support. The 5 wt % content of Ru was chosen to compare with the commercially available 5 wt % Ru/C catalyst and thus validate the all-in situ method. Reaction conditions: 0.1 g L−1 catalyst, pH 7, 1 atm H2, 20° C.



FIGS. 5A-5D: (A) Profiles and TOF0 for 1 mM ClO reduction by three catalysts prepared by the all-in situ method. (B) Chlorine balance during ClO3 reduction. (C) Reduction of 100 mM ClO3 by Ru—Pd/C and Ru/C. (D) pH dependence of Ru—Pd/C, Ru/C, and 5 wt % Rh/C and 5 wt % Mo-5 wt % Pd/C. First order rate constants were normalized by the loading of PGM. Reaction conditions 0.1 gL−1 catalyst (1 wt % Ru and/or Pd), pH 7, 1 atm H2, 20° C.



FIGS. 6A-6N: HAADF-STEM imaging of (A, B, G, and H) 1 wt % Ru-1 wt % Pd/C, (J) 1 wt % Pd/C, and (K) 1 wt % Ru/C prepared by the all-in situ method. (C, D, E, F and I) EDS elemental mapping. (L) EDS mapping of 1 wt % Pd-1 wt % Ru/C. The dotted areas in D, E, F and I highlight the Ru coverage outside Pd., Conceptual illustration of Ru and Pd particle formation in (M) Ru—Pd/C and (N) Ru/C.



FIG. 7 illustrates the catalytic reduction of 1 mM ClO3 by 1 wt % Ru-1 wt % Pd/C and 1 wt % Ru-1 wt % Rh/C. Reaction conditions: 0.1 g L−1 catalyst, pH 7, 1 atm of H2, and 20° C.



FIG. 8: Catalytic reduction of 1 mM ClO3 by Ru—Pd/C and Pd—Ru/C catalysts (both contain 1 wt % Pd and 1 wt % Ru). Reaction conditions: 0.1 g L−1 catalyst, pH 7, 1 atm H2, and 20° C. The filled data points (C/C0>0.2) were used for the zeroth-order kinetic).



FIGS. 9A-9D: Profiles of (A) ClO reduction by individual Ru/C, Pd/C, and 1:1 mixed Ru/C+Pd/C at pH 7; (B) ClO2 reduction at pH 7.9 by Ru/Al2O3 and Pd/Al2O3; (C) ClO2 and ClO3 reduction in the mixture by Ru/Al2O3 at pH 7.9. Default reaction conditions: 0.1 g L−1 of individual catalyst containing 1 wt % of metal, 1 mM ClO3 or ClO2, 1 atm H2, 20° C. (D) Illustrated reaction mechanisms on the Ru—Pd/C catalyst surface.



FIGS. 10A-10D: Profiles of ClO3 reduction (A) in the presence of Cl and SO42−, (B) in tap water (pH 7.9) and DI water with and without 0.4 mM NO3− (pH 8.0) at 0.5 gL−1 catalyst loading, (C) in the first and fifth ClO3− spike (1 mM) in the same batch reactor, and (D) in the as prepared catalyst suspension and by the centrifuge-collected and redispersed catalyst. Default reaction conditions: 0.1 g L−1 of wt % Ru-1 wt % Pd/C, 1 mM ClO3, pH 7, 1 atm H2, 20° C.



FIG. 11: Catalytic reduction of 1 mM ClO3 by 1 wt % Ru-1 wt % Pd/C in the presence of 2 M NaCl. Reaction conditions: 0.1 g L−1 catalyst, pH 7, 1 atm H2, and 20° C. FIG. 10A shows the first 1 hour of the data set.



FIGS. 12A-12B: Catalytic reduction of (A) 0.4 mM NO3 and (B) 1 mM ClO3 by 1 wt % Ru-1 wt % Pd/C in the tap water and DI water. The first 15 minutes time windows are marked to highlight the slow NO3 reduction and relatively fast ClO3 reduction. Reaction conditions: 0.5 g L−1 catalyst, pH 8, 1 atm H2, and 20° C.



FIG. 13: Shows data from Example 5 for representative catalysts Ru—Pd/C, Ru—Rh/C, Ru—Pt/C, and Ru—Ir/C.



FIG. 14. Shows data from Example 6. NO3− Reduction Activity of M-Pd/C Catalysts Containing Group V-VIII Metals (M). See reaction conditions in Table 1.



FIG. 15. Shows data from Example 6. NO3− Reduction Activity of M-Ru/C Catalysts Containing Group V-VIII Metals (M). See reaction conditions in Table 2.



FIG. 16. Shows data from Example 6. NO3− Reduction Activity of Ru-M0/C Catalysts Containing Variable Hydrogenation Metal Nanoparticles (M0). See reaction conditions in Table 3.



FIG. 17. Shows data from Example 6. 1 mM ClO3− reduction by 0.1 g L−1 1 wt % Ru-1 wt % Pd/C and 0.2 g L−1 5 wt % Ru-5 wt % Ru/C. Reaction conditions: pH 3, 1 atm H2, 20° C.



FIG. 18. Shows data from Example 6. 1 mM NO3 reduction by Ru—Pd/C. Reaction conditions: pH 3, 1 atm H2, 20° C.



FIG. 19. Shows data from Example 6. 1 mM NO3− reduction by 0.1 g L−1 Ru—Pd/C. Reaction conditions: pH 3, 1 atm H2, 20° C.



FIG. 20. Shows data from Example 6. Salt Effect on NO3− Reduction Activity with M-Pd/C Catalysts. Reaction conditions: catalyst (5 wt % Pd and 5 wt % M, 0.5 g L−1 for In and Re—Pd/C, 2 g L−1 for other catalysts); NaNO3 (1 mM); pH 3 (5 mM H3PO4/NaH2PO4 buffer, adjusted by HCl in deionized water); H2 (1 atm); 25° C.





DETAILED DESCRIPTION OF THE INVENTION

The solid support can be any material that is suitable for providing a catalyst that can be used to reduce ClO3 to Cl. For example, in one embodiment, the solid support can comprise porous carbon, alumina, silica, a zeolite, a clay, TiO2, CeO2, ZrO2, Mxene (Nanoscale Horiz., 2020, 5, 235-258), a carbon nanotube, graphene, or biochar. In one embodiment, the solid support comprises carbon. In one embodiment, the solid support is carbon.


In one embodiment, the catalyst comprises highly dispersed Ru0. The term highly dispersed Ru0 includes dispersions from a single atom to 20 nm. In one embodiment, the term includes dispersions from a single atom to about 5 nm. In one embodiment, the term includes dispersions from a single atom to about 3 nm. In one embodiment, the term includes dispersions from about 3 nm to about 5 nm.


As used herein, the term “turnover frequency” means the number of ClO3 that are reduced to Cl:per minute:per Ru atom on the surface of the catalyst. In one embodiment, the turnover frequency of the catalyst is at least about 10. In one embodiment, the turnover frequency of the catalyst is at least about 11. In one embodiment, the turnover frequency of the catalyst is at least about 12. In one embodiment, the turnover frequency of the catalyst is at least about 13. In one embodiment, the turnover frequency of the catalyst is less than about 35. In one embodiment, the turnover frequency of the catalyst is less than about 30. In one embodiment, the turnover frequency of the catalyst is less than about 25. In one embodiment, the turnover frequency of the catalyst is in the range of from about 0.2 to about 30. In one embodiment, the turnover frequency of the catalyst is in the range of from about 5 to about 30. In one embodiment, the turnover frequency of the catalyst is in the range of from about 5 to about 20. In one embodiment, the turnover frequency of the catalyst is in the range of from about 5 to about 15. In one embodiment, the turnover frequency of the catalyst is in the range of from about 10 to about 15.


Specific values identified below are for illustration only; they do not exclude other defined values or other values within defined ranges. It is to be understood that two or more specific values may be combined. It is also to be understood that the specific values listed herein below (or subsets thereof) can be excluded from the invention.


Specifically, the water sample can also comprise ClO2 and/or ClO4.


Specifically, the method can be carried out at a pH in the range from about 3 to about 11.


Specifically, the method can be carried out at a pH in the range from about 3 to about 8.


Specifically, the method can be carried out at a pH in the range from about 5 to about 8.


Specifically, the method can be carried out at a pH of at least about 6.


Specifically, the method can be carried out at a pH of at least about 7.


Specifically, the sample can further comprise sulfate, phosphate, silicate, chloride, bromide, iodide, nitrate, nitrite, perchlorate, fluoride, or bromate.


Specifically, the sample can further comprise sulfate.


Specifically, the catalyst has a turnover number of at least 9,000 for ClO3.


Specifically, the catalyst has a turnover number of at least 10,000 for ClO3.


Specifically, the catalyst has a turnover number of at least 11,000 for ClO3.


Specifically, the catalyst has a turnover number of at least 15,000 for ClO3.


Specifically, the catalyst has a turnover number of at least 20,000 for ClO3.


Specifically, the Ru0 and the Pd0, Pt0, Rh0, or Ir0 together provide synergy.


Specifically, the reducing is carried out at a temperature in the range from about 4° C. to about 95° C.


Specifically, the reducing is carried out at a temperature in the range from about 10° C. to about 80° C.


Specifically, the reducing is carried out at a temperature in the range from about 10° C. to about 45° C.


Specifically, the reducing is carried out at a temperature in the range from about 15° C. to about 25° C.


Specifically, the reducing is carried out at a temperature of about 20° C.


Specifically, the reducing is carried out under 1-3 atm of H2. In one embodiment, the reaction can be carried out in a mixture of an inert gas (e.g., N2 or Ar) and H2. In one embodiment, the ratio of inert gas to H2 is about 90:10. In one embodiment, the ratio of inert gas to H2 is about 80:20. In one embodiment, the ratio of inert gas to H2 is about 70:30. In one embodiment, the ratio of inert gas to H2 is about 60:40. In one embodiment, the ratio of inert gas to H2 is about 50:50. In one embodiment, the ratio of inert gas to H2 is about 40:60. In one embodiment, the ratio of inert gas to H2 is about 30:70. In one embodiment, the ratio of inert gas to H2 is about 20:80. In one embodiment, the ratio of inert gas to H2 is about 10:90. In one embodiment, the ratio of inert gas to H2 is about 5:95.


Specifically, the reducing is carried out under about 1 atm of H2.


In one embodiment, the term “about” means ±20%.


The invention will now be illustrated by the following non-limiting Examples.


EXAMPLES
Example 1. Catalyst Preparation

Both aqueous RuIII (from RuCl3·xH2O) and PdII (from Na2PdCl4) can be reduced into Ru0 and Pd0 precipitates, respectively, by direct exposure to 1 atm H2 in the headspace at 20° C., but the reduction of RuIII is much slower than that of PdII. While the yellow PdII solution was fully converted into Pd black (i.e., large Pd0 solids) and colorless liquid within 35 min (FIG. 2A) (Gao, J. et al., ACS ES&T Eng. 1, 562-570 (2021), the same procedure could not complete the reduction of RuIII by 12 h (FIG. 2B). However, RuIII added to the Pd black suspension underwent fast color fading and particle formation within 2.5 h (FIG. 2C). Hence, Pd0 accelerated the H2 reduction of RuIII into Ru0. The adsorption behaviors of aqueous RuIII and PdII on porous carbon support (1300 m2 g−1) were also different. While >98% of PdII was immobilized on carbon within 5 minutes, only 84% of RuIII was immobilized in the same time frame.


A bimetallic catalyst was prepared by adding RuIII into an all-in situ prepared Pd0/C (FIG. 2D). In the nanoscale carbon pores, the highly dispersed Pd0 particles were expected to provide a faster reduction of RuIII than the bulky Pd black aggregates. The adsorption and reduction of RuIII (onto Pd0/C by 1 atm H2 at 20° C.) for only 10 minutes yielded the full activity of ClO3 reduction (FIG. 2E). The dissolved Pd and Ru in the aqueous phase were below 1 μg L−1, showing that >99.9% of the two PGMs were immobilized. X-ray photoelectron spectroscopy (XPS) measured the binding energy (BE) of Pd 3d5/2 at 335.8 eV and Ru 3p3/2 at 461.8 eV (FIG. 2G and FIG. 2H), confirming the reduction of both metals to the metallic state. Therefore, only 20 minutes is needed to prepare Ru0—Pd0/C all-in situ from aqueous RuIII, PdII, carbon support, and 1 atm H2 at 20° C.


RuIII was also directly immobilized on the same carbon support without Pd0. The adsorption of >95% RuIII required 1 hour (FIG. 3A). Upon H2 exposure for another 1 hour, the fraction of dissolved Ru was further lowered to 0.2% (FIG. 3B). XPS characterization confirmed the yield of Ru0/C (FIG. 2I, Ru 3p3/2 BE at 461.8 eV, the same as in Ru0—Pd0/C). More than 1 hour of H2 exposure was required to maximize the ClO3 reduction activity (FIG. 2F). The three catalysts had consistent metal contents −0.64 wt % of Pd in Pd0/C, 0.68 wt % of Pd+0.98 wt % of Ru in Ru0—Pd0/C, and 0.94 wt % of Ru in Ru0/C. These values show the reliability of the all-in situ preparation method, which provides a fair basis for activity comparison. The all-in situ adsorption-reduction method for Ru was further validated by comparing it with the conventional incipient wetness+heated H2 reduction method and a commercial catalyst (FIG. 4).


Example 2. Chlorate Reduction Performance

Based on the chemisorption data (Table 1) and ClO3 reduction time profile (FIG. 5A), the initial turnover frequency (TOF0, the average number of ClO3 anions reduced by individual surface metal atoms upon the first sampling time) was calculated. At pH 7, Pd/C barely catalyzed ClO3 reduction (TOF0=0.4 min−1 on Pd0) while Ru/C was much more active (TOF0=9.0 min−1 on Ru0). Surprisingly, the Ru—Pd/C was substantially more active than Ru/C. A conservative estimation of TOF0 (see later sections for the estimation of Ru dispersion) is 13.9 min−1 on Ru0. A good mass balance of Cl was established between ClO3 and Cl (FIG. 5B), indicating no accumulation of other chlorine species.









TABLE 1







Metal Dispersion from CO Chemisorption.ª











Metal


Entry
Catalyst
dispersion





1
Pd/C
20.9%


2
Ru/C
15.0%


3
Ru/Cb
37.1%


4
Ru—Pd/C
36.5-42.5%c


5
Pd—Ru/C
25.0-29.2%c






aUnless specified, the catalysts were prepared by the all-in situ method with a nominal 1 wt % content for each metal. The stoichiometries for Ru:CO and Pd:CO are 12:740 and 2:1,41 respectively.




bPrepared by conventional method involving incipient wetness impregnation and reduction with heated H2 (see Method section for details).




cThe lower and higher limits were calculated assuming all-Ru and all-Pd scenarios in the bimetallic system.







Ru and Pd contents from 0.1 to 5 wt % were extensively screened to identify the roles of both metals. First, the addition of Ru as low as 0.1 wt % can significantly enhance the activity of monometallic Pd/C (Table 2, entry 4 versus entry 5) and vice versa (entry 6 versus entry 7), suggesting the synergy between Ru and Pd. The highest activity was shown when the two metals were both at 1 wt %. An unexpected advantage of Ru—Pd/C over Ru/C was observed from the treatment of concentrated ClO3.


The use of 0.1 g/L Ru—Pd/C achieved 99.9% reduction of 100 mM ClO3 (FIG. 2C) with a turnover number (TON, the total number of ClO3 anions reduced by each surface Ru atom) of 11,970. In contrast, Ru/C was substantially inhibited by the concentrated ClO3 (FIG. 2C versus A). Second, increasing the metal contents above 1 wt % did not proportionally accelerate ClO3 reduction (Table 2, entries 1-3). Instead, the rate constant normalized by the total mass of Ru and Pd became lower, probably due to the decreased metal dispersion. Third, decreasing the metal content below 1 wt % did not further increase the normalized rate (Table 2, entries 3, 8, and 9). In other words, in order to support the same amount of metal to achieve the same reaction rate, more carbon material needs to be used. Thus, a 1 wt % formulation was used for both Ru and Pd.









TABLE 2







Rate Constants of ClO3 Reduction by RU—Pd/C


Catalysts with Variable Formulations.
















Apparent
Metal-normalized






rate
rate




Pd
Ru
(mM
(mmol gmetal−1



Entry
(wt %)
(wt %)
min−1)a
min−1)b

















1
5
5
0.116
11.6



2
3
3
0.114
19.1



3
1
1
0.091
45.3



4
1
0.1
0.010
9.2



5
1
0
0.001
0.8



6
0.1
1
0.024
21.5



7
0
1
0.014
14.3



8
0.5
0.5
0.040
40.1



9
0.1
0.1
0.009
44.0








aReaction conditions: 0.1 g L−1 of Ru—Pd/C, 1 mM ClO3 , pH 7, 1 atm H2, 20° C.





bNormalized to the total mass of Ru and Pd.







The Ru—Pd/C outperforms all reported PGM-based catalysts for ClO3 reduction in a wide pH range from 3 to 8. At pH 7, the metal-normalized first-order rate constant is more than two orders of magnitude higher than those of Rh/C (Chen, X. et al. Chem. Eng. J. 313, 745-752 (2017)) and Mo—Pd/C (Ren, C. et al., ACS Catal. 10, 8201-8211 (2020)) (FIG. 5D), both of which show the highest activity at pH 3. However, the activity of Ru—Pd/C also increased at pH 3, still being 5-7 times more active than Rh/C and Mo—Pd/C under the acidic condition. A comprehensive comparison of all PGM catalysts is provided in FIG. 1. Although Rh/C is much more active than Pd/C for ClO3 reduction at pH 7 (FIG. 1, entry 4 versus 3), the Ru—Rh/C catalyst prepared by the same all-in situ method showed almost identical activity as Ru—Pd/C (FIG. 7), corroborating the dominant role of Ru in the reaction with ClO3. Thus, the use of Ru avoids the use of Rh and saves a substantial amount of Pd. Rh is 28 times more expensive than Ru (platinum.matthey.com), while Pd is 3 times more expensive but 3-4 orders of magnitude less active than Ru. However, in comparison to the monometallic Ru/C, the significantly enhanced activity, robustness with concentrated ClO3, and different pH dependence of Ru—Pd/C also suggest important roles of Pd in catalyst structure and reaction scheme.


Example 3. Mechanistic Elucidation

High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) characterization of Ru—Pd/C observed fine metal particles with an average size of 2.3 nm on the carbon support (FIGS. 6A, 6B, 6G, and 6H). The size is apparently larger than Pd particles of Pd/C (<2 nm on average; Gao, J. et al., ACS ES&T Eng. 1, 562-570 (2021)) before accommodating Ru (FIG. 6J versus 6A). Energy-dispersive X-ray spectroscopy (EDS) elemental mapping (FIGS. 6C-6F) confirmed the bright spots as overlapped Pd and Ru. The relatively large particles clearly show the mixing pattern as a layer of Ru on the outside of Pd particles (FIGS. 6F and 6I). Hence, RuIII was rapidly reduced as Ru0 on Pd0 particles by the on-surface active H (FIG. 6M; Conrad, H., Ertl, G. & Latta, E., Surf Sci. 41, 435-446 (1974)). Besides, one cannot exclude the scenario that a minor fraction of Ru was reductively immobilized in the vicinity of Pd0 particles by spilled-over H on the carbon support (Prins, R. Chem. Rev. 112, 2714-2738 (2012)).


In stark contrast, the direct reduction of RuIII on carbon resulted in large aggregates (FIG. 6K). In comparison to PdII, the poorer adsorption of RuIII and much slower reduction (by H2 at 20° C.) could be responsible for such morphology (FIG. 6N). Relatively few Ru0 seeds could form at the beginning, and the remaining RuIII could migrate to the surface or vicinity of existing Ru0 particles, slowly react with the active H, and gradually yield the bulky Ru0 solid. The integration of Pd0 with Ru0/C was also achieved. Elemental mapping of Pd—Ru/C found the overlapping of Pd and Ru in much larger aggregates than in Ru—Pd/C (FIG. 6L versus 6F, note the scale). Not surprisingly, the ClO3 reduction activity of Pd—Ru/C was only 43% of that of Ru—Pd/C (FIG. 8).


Chemisorption data indicate substantially enhanced metal dispersion in the bimetallic catalysts (FIG. 1). While the Pd dispersion in Pd/C was 20.9%, the total metal dispersion in Ru—Pd/C ranged from 36.5% (all Ru) to 42.5% (all Pd), based on the stoichiometry for Ru:CO (12:7) and Pd:CO (2:1). Without a high-temperature process, the later added Ru is less likely to change the morphology of Pd. Thus, the enhanced dispersion in Ru—Pd/C should be attributed to high dispersion of Ru ranging from 55% (if Pd and Ru did not overlap at all) to 85% (if Pd was completely covered by Ru). This value is higher than Ru/C prepared by the conventional method (37.1%), which accelerated metal reduction by heated H2. Similarly, the total metal dispersion in Pd—Ru/C (25.0-29.2%) is also higher than Ru/C (15.0%). In comparison to Pd/C (20.9% dispersion), Ru enhanced the Pd dispersion in Pd—Ru/C to the calculated 35-58%.


The >55% dispersion of Ru in Ru—Pd/C could contribute to the higher ClO3 reduction activity than Pd—Ru/C (dispersion of Ru<15% due to Pd coverage). However, the higher activity of Pd—Ru/C (k=2.3 mM h−1) than Ru/C (k=0.9 mM h−1) suggests other critical roles of Pd. Despite the negligible activity of Pd/C at pH 7, a 1:1 mixture of the two monometallic Pd/C and Ru/C catalysts exhibited a higher activity than using Ru/C only (FIG. 9A). Thus, the individual activity of Ru and Pd with ClO2 was investigated, since ClO2 is the most probable first intermediate from ClO reduction. Since ClO2 can react with carbon (Voudrias, E. et al. Water Res. 17, 1107-1114 (1983); and Gonce, N. & Voudrias, E. A., Water Res. 28, 1059-1069 (1994), Pd/Al2O3 and Ru/Al2O3 were prepared with the same all-in situ method (Gao, J. et al., ACS ES&T Eng. 1, 562-570 (2021)). Pd/Al2O3 showed 5.4-fold higher activity of ClO reduction than Ru/Al2O3 (FIG. 9B). More interestingly, when 1:1 of ClO3 and ClO2 were added together to Ru/Al2O3, the reduction of ClO was largely inhibited until >90% of ClO was reduced first (FIG. 9C). Therefore, the loss of activity of Ru/C upon treating 100 mM ClO3 (FIG. 5C) is likely attributed to the accumulation of ClO shortly after the reaction. Reactions of Ru species with ClO (and the potential daughter product ClO) are complex. For example, bulk Ru0 can be oxidized by concentrated ClO toward dissolution (Howe, J. L. & Mercer, F. N., J. Am. Chem. Soc. 47, 2926-2932 (1925)); oxidized Ru species such as RuII and H+ can trigger complex decomposition of ClO into ClO, HOCl, ClO, and ClHu, Z. et al., ChemComm 48, 475, 1102-1104 (2012); and Deshwal, B. R., et al., Can. J. Chem. Eng. 82, 619-623 (2004)). Details of these reactions remain largely unexplored. Nevertheless, the findings above clearly suggest the synergy between Ru0 and Pd0; Ru reduces ClO3 for a fast overall reaction while Pd rapidly scavenges ClO and generates active H to minimize the oxidative deactivation of Ru (FIG. 9D). The oxidized Ru can still be reduced back to Ru0 by Pd-activated H2 (FIGS. 2B and 6N).


Example 4. Catalyst Robustness

The performance of Ru—Pd/C for ClO3 reduction was assessed in typical application scenarios, such as (i) chloralkali NaCl brines containing the undesirable ClO3 byproduct from the anode (Van Santen, R., et al., U.S. Pat. No. 6,270,682 (2001)), (ii) waste stream from reverse osmosis or ion-exchange that enriched ClO3 from source water, and (iii) drinking water containing ClO3 from source water or disinfection operations (Gorzalski, A. S. & Spiesman, A. L. J. Am. Water Works Assoc. 107, E613-E626 (2015). Because modern water treatment usually involves sequential processes and does not expose advanced systems to raw water or known poisoning/destructive species (Ren, C., e t al , ACS ES&T Eng. 2, 181-188 (2022); and Anis, S. F., et al., Water Res. 452, 159-195 (2019)), the catalyst was not challenged with sulfide (a potent PGM catalyst poison but readily oxidizable, Tomar, M. & Abdullah, T. H., Water Res. 28, 2545-2552 (1994)) or humic acid (a common fouling species but readily adsorbable, Chaplin, B. P. et al., Environ. Sci. Technol. 46, 3655-3670 (2012)). Instead, anions such as Cl and SO42− are ubiquitous co-existing species. The presence of 1.0 M SO42−, 0.1 M Cl, and 2.0 M Cl decreased the rate of 1 mM ClO3 reduction for 30%, 58%, and 94%, respectively (FIG. 10A). Even under the inhibition by 2.0 M Cl, the reduction of 1 mM ClO3 was completed within 8 hours (FIG. 11). Thus, a higher loading of catalyst can be used to proportionally boost the apparent reaction rate (Ren, C., et al., ACS ES&T Eng., 2, 181-188 (2022); and Ren, C. et al., J. Am. Chem. Soc. 143, 7891-7896 (2021)).


Ru—Pd/C was also tested in a tap water sample from Southern California, where the groundwater occasionally contained ClO3 slightly higher than the minimum reporting level (0.02 mg L−1, or 0.24 μM). The tap water had an initial pH of 7.9 and contained 0.4 mM of NO3. The use of 0.5 g L−1 Ru—Pd/C reduced the spiked 1 mM ClO3 for 99, 99.95, and >99.99% (i.e., lower than the detection limit of 0.1 μM) within 30, 45, and 60 minutes, respectively (FIG. 10B). However, the reaction rate was slower than in the deionized (DI) water. The addition of 0.4 mM NO3 in DI water resulted in a very similar level of inhibition as in the tap water. Therefore, other constituents in the tap water were not significant inhibitors of Ru—Pd/C. This catalyst showed a relatively low activity for NO3 reduction at pH 8, and NO3 reduction barely proceeded before the majority of ClO3 was reduced (FIG. 12).


Preliminary reuse tests show that Ru—Pd/C did not lose activity after five spikes of 1 mM ClO (FIG. 10C), because the inhibition by Cl in the mM concentration range is negligible. Centrifugation and handling in air did not deactivate the recycled catalyst (FIG. 10D). The PGM leaching into water was below the detection limit (10 μg L−1; i.e., <1% of the immobilized 1 wt % Ru or Pd) from these operations


Methods

Chemicals and Materials. RuCl3·xH2O (99.98%), Na2PdCl4 (≥99.99%), NaClO3 (≥99%), and NaClO2 (technical grade, 80%) were used as received from Sigma-Aldrich. The activated carbon support (Norit GSX, steam activated and acid-washed, surface area 1300 m2 g−1) from Alfa Aesar (#L11860). The alumina support was received as ⅛″ pellets from Alfa Aesar (#43855) and ground into powders before use. A commercial 5 wt % Ru/C was used as received from Alfa Aesar (#44338). Except for the tap water, all aqueous solutions were prepared with Milli-Q water (resistivity >18.2 MΩ cm).


Catalyst Preparation. All-in situ Method for Pd/C and Ru—Pd/C: A 50 mL flask was sequentially loaded with a magnetic stir bar, 5 mg of carbon powder, 50 mL of DI water, and Na2PdCl4 stock solution. The mixture was sonicated for 1 min to disperse the carbon particles and stirred at 350 rpm for 4 minutes to allow the adsorption of PdII. The flask was capped by a rubber stopper. A 16-gauge needle penetrating the stopper was connected to the H2 gas supply (2-3 mL min−1), and the needle tip was pushed under water. The other needle had the tip above the water as the gas outlet to the atmosphere. After 5 minutes of H2 sparging at 20° C., all adsorbed PdII was reduced to Pd0. The Pd0/C suspension was added with RuCl3 stock solution and sparged with H2 for another 10 minutes at 20° C. to reduce adsorbed RuIII to Ru0, yielding Ru0—Pd0/C.


All-in situ Method for Ru/C and Pd—Ru/C. The preparation followed the same procedure as detailed above. However, the direct immobilization of RuIII onto carbon took 1 hour, and the subsequent reduction by H2 took 4 hours to yield Ru0/C. The immobilization of Pd0 onto the resulted Ru0/C still took 5 minutes for PdII adsorption and 5 minutes for the reduction by H2, yielding Pd0—Ru0/C.


Conventional Method for Ru/C. The RuIII precursor was impregnated into the same carbon support material by incipient wetness. The wet paste was dried in an oven at 75° C. for 12 hours and reduced with 90/10 (v/v) N2/H2 at 450° C. for 6 hours to yield Ru0/C (Lin, B. et al. ACS Catal. 9, 1635-1644 (2019); and Lin, B., et al., ChemCatChem 5, 1941-1947 (2013)).


Catalyst Characterization. The solid catalyst was collected from the water suspension by filtration under vacuum. The filter paper with the black paste was dried at 20° C. by the airflow in a fume hood. No inert gas protection was involved for catalyst handling or transportation. The Ru and Pd content in the catalysts were measured by inductively coupled plasma-optical emission spectrometry (ICP-OES, PerkinElmer Optima 8300) after digestion at the Microanalysis Laboratory, University of Illinois at Urbana-Champaign. The oxidation state of Ru and Pd was characterized by X-ray photoelectron spectroscopy (XPS, Kratos AXIS Supra). The sp2 C 1 s peak (284.5 eV) of the carbon support was used for binding energy (BE) calibration. XPS spectra in the resolution of 0.1 eV were fitted using CasaXPS (version 2.3.19). Microscopic characterization was conducted using the scanning transmission electron microscopy (STEM, FEI Titan Themis 300) equipped with an energy dispersive X-ray spectrometer (EDS) system. The catalyst powder was resuspended and sonicated in distilled water to further reduce the size. The STEM images were acquired with a high-angle annular dark-field (HAADF) detector. Nano Measurer software package was used for the statistical analysis of average particle size in the STEM images. The specific surface areas of Ru and Pd were determined by CO pulse titration experiments on a Quantachrome Autosorb-iQ physisorption-chemisorption instrument. The calculation of metal dispersion used the surface Ru:CO and Pd:CO stoichiometry of 12:7 (Chen, Q. et al. J. Phys. Chem. C 119, 8626-8633 (2015)) and 2:1 (Hooshmand, Z., Le, D. & Rahman, T. S. Surf Sci. 655, 7-11 (2017)), respectively.


Chlorate and Chlorite Reduction. During the all-in situ catalyst preparation, the solution pH was significantly lowered from pH 6.5 (DI water with dissolved CO2) because of the hydrolysis and reduction of [PdIICl4]2− and RuIIICl3 (Gao, J. et al., ACS ES&T Eng. 1, 562-570 (2021); and Lebedeva, A. et al., Inorg. Chem. 58, 4141-4151 (2019)). Therefore, the solution pH was adjusted by NaOH to pH 7.0 before adding ClO3 or ClO2. The addition of 1 mM NaClO2 stock solution increased the pH from 7.0 to 7.9 due to the hypochlorite impurity. However, as ClO2 is sensitive to acidic conditions (Deshwal, B. R., et al., Can. J. Chem. Eng. 82, 619-623 (2004)), the pH was not adjusted back to 7.0 after the addition. The catalytic reduction of ClO3 and ClO2 started upon their spike into the catalyst suspension. The flow of 1 atm H2 was maintained at 2-3 mL min−1, and the flask reactor was placed on the benchtop (20° C.). Aliquots were collected through the H2 outlet needle with a 3 mL plastic syringe and immediately filtered through a 0.22-μm cellulose acetate membrane.


The experiment in tap water (containing 0.4 mM NO3) used a 50 mL double-neck flask. Both necks were capped with rubber stoppers. One stopper accommodated two needles as the H2 inlet and outlet/sampling port, respectively. The other stopper accommodated a Fisherbrand accumet gel-filled pencil-thin pH combination electrode to monitor the pH during the reaction. While the reduction of ClO3 and ClO2 do not consume H+, the reduction of NO3 consumes H+ and may elevate the pH Liu, J., et al., Water Res. 47, 91-101 (2013)). To maintain the solution pH at 7-8, H2SO4 (0.1 M) was added via the sampling needle when the pH reading went higher than 8.0.


Sample Analysis and Kinetic Evaluation. The concentrations of ClO3 and ClO2 were determined by ion chromatography (Dionex ICS-5000) equipped with a conductivity detector and an IonPac AS19 column. The column temperature was set at 30° C., with 20 mM KOH eluent at 1 mL min−1. The concentrations of Ru and Pd in aqueous samples were analyzed by ICP-OES (detection limit 10 μg L−1).


When all ClO3 was reduced, the turnover number (TON) was calculated as





TON=[ClO3]0×Mw/(Lcat.×Cmetal×Dmetal)


where [ClO3]0 is the initial concentration of chlorate (mol L−1), Mw is the atomic mass of Ru or Pd (g mol−1), Lcat. is the loading of catalyst powder (g L−1), Cmetal is the metal content, and Dmetal is the metal dispersion.


The initial turnover frequency (TOF0), in min−1) was calculated as





TOF0=([ClO3]0−[ClO3]tMw/(Lcat.×Cmetal×Dmetal×t)


where [ClO3]t is the concentration at the first sampling point of reaction time t (min).


Conclusions

A highly active Ru—Pd/C catalyst was conveniently prepared by sequential all-in situ adsorption-reduction of PdII and RuIII precursors on the carbon support. The preparation only takes 20 minutes using 1 atm H2 at 20° C. without heating procedures. The Pd0 nanoparticles enhance the reduction of RuIII into highly dispersed (>55%) Ru0. The resulting Ru—Pd/C catalyst shows a substantially higher activity of ClO3 reduction into Cl than any reported catalyst at both neutral and acidic pH. The catalyst allows complete reduction of ClO3 in the presence of concentrated SO42− and as well as in the tap water matrix. While Ru shows high reactivity with ClO3, Pd is more reactive with ClO2, an inhibitor of Ru. The synergy between Ru and Pd makes Ru—Pd/C superior to monometallic Ru/C, especially in reducing concentrated ClO3.


Example 5. Additional Catalysts

Additional catalysts were prepared and tested as follows.


Reaction Conditions

Catalyst including 1 wt % Ru added in 1 wt % M/C (M=Pd, Rh, Pt, and Ir). Pd/C, Rh/C, and Pt/C were prepared as described by Gao, J. et al., ACS ES&T Eng. 1, 562-570 (2021); Ir/C was purchased from Alfa-Aesar. For all catalysts, the Ru metal was immobilized in situ as described herein above. The catalyst loading was 0.1 g/L of the powder material in water at pH 7 and 20° C. The initial concentration of chlorate was 1 mM.


Conclusion

The catalysts were found to be highly effective for chlorate reduction at pH 7. Data for representative catalysts of the invention is provided in FIG. 13.


All publications, patents, and patent documents are incorporated by reference herein, as though individually incorporated by reference. The invention has been described with reference to various specific and preferred embodiments and techniques. However, it should be understood that many variations and modifications may be made while remaining within the spirit and scope of the invention.


Example 6. Nitrate Reduction

Catalysts were prepared using the same procedure as described above for Ru—Pd/C, with the exception of using different metal salts.


M-Pd/C Catalysts Containing Group 5-8 Metals (M). Bare Pd/C exhibited no activity in NO3 reduction (Table 3 and FIG. 14). The incorporation of Group 5-8 metals enabled the reaction. Notably, among the catalysts, Ru—Pd/C showed the activity.









TABLE 3







NO3 Reduction Activity of M—Pd/C Catalysts


Containing Group V-VIII Metals (M)ª












NO3 reduction
k



Catalyst
after 1 h
(L h−1 gcat−1)







Pd/C
No reaction




V—Pd/C
   3.9%
NAc



Mo—Pd/C
  68.4%
0.481 ± 0.021



W—Pd/C
No reaction
NAc



Re—Pd/C
  83.4%
0.744 ± 0.056



Os—Pd/C
  92.6%
 1.32 ± 0.06



Ru—Pd/C
>99.9%
 8.66 ± 0.65d



In—Pd/Cb
>99.9%
 2.97 ± 0.23d








aReaction conditions: catalyst (5 wt % Pd and 5 wt % M, 2 g L−1); NaNO3 (1 mM); pH 3 (5 mM H3PO4/NaH2PO4 buffer, adjusted by HCl in deionized water); H2 (1 atm); 25° C.





bPrepared with impregnation and calcination of In(NO3)3 in Pd/C.





cNo further kinetic measurement conducted due to the low reactivity shownin the 1-h screening.





dKinetics measured with a reduced catalyst loading of 0.5 g L−1.







M-Ru/C Catalysts Containing Group 5-8 Metals (M). Bare Ru/C effectively catalyzed nitrate reduction at a moderate rate (Table 4 and FIG. 15). Incorporating Group 8 (Ru and Os) metals enhanced the reaction rate, whereas the introduction of Group 6 (Mo) and 7 (Re) metals decelerated the reaction.









TABLE 4







NO3 Reduction Activity of M—Ru/C Catalysts


Containing Group V-VIII Metals (M)ª












NO3 reduction
k



Catalyst
after 1 h
(L h−1 gcat−1 )







Ru/C
65.4%
 2.29 ± 0.11



Re—Ru/C
39.3%
 1.13 ± 0.03



Mo—Ru/C
38.1%
0.917 ± 0.006



Os—Ru/C
99.8%
 8.81 ± 1.26



Ru—Ru/C
99.8%
 10.5 ± 2.2








aReaction conditions: catalyst (5 wt % Pd and 5 wt % M, 0.5 g L−1); NaNO3 (1 mM); pH 3 (5 mM H3PO4/NaH2PO4 buffer, adjusted by HCl in deionized water); H2 (1 atm); 25° C.







Ru-M0/C Catalysts Containing Variable Hydrogenation Metal Nanoparticles)(M0). Among the five hydrogenation metals (Ru, Rh, Pd, Ir and Pt), only Ru and Rh exhibited NO3 reduction activity (Table 5 and FIG. 16). Introducing Ru to all these metals substantially enhanced the reaction rate, with Ru—Ru/C displaying the highest activity.









TABLE 5







NO3 Reduction Activity of Ru—M0/C


Catalysts Containing Variable


Hydrogenation Metal Nanoparticles (M0)a












NO3 reduction
k



Catalyst
after 1 h
(L h−1 gcat−1)







Ru/C
65.4%
 2.29 ± 0.11



Ru—Ru/C
99.8%
 10.5 ± 2.2



Rh/C
23.8%
0.426 ± 0.046



Ru—Rh/C
99.5%
 7.42 ± 0.46



Ir/Cb
No reaction




Ru—Ir/Cb
98.4%
 4.97 ± 0.44



Pt/C
No reaction




Ru—Pt/C
94.4%
 5.76 ± 0.46



Pd/C
No reaction




Ru—Pd/C
98.0%
 8.66 ± 0.65








aReaction conditions: catalyst (5 wt % M0 and 5 wt % Ru, 0.5 g L−1); NaNO3 (1 mM); pH 3 (5 mM H3PO4/NaH2PO4 buffer, adjusted by HCl in deionized water); H2 (1 atm); 25° C.





b1 wt % Ir0 in catalyst.







Optimization of the Ru—Pd/C Catalyst.

Among the screened catalysts, Ru—Ru/C and Ru—Pd/C exhibited the highest activity in NO3 reduction (Table 3-5). However, Ru—Ru/C showed limited activity in ClO3 reduction (FIG. 17).


The NO3 reduction with Ru—Pd/C catalyst is even faster without phosphate buffer (FIG. 18). Using 0.1 g L−1 of catalyst without buffer yields a reaction rate equivalent to that achieved by using 0.5 g L−1 of catalyst with buffer (FIG. 18). However, the solution pH increased from 3 to 4 without buffer.


By keeping Pd content constant and reducing Ru content from 5% to 1%, the reaction rate only reduced by 26%. (FIG. 19).


Salt Resistance of the Re-Based Catalysts.

While Ru—Pd/C demonstrated superior activity in NO3 reduction, its performance is substantially hindered in concentrated salt solutions (FIG. 20). In stark contrast, the Re—Pd/C catalyst showed even higher activity in concentrated salt.

Claims
  • 1. A catalyst comprising 1) Ru0 and 2) Pd0, Pt0, Rh0, or Ir0, on a solid support.
  • 2. The catalyst of claim 1, which comprises Ru0 and Pd0 on a solid support.
  • 3. The catalyst of claim 1, wherein the solid support comprises porous carbon, alumina, silica, a zeolite, a clay, TiO2, CeO2, ZrO2, Mxene, a carbon nanotube, graphene, or biochar.
  • 4. The catalyst of claim 1, which comprises highly dispersed Ru0.
  • 5. The catalyst of claim 1, which can reduce ClO3− to Cl− in water at a pH in the range from about 5 to about 8.
  • 6. The catalyst of claim 1, which can reduce ClO3− to Cl− in water at a pH of at least about 6.
  • 7. The catalyst of claim 1, wherein the catalyst has a turnover number of at least 9,000 for ClO3−.
  • 8. The catalyst of claim 1, which has a turnover frequency of at least 10 ClO3− to Cl−:per minute:per Ru.
  • 9. The catalyst of claim 1, comprises Ru0 and Pd0 on a support that comprises carbon, and which can reduce ClO3− to Cl− in water at a pH in the range from about 5 to about 8, with a turnover frequency of at least 11 ClO3− to Cl−:per minute:per Ru.
  • 10. A method comprising contacting a water sample that comprises ClO3− with a catalyst as described in claim 1, under conditions such that at least some of the ClO3− is reduced to Cl−.
  • 11. The method of claim 10, which is carried out at a pH of at least about 7.
  • 12. The method of claim 10, wherein the water sample further comprises sulfate.
  • 13. A method comprising sequentially reducing 1) PdII, PtII, RhIII, or IrIII and 2) RuIII on a support to provide a catalyst.
  • 14. The method of claim 13, wherein PdII is reduced to Pd0 on a support that comprises carbon and RuIII is reduced to Ru0 on the support to provide a Ru—Pd/C catalyst.
  • 15. The method of claim 14, wherein the reducing is carried out at a temperature in the range from about 4° C. to about 25° C.
  • 16. The method of claim 13, wherein the reducing is carried out under a mixture of an inert gas (e.g., N2 or Ar) and H2.
  • 17. A method comprising contacting a water sample that comprises NO3− with a catalyst as described in claim 1, under conditions such that at least some of the NO3− is reduced to NH4+.
  • 18. A method comprising contacting a water sample that comprises NO3− with a catalyst that comprises in situ reduced Ru particles and another metal, under conditions such that at least some of the NO3− is reduced to NH4+.
  • 19. The method of claim 18, wherein the catalyst comprises Os—Ru/C, Ru—Ru/C, Ru—Rh/C, Ru—Ir/C, Ru—Pt/C or Ru—Pd/C.
  • 20. A method comprising contacting a water sample that comprises NO3− with a catalyst that comprises Ru—Pd/C or In—Pd/C, under conditions such that at least some of the NO3− is reduced to NH4+.
STATEMENT REGARDING FEDERALLY SPONSORED RESEARCH

This invention was made with government support under 1932942 awarded by the National Science Foundation. The government has certain rights in the invention.

Provisional Applications (1)
Number Date Country
63410083 Sep 2022 US