Electrocatalytic Conversion of Nitrates and Nitrites to Ammonia
Electrocatalytic reduction of waste nitrates (NO3−) enables the synthesis of ammonia (NH3) in a carbon neutral and decentralized manner. The present invention uses atomically dispersed transition metal-nitrogen-carbon (M-N—C) catalysts with varying metal centers to uniquely favor mono-nitrogen products (e.g., NH3) via the electrochemical nitrate reduction reaction (NO3RR).
This application claims the benefit of U.S. Provisional Appl. No. 63/324,537, filed Mar. 28, 2022, which is incorporated herein by reference.
STATEMENT OF GOVERNMENT INTERESTThis 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.
STATEMENT REGARDING PRIOR DISCLOSURES BY THE INVENTOR OR A JOINT INVENTORThe following disclosure is submitted under 35 U.S.C. 102(b)(1)(A): Eamonn Murphy, Yuanchao Liu, Ivana Matanovic, Shengyuan Guo, Peter Tieu, Ying Huang, Alvin Ly, Suparna Das, Iryna Zenyuk, Xiaoqing Pan, Erik Spoerke, and Plamen Atanassov, “Highly Durable and Selective Fe- and Mo-based Atomically Dispersed Electrocatalysts for Nitrate Reduction to Ammonia via Distinct and Synergized NO2− Pathways,” ACS Catalysis 12, 6651 (2022). The subject matter of this disclosure was conceived of or invented by the inventors named in this application.
FIELD OF THE INVENTIONThe present invention relates to ammonia synthesis and, in particular, to the electrocatalytic conversion of nitrates and nitrites to ammonia.
BACKGROUND OF THE INVENTIONRecently, there has been a surge in global research efforts towards the electrochemical production of ammonia (NH3). See R. Society, Ammonia: Zero-Carbon Fertiliser, Fuel and Energy Store, The Royal Society London, UK (2020); and Y. Ren et al., Energy Environ. Sci. 14(3), 1176 (2021). NH3 is critical to produce N-based fertilizers and shows promise as an energy vector and carbon free liquid fuel. See R. Society, Ammonia: Zero-Carbon Fertiliser, Fuel and Energy Store, The Royal Society London, UK (2020); S. Giddey et al., Int. J. Hydrogen Energy 38(34), 14576 (2013); A. Grinberg Dana et al., Angew. Chemie—Int. Ed. 55(31), 8798 (2016); and C. Smith et al., Energy Environ. Sci. 13(2), 331 (2020). Currently, the near century-old Haber-Bosch (H-B) process is the only industrial NH3 manufacturing approach. However, the H-B process operates at high temperatures and pressures and utilizes gray hydrogen. See A. Klerke et al., J. Mater. Chem. 18(20), 2304 (2008). As a result, the process consumes 1-2% of global energy and produces 1-2% of global CO2 emissions. See C. Smith et al., Energy Environ. Sci. 13(2), 331 (2020); and J. Norskov and J. Chen, “Sustainable Ammonia Synthesis”, Report, DOE Roundtable 2016, 1-23. Electrochemical methods offer carbon-neutral pathways for NH3 synthesis at ambient conditions and in a distributed manner. To date, the electrocatalytic nitrogen reduction reaction (eNRR) from dinitrogen gas is the most heavily researched. However, eNRR faces several challenges such as low solubility in aqueous electrolytes, competition from the more facile hydrogen evolution reaction (HER) and large thermodynamic activation barriers. See C. Tang and S. Z. Qiao, Chem. Soc. Rev. 48(12), 3166 (2019); D. R. MacFarlane et al., Joule 4(6), 1186 (2020); and B. H. R. Suryanto et al., Nat. Catal. 2(4), 290 (2019). As a result, the Faradaic efficiencies (FEs) and yields of NH3 in aqueous systems remain low, even proving difficult for reliable analytical detection. See J. Choi et al., Nat. Commun. 11(1), 1 (2020); Y. Chen et al., Nat. Catal. 3(12), 1055 (2020); and S. Z. Andersen et al., Nature 570(7762), 504 (2019).
To circumvent these challenges, oxidized nitrogen species, such as nitrate (NO3−) can be utilized as feedstock for efficient electrocatalytic NH3 production. NO3− is the second most abundant form of nitrogen as an environmental pollutant present in industrial waste streams and water runoff at concentrations up to 2 M. See I. Katsounaros et al., J. Hazard. Mater. 171(1-3), 323 (2009); R. Chauhan and V. C. Srivastava, Chem. Eng. J. 386(3), 122065 (2020); and T. T. P. Nguyen et al., ECS Trans. 53(16), 41 (2013). Nitrate's appeal is owed to its large solubility in aqueous electrolytes and relatively low bond dissociation energy (204 kJ/mol), facilitating significantly more favorable kinetics compared to eNRR processes. See D. Xu et al., Front. Environ. Sci. Eng. 12(1), 1 (2018); W. T. Mook et al., Desalination 285, 1 (2012); and L. Su et al., Nano Lett. 19(8), 5423 (2019). The electrocatalytic nitrate reduction reaction (NO3RR) has been studied extensively for denitrification and water purification. See D. Xu et al., Front. Environ. Sci. Eng. 12(1), 1 (2018); W. T. Mook et al., Desalination 285, 1 (2012); and H. Liu et al., ACS Catal. 2(3), 8431 (2021). Early NO3RR research focused on selective reduction of NO3− to ½N2 via a 5e− transfer pathway (E0=1.25 V vs. RHE) and has been validated on a variety of bulk metals, heterogeneous catalysts (Ru, Rh, Pd, Ir, Pt, Cu, Ag and Au), and their alloys. See L. Su et al., Nano Lett. 19(8), 5423 (2019); J. Martinez et al., Appl. Catal. B Environ. 207, 42 (2017); M. T. De Groot et al., J. Electroanal. Chem. 562, 81 (2004); M. D'Arino et al., Appl. Catal. B Environ. 53(3), 161 (2004); A. Pintar et al., Appl. Catal. B Environ. 52(1), 49 (2004); X. Huo et al., Appl. Catal. B Environ. 211, 188 (2017); A. H. Pizarro et al., J. Environ. Chem. Eng. 3(4), 2777 (2015); M. Yanauchi et al., J. Am. Chem. Soc. 133, 1150 (2011); E. Perez-Gallent et al., Electrochim. Acta 227, 77 (2017); J. Yang et al., Chem. Commun. 50, 2148 (2014); and F. Calle-Vallejo et al., Phys. Chem. Chem. Phys. 15, 3196 (2013).
Recently, significant attention has been shifted to selectively reducing NO3− to NH3 via an 8e− transfer pathway (E0=0.82 V vs. RHE). The 8e− transfer pathway is highly complex, involving many potential reaction intermediates (NO2, NO2−, NO, N2O, N2, NH3, NH2OH and N2H4). The initial 2e− transfer reduction of adsorbed *NO3− to *NO2− is thought to be universal in the NO3RR pathways and is regarded as the rate-limiting step. See Z. Wang et al., Catal. Sci. Technol. 11, 705 (2021); Y. Zeng et al., Small Methods 4(12), 1 (2020); and M. Duca and M. T. Koper, Energy Environ. Sci. 5, 9726 (2012). Further reduction to *NO is regarded as the selectivity determining step on extended lattice surfaces. From *NO, the reaction pathways diverge selectively towards NH3 or N2, depending on electrolyte conditions and intermediates' adsorption strength on catalyst surfaces. Several recent computation studies have examined the selectivity of the *NO to subsequent reaction products (N2, NH2OH, and NH3) over a variety of transition metal surfaces. See C. A. Casey-Stevens et al., Appl. Surf. Sci. 552, 149063 (2021); H. J. Chun et al., ACS Catal. 7, 3869 (2017); and H. Wan et al., Angew. Chem. 133, 22137 (2021). By contrast, single-atom catalysts (SACs) provide an intrinsic advantage favoring NH3 production over N2 products. By creating isolated single-atom active sites, the possibility for two mono-nitrogen moieties can be activated on adjacent sites and form coupled N2 products (N2O, N2, N2H4) is minimized, narrowing product selectivity largely to NH3. See Z. Wu et al., Nat. Commun. 12(1), 1 (2021). Recent work reported the feasibility of Fe-based SACs for NO3RR and some theoretical work indicated that single-atom active sites preferentially adsorb NO3− over H+, reducing competition from the HER. See Z. Wu et al., Nat. Commun. 12(1), 1 (2021); P. Li et al., Energy Environ. Sci. 14, 3522 (2021); H. Niu et al., Adv. Funct. Mater. 31(11), 2008533 (2020); J. Wu et al., Catal. Sci. Technol. 11, 7160 (2021); and J. Wu et al., J. Phys. Chem. Lett. 12, 3968 (2021). Although good progress has been made in the field of nitrate reduction for NH3 production, developing well-established reaction mechanisms and electrocatalysts with long-term operational stability remains a significant challenge. See P. H. van Langevelde et al., Joule 5(2), 290 (2021).
SUMMARY OF THE INVENTIONThe present invention is a method for the electrocatalytic conversion of nitrates and nitrites to ammonia, comprising providing an electrochemical cell, the electrochemical cell comprising a working electrode comprising an atomically dispersed transition metal-nitrogen-carbon (M-N—C) catalyst, a counter electrode, a hydrogen permeable membrane between the working and counter electrodes, and an electrolyte comprising an aqueous solution of nitrate or nitrite; and applying an electrode potential between the working electrode and the counter electrode sufficient to electrocatalytically reduce the nitrate or nitrite to ammonia at the working electrode. For example, the transition metal can comprise Fe, Mo, Cr, Mn, Co, Ni, Cu, Ru, Rh, Pd, W, La, or Ce, or a combination thereof. For example, the M-N—C catalyst can be a monometallic catalyst or a bimetallic catalyst, such as FeMo. The electrode potential can be more negative than −0.1 V.
The electrochemical nitrate reduction reaction (NO3RR) not only promises an effective route to sustainable ammonia synthesis in a carbon-neutral and decentralized manner, but also offers potential impact to critical wastewater remediation. The bioinspired, atomically dispersed catalysts of the present invention have exception efficacy for NO3RR in aqueous media via a catalytic cascade. Atomically dispersed catalysts have intrinsic selectivity towards mono-nitrogen species over their dinitrogen counterparts. As an example, a series of nitrogen-coordinated mono- and bimetallic, atomically dispersed, iron- and molybdenum-based electrocatalysts were evaluated for ammonia synthesis via the NO3RR. The key role of the nitrite *NO2/NO2− intermediates was identified both computationally and experimentally, wherein the Fe—N4 sites and Mo—N4/*O—Mo—N4 sites carried distinct associative and dissociative adsorption of NO3− molecules, respectively. By integrating individual Fe and Mo sites on a single bimetallic catalyst, the unique reaction pathways were synergized, achieving a Faradaic efficiency of 94% toward ammonia. The utilization of catalytic cascades, synergizing distinct reaction pathways on heterogeneous single-atom sites, is largely unconstrained by linear scaling relations of reaction intermediates. The invention enables electrocatalysts for highly selective, efficient, and durable ammonia synthesis.
The detailed description will refer to the following drawings, wherein like elements are referred to by like numbers.
In nature, the NO3RR pathway is carried by enzymatic cascades, wherein the conversion of NO3− to NO2− and subsequent reduction of NO2− to NH3 are catalyzed by the nitrate reductase Mo cofactor and nitrite reductase Fe cofactor, respectively. See R. D. Milton et al., Chempluschem 82(4), 513 (2017); and L. Y. Stein and M. G. Klotz, Curr. Biol. 26(3), R94 (2016). This indicates the unique role of different metal elements for nitrogen transformation. As described below, an embodiment of the present invention uses a highly efficient synthetic electrocatalytic cascade utilizing chemical dissociation of the NO3− ion to NO2− over single-atom Mo sites, followed by a kinetically fast 6e− transfer reduction of NO2 to NH3 at 100% FE over single-atom Fe sites.
The biological enzymatic process indicates the unique role of different metal elements in nitrogen transformation. An embodiment of the invention uses atomically dispersed mono- and bimetallic Fe- and Mo-based electrocatalysts for the NO3RR. As will be described below, distinct NO3− to *NO2/NO2− intermediate pathways were identified both computationally and experimentally. The initial NO3− adsorption was found to be dissociative on Mo sites and associative on Fe sites, showing great biological similarity. By integrating both Mo and Fe sites on a bimetallic catalyst, the two reaction mechanisms were synergized, and the synthetic catalytic cascade achieved a NO3RR FE for NH3 of 94% and a yield of 18.0 μmol cm−2 hr−1 (153 μgNH3 mgcat−1 h−1). Moreover, durability studies showed a well-maintained FE above 90% over a 60 h electrolysis. These properties enable synergizing distinct reaction pathways on atomically dispersed electrocatalysts for highly selective and efficient NH3 synthesis.
Synthesis of Fe, Mo and FeMo—N—C CatalystsThe transition metal-nitrogen-carbon (M-N—C) catalysts are based on transition metal (M) ions coordinated with nitrogen (N) atoms and embedded into a graphitic carbon (C) matrix. See T. Asset et al., Joule 4, 33 (2020); and K. Kumar et al., ACS Catal. 11, 484 (2021), which are incorporated herein by reference. Atomically dispersed mono- and bimetallic catalysts can be synthesized using the well-established sacrificial support method (SSM). See A. Serov et al., Adv. Energy Mater. 4(10), 201301735 (2014); A. Serov et al., Nano Energy 16, 293 (2015); T. Asset et al., ACS Catal. 9(9), 7688 (2019); and A. Serov et al., Electrochim. Acta 179, 154 (2015), which are incorporated herein by reference. The SSM class of M-N—C catalysts has been widely studied due to their good activity and stability for the oxygen reduction reaction (ORR), where the oxidative cathodic potentials (0.6-1 V) are even harsher for carbon corrosion as compared to the reductive potentials in NO3RR (−0.8 to −0.2 V). See T. Asset et al., Joule 4, 33 (2020); and M. Primbs et al., Energy Environ. Sci. 13, 2480 (2020). The outstanding ORR performance for the SSM M-N—C was largely attributed to its high degree of graphitization and robust M-Nx sites, which also greatly contributed to the excellent stability of the FeMo—N—C for NO3RR, as will be described below. See Y. Chen et al., Mater. Today 53, 58 (2020).
As examples of the invention, atomically dispersed Fe—N—C and Mo—Ni—C monometallic catalysts and FeMo—N—C bimetallic catalysts were synthesized using the sacrificial support method. Iron nitrate nonahydrate and ammonium molybdenum tetrahydrate were selected as the iron and molybdenum precursors, respectively, for the catalyst synthesis. First, a calculated amount of metal nitrate precursor (Fe—N—C precursor=0.6 g, Mo—N—C precursor=0.26 g) was combined with silica of varying surface areas, consisting of in-house Stöber spheres, LM-150 and OX-50 (0.5, 1.25, and 1.25 g, respectively). Nicarbazin (6.25 g) was added to the precursor mixture. For the molybdenum-containing catalysts, a calculated amount of urea was added to assist in reducing the molybdenum precursor. The amount of metallic precursor was calculated such that the number of metal atoms between the mono- and bimetallic catalysts was constant. Next, MilliQ water was added to the precursor ratio and sonicated for 30 min. The viscous mixture was then set at 45° C. under constant stirring until dry. The material was then ground and ball-milled for 1 h at 45 Hz. A first pyrolysis was performed at 650° C. for 45 min under a reductive 7% H2-93% Ar atmosphere. The material was then etched 4 days in a 40 wt % hydrofluoric acid mixture (2:1) HF/H2O. The material was then washed to a neutral pH before being dried and again ball-milled. Next the material underwent a second pyrolysis at 650° C. for 30 min under a 10% NH3-90% Ar atmosphere, followed by a final ball-milling.
Electrochemical MeasurementsThe electrocatalytic conversion of nitrates and nitrites to ammonia can be accomplished in an electrochemical H-cell, as shown in
The working electrode was a carbon paper cut to a geometric surface area of 0.45 cm2. A catalyst ink was prepared by dispersing 5 mg of the M-N—C catalyst in 300 μL of MilliQ water, 670 μL of isopropanol and 30 μL of a 5 wt % Nafion solution and ultrasonicated for 1 hour. 100 μL of the catalyst ink was drop cast onto a carbon paper working electrode for a total catalyst loading of 0.5 mg/cm2 and dried in a vacuum oven at 60° C. overnight. A standard three-electrode system was used with a reversible hydrogen electrode and a graphite rod as the reference and counter electrode, respectively. Electrochemical nitrate and nitrite reduction reactions were carried out in a customized glass electrochemical H-cell comprising working and counter electrode chambers separated by a Nafion 211 membrane at ambient conditions. The electrolyte used for all nitrate reduction experiments is 0.05 M phosphate buffer solution (PBS) and 0.16 M NO3− (KNO3). Prior to an electrochemical test, N2 gas was purged in both the working and counter chambers for 30 min at 80 sccm, which contain 30 mL and 25 mL of electrolyte, respectively. Electrochemical nitrite reduction experiments were performed analogously to nitrate reduction experiments, utilizing the same H-cell and three electrode system. The electrolyte was 0.05 M PBS and 0.01 M NO2− (KNO2). Potentiostatic tests were performed at applied potentials of between −0.2 and −0.70 V vs RHE.
Distinct *NO2/NO2− Intermediate Pathways Over Fe and Mo Active Sites
In biological NO3− reduction pathways, Mo-based active centers reduce NO3− to NO2− and Fe-based active centers further reduce NO2− to NH3 (or N2). To evaluate the catalytic activity of atomically dispersed Fe and Mo sites toward the NO3RR to NH3, a series of atomically dispersed M-N—C(M=Fe, Mo, and FeMo) electrocatalysts were employed.
Density Functional Theory (DFT) calculations and experimental NO3− reduction time-course studies predicted and confirmed two different reaction mechanisms over the M-N4 sites. All the DFT calculations were performed using a generalized gradient approximation approach and projector augmented-wave pseudopotentials as implemented in the Vienna Ab initio Simulation Package. See P. E. Blöchl et al., Phys. Rev. B 49(23), 16223 (1994); D. Joubert, Phys. Rev. B 59(3), 1758 (1999); G. Kresse and J. Hafner, Phys. Rev. B 47(1), 558 (1993); G. Kresse and J. Hafner, Phys. Rev. B 49(20), 14251 (1994); and G. Kresse and J. Furthmüller, Comput. Mater. Sci. 6(1), 15 (1996). As shown in
By integrating both Mo and Fe sites in a bimetallic catalyst (FeMo—N—C), a synergized pathway was proposed to optimize the NH3 production with a catalytic cascade, as shown in
By mimicking biological pathways and synergizing Mo and Fe sites, the activity of the bimetallic catalyst for the NO3RR was significantly increased compared with its monometallic counterparts. To further explore the catalyst activity and reaction mechanism, detailed potential-dependent studies were performed for the nitrite reduction reaction (NO2RR) and NO3RR processes.
The cathodic current and reaction onset potential were first evaluated by linear sweep voltammetry (LSV) in a standard 0.05 M phosphate-buffered solution (PBS) electrolyte. Employing the bioinspired bimetallic FeMo—N—C as an example, the LSV curves in
The 6e− transfer reduction of NO2− ions or a *NOOH intermediate to NH3 is a critical step in the NO3RR and the key in synergizing the dissociative and associative reaction mechanisms as shown in
Based on this understanding of the NO2RR activities, the more complex electrochemical nitrate (NO3−) reduction was investigated.
In agreement with
Therefore, by integrating heterogenous Mo sites, which readily reduce NO3− to NO2−, and Fe sites, which are limited by NO3− to NO2− reduction, into a single bimetallic catalyst (FeMo—N—C), a synergistic catalyst system can be employed to optimize this multistage NO3RR chemistry. As shown in
A recent review by Koper and co-workers highlighted that long-term durability studies are an immediate need for the future development of ammonia synthesis via the NO3RR. See P. H. van Langevelde et al., Joule 5(2), 290 (2021). Usually, durability was claimed from single or several cycles of short-term electrolysis (0.5-4 h). For example, titanium electrodes were reported to achieve a FE of 82% for the first 2 hours, after which the FE dropped to 40%, showing change of catalytic performance over long-term electrolysis. See J. M. McEnaney et al., ACS Sustain. Chem. Eng. 8(7), 2672 (2020). Therefore, durability studies were performed on the FeMo—N—C catalyst at −0.45 V for a total operation time of 60 h, broken into five 12 h segments, each with a refreshed electrolyte, as shown in
Isotopic analysis is essential to confirm the electrochemical transformation of N species to NH3, especially with N-doped catalysts. A 6 h electrolysis was conducted at −0.45 V with a 15NO3− feed at the same concentration.
Inductively coupled plasma mass spectrometry (ICP-MS) results showed a metal loading of 0.73 wt % for Fe—N—C and 1.36 wt % for Mo—N—C, while the bi-metallic FeMo—N—C preserved 0.38 wt % Fe and 0.77 wt % Mo. Raman spectroscopy showed that the three electrocatalysts had a similar degree of graphitization. The resulting carbon matrix exhibited a hierarchical porous structure with leading mesoporosity around 20 nm and 100 nm, as shown by nitrogen physisorption results and scanning electron microscopy (SEM) images. The N2 sorption isotherm also confirmed a large amount of micropores that contributed to a Barrett-Emmett-Teller (BET) surface area of ca. 600 m2/g for all catalysts. The hierarchical pore structure was favorable for promoting mass transport of the electrolyte to the active sites. X-ray diffraction (XRD) patterns for the mono- and bi-metallic catalysts (
DFT was further used to study the mechanism of NO3− conversion to NO2− and *NO2 on the M-N4 centers of the Fe—N—C and Mo—N—C catalysts, illustrated in
Both associative and dissociative adsorption of NO3− molecules were considered as the initial step of the NO3RR (
In contrast, the Mo—N4 sites showed a very strong oxygen affinity, such that when the NO3− molecule was coordinated on the Mo site through a single O-atom, the N—O bond was chemically cleaved (−4.2 eV), leading to an oxygen atom adsorbed (*O) on Mo site and a desorbed NO2− ion, as shown in
Interestingly, further investigation showed that due to the strong O affinity of the Mo active site, high coordination number, and orientation of its 4d-orbitals, the oxygenated *O—Mo site can simultaneously bind an additional NO3− molecule (
In the calculations, the dissociative-adsorption pathway was also found to be thermodynamically plausible over Fe sites, as shown by
As described above, the experimental results gave strong evidence of the excellent NO2− reduction capability of Fe—N—C catalyst, good NO2− generation capability of Mo—N—C catalysts, and synergistic effect of bimetallic FeMo—N—C catalysts (
In summary, taking inspiration from the cascade mechanism of biological nitrate/nitrite reductase enzymes, distinct NO3RR mechanisms were identified for single-atom Mo—N4 and Fe—N4 active sites, where Mo sites preferentially followed a dissociative-adsorption pathway, spontaneously breaking the N—O bond, releasing NO2− into the bulk and creating the *O—Mo active site for the continuous catalytic cycle of NO2− generation. Then, the NO2− molecules could be re-adsorbed and further reduced to NH3, following a 2e−+6e− process. Over Fe sites, an associated-adsorption and direct 8e− transfer pathway was favorable in which NO3− adsorbs and was directly reduced to NH3. Additionally, independent NO2RR electrolysis revealed the ability of the Fe sites to reduce NO2− to NH3 at 100% FENH3 and high yields. DFT results further revealed that the Mo—N—C catalyst was more efficient in the first 2e− transfer step than its Fe—N—C counterpart, which enables the synergistic effect of the bimetallic FeMo—N—C catalysts in terms of both NH3 selectivity and yield rate. Specifically, the cascade NO3RR synergized the favorable dissociation of NO3− to NO2− over Mo sites with the fast kinetics of Fe sites in reducing NO2− to NH3. As a result, selective NO3RR to NH3 at a high FENH3 of 94% was achieved at −0.45 V, outperforming either of the Fe—N—C or Mo—N—C mono-metallic catalysts. Additionally, long-term durability tests over 60 hours demonstrated the robustness of the FeMo—N—C catalyst, maintaining a FENH3 over 90%. Therefore, heterogeneous single-atom sites utilizing distinct reaction pathways can be synergized on a single electrocatalyst to achieve highly selective and efficient NO3− reduction to NH3.
Other Transition Metal M-N—C CatalystsOther atomically dispersed transition metal and rare earth element M-N—C catalysts (M=Cr, Mn, Fe, Co, Ni, Cu, Mo, Ru, Rh, Pd, W, La, and Ce) can also be used for the electrocatalytic reactions. In order to characterize the utility of these catalysts for the electrocatalytic conversion of nitrates and nitrites, electrochemical descriptors were derived experimentally through the reaction onset potential (HER, NO3RR and NO2RR), the NO3RR selectivity to both NO2− and NH3, and the NO2RR selectivity to NH3. Isotopically doped 15NO2− in the NO3RR elucidated the complex production/consumption mechanism of the NO2− intermediate. Density functional theory (DFT) was employed to evaluate the Gibbs free energy for the NO3RR and NO2RR following either the dissociative adsorption or associative adsorption pathway. By relating the experimentally determined and computationally determined activity descriptors, strong correlations were observed for the NO2RR (R=0.72) the NO3RR (R=0.73) selectivity to NH3. This work bridges the gap between computation and experiment for the NO3− and NO2− reduction reactions over atomically dispersed M-N—C catalysts by providing a powerful set of activity descriptors that correlate strongly with the experimentally observed activity. These descriptors can be utilized to guide future atomically dispersed M-Nx catalyst development for highly active and efficient NO3− reduction to NH3.
Synthesis of Transition Metal M-N—C CatalystsAtomically dispersed M-N—C catalysts were synthesized through the well-established sacrificial support method (SSM). See A. Serov et al., Nano Energy 16, 293 (2015); M. M. Hossen et al., J. Power Sources 375, 214 (2018); and A. Serov et al., Electrochim. Acta 179, 154 (2015). As described above, SSM has been extensively applied for atomically dispersed Fe—N—C catalysts for the ORR and CO2RR. See also T. Asset et al., ACS Catal. 9(9), 7668 (2019); and T. Asset and P. Atanassov, Joule 4(1), 33 (2020). Extending beyond Fe—N—C to a variety of 3d-, 4d-, 5d- and f-block metals, a set of atomically dispersed M-N—C catalysts (M=Cr, Mn, Fe, Co, Ni, Cu, Mo, Ru, Rh, Pd, La, Ce, and W), as shown in
Synthesis of Mn, Fe, Co, Ni, Cu, Mo and W—N—C catalysts. First a slurry of a carbon-nitrogen containing precursor, Nicarbazin (6.25 g), a silica sacrificial support comprising LM-150 fumed silica (1.25 g), OX-50 (1.25 g) and Stöber spheres (0.5 g), and the corresponding metal salt precursor (manganese(II) nitrate tetrahydrate, Mn=0.266 g; iron(III) nitrate nonahydrate, Fe=0.60 g; cobalt(II) nitrate hexahydrate, Co=0.272 g; nickel(II) nitrate hexahydrate, Ni=0.271 g; copper(II) nitrate hemi pentahydrate, Cu=0.345 g; ammonium molybdate tetrahydrate, Mo=0.262 g; or ammonium paratungstate, W=0.095 g), in 50 mL of MilliQ water was created. Next, the slurry was sonicated for 30 minutes before being dried overnight at 45° C., under constant stirring. The mixture was then further dried in an oven at 45° C. for 24 hours. The resulting powder was then ball milled at 45 Hz for 1 hour. The catalyst powder was then pyrolyzed at 975° C. (with a ramp rate of 15° C./min) under a reductive H2/Ar (7%/93%) atmosphere for 45 min. The pyrolyzed powder was then ball milled a second time at 45 Hz for 1 hour. The silicate template was then etched in a hydrofluoric acid (15M) solution for 96 hours. The catalyst was then recovered by filtration and washed to neutral pH, followed by drying at 60° C. overnight. The catalyst underwent a second pyrolysis in a reductive NH3/N2 (10%/90%) atmosphere at 950° C. for 30 min (with a ramp rate of 20° C./min). The catalyst was then ball milled a final time at 45 Hz for 1 hour.
Synthesis of Cr—N—C catalyst. The synthesis is identical to the previous procedure, with only the pyrolysis temperatures being reduced to 650° C. in both the first and second pyrolysis, to maintain an atomic dispersion of the Cr metal. The metal salt precursor loading of chromium(III) acetylacetonate was Cr=0.519 g.
Synthesis of Ru—N—C catalyst. The synthesis is identical to the previous procedure, with the pyrolysis temperatures being reduced to 650° C. and an inert argon pyrolysis atmosphere, to maintain an atomic dispersion of the Ru metal. The metal salt precursor loading of ruthenium(III) nitrosylnitrate was Ru=0.235 g.
Synthesis of Rh and Pd—N—C catalysts. The synthesis is identical to the previous procedure, with the high pyrolysis temperatures of 975° C. and 950° C. for the first and second pyrolysis, respectively; however the pyrolysis is performed in an inert argon atmosphere, to maintain an atomic dispersion of the Rh and Pd metals. The metal salt precursor loading of rhodium(III) nitrate hydrate was Rh=0.215 g, and for palladium(II) nitrate dihydrate was Pd=0.198 g.
Synthesis of the La and Ce—N—C catalysts. The synthesis is identical to the previous procedure, with the pyrolysis temperatures being reduced to 650° C. under an inert argon pyrolysis atmosphere. Furthermore, the etching of the silica template was performed in an alkaline 4M NaOH environment at 80° C. for 96 hours. If etched using a hydrofluoric acid environment, the La and Ce readily form LaF3 and CeF3 nanoparticles. The metal salt precursor loading of lanthanum(III) nitrate hexahydrate was La=0.060 g, and for cerium(III) nitrate hexahydrate was Ce=0.061 g.
Physical Structure of Single-Atom Metal CentersX-ray diffraction (XRD) shows only the characteristic (002) and (100) peaks for carbon, confirming the absence of bulk metallic phases.
The electron energy loss spectroscopy (EELS) in
X-ray photoelectron spectroscopy (XPS) was used to further examine the metal-nitrogen coordination and other nitrogen moieties present. The representative N 1s spectra for the Fe—, Rh—, and La—N—C catalysts in
In summary, 13 atomically dispersed M-N—C catalysts were synthesized using the SSM. By combining AC-STEM, EELS, XAS and XPS, the atomically dispersed nature of each metal site was comprehensively visualized and confirmed to be in a M-Nx coordination. Critically, the well-deciphered coordination environment of the metal center allows elucidation of the intrinsic activity of the different M-Nx moieties towards the electrocatalytic transformation of reactive N-species.
Electrochemical PerformanceThe NO3RR is somewhat universal in that even metal-free nitrogen-doped-carbon (N—C) showed limited but observable activity towards the generation of NH3 and NO2− under an isotopically labeled 15NO3− feed at neutral pH. With the addition of nitrogen-coordinated metal sites Cr—N—C as an example, the FENH3 increased to 91.2%. The introduction of Fe—Nx sites boosted the FENH3 to 99% at −0.4 (V vs. RHE) and the corresponding YieldNH3 increased from 1.3 to 10.0 μmol h−1 cm−2.
It should be noted that, for the NO3RR activity, as compared to neutral media, the current density and NH3 yield rate could be significantly enhanced under alkaline conditions, while maintaining a high FENH3. Given Fe—N—C as an example, when using a 1M KOH electrolyte (pH=14), the NH3 partial current density increased from 2 mA cm−2 (11.4 μmol h−1 cm−2) to 35 mA cm−2 (156.5 μmol h−1 cm−2) at −0.4 V and further to 175 mA cm−2 (750.4 μmol h−1 cm−2) at −0.6 V, both with a FENH3 above 95%. Other work on alkaline NO3RR also found similar trends. See Y. Wang et al., J. Am. Chem. Soc. 142(12), 5702 (2020); P. Li et al., Energy Environ. Sci. 14(6), 3522 (2021); and J. Li et al., J. Am. Chem. Soc. 142(15), 7036 (2020). However, a neutral pH (pH 6.3) was used herein in order to evaluate the intrinsic activity of the M-Nx sites toward the electrocatalytic NO3RR and NO2RR.
Regardless of the NO3RR pathways, the surface intermediates *NO2 and *NOOH are inevitable before the reaction pathway diverges to its several possible products. See Y. Wang and M. Shao, ACS Catal. 12(9), 5407 (2022); H. Niu et al., Adv. Funct. Mater. 2008533(3), 1 (2020); and S. Wang et al., Nano Energy, 107517 (2022). The NO2− molecule is also the first desorbable reaction intermediate, playing a key role in the NO3RR via a cascade pathway. See E. Murphy et al., ACS Catal. 12(11), 6651 (2022); and H. Liu et al., ACS Catal. 2(3), 8431 (2021). For the present set of M-N—C catalysts, the NO2RR displayed an earlier activation than the NO3RR by linear sweep voltammetry (LSV), as shown in
To investigate the NO3RR activity and selectivity towards NH3 and NO2−, a series of 2-hour potential holds were performed at potentials between −0.2 V and −0.8 V. As shown in
For the role of potential nitrite (as by-product or intermediate), previous reports usually considered the NO3RR as a direct 8e− transfer pathway with certain irreversible NO2− desorption or leaching. See Z. Wu et al., Nat. Commun. 12(1), 2870 (2021); P. Li et al., Energy Environ. Sci. 14(6), 3522 (2021); and G. F. Chen et al., Nat. Energy 5(8), 605 (2020). Here, doping of isotopic 15NO2− in the NO3RR, schematically shown in
To examine the relationship between the NO2RR and NO3RR,
Computational Descriptors for the NO3RR and NO2RR
As discussed above, the first 2e− transfer is the rate limiting step in the NO3RR, therefore DFT was used to study the energetics of key intermediates in the conversion of NO3−-to-NO2− on various catalytic active sites. This included the *NO3 surface intermediate formed by oxidative associative adsorption of NO3−, the *O surface intermediate formed by neutral dissociative adsorption of NO3−, and the *NO2 surface species formed in the reductive adsorption of NO3−; these result in four thermodynamic reaction descriptors, as shown in
The quadrant plot in
To evaluate the practical relevance of these computational descriptors, a set of correlations were developed between the DFT-derived free energies (ΔrG) in
However, the NO3RR FENO2− showed minimum correlations with either ΔrG [*NO2→*+NO2− ] or ΔrG [*+NO3−→*O+NO2−]. This agrees well with the above-mentioned poor correlation between the NO3RR FENO2− and NO2RR YieldNH3 (
In summary, a rich set of atomically dispersed 3d-, 4d-, 5d- and f-block M-N—C catalysts with a well-established M-Nx coordination was synthesized. The gap analysis plot revealed diverse NO3RR performance, wherein Cr—N—C and Fe—N—C were the most selective catalysts, achieving near 100% FENH3, while Mn—, Cu— and Mo—N—C were highly selective for NO2−. For the NO2RR, several elements including Cr—, Fe—, Co—, Ni—, Cu— and La—N—C achieved a FENH3 of 100% at high overpotentials, with Fe and Co—N—C showing 100% FENH3 over the entire potential range. Isotopically doped 15NO2− in concentrated 14 NO3− demonstrated the ease at which minute concentrations of NO2− in the bulk electrolyte can preferentially reduce to NH3, convoluting the possibility of a direct 8e− pathway or a 2e−+6e− cascade pathway with NO2− as transient intermediate for the NO3RR. The correlation between experimental NO3RR ammonia selectivity and experimental NO2RR activity suggested a universal contribution of the NO2− intermediate in NO3RR (2e−+6e−). The DFT-derived thermodynamic descriptors theoretically explain the electrocatalytic selectivity for each metal center. Furthermore, these computational descriptors showed strong correlations with the experimental performances, such that a simple computationally evaluated descriptor can be utilized to estimate the activity of M-N—C catalysts for the NO3RR and NO2RR. While these computational-experimental activity descriptors provide strong correlations for the NO3RR and NO2RR activity, these correlations are limited when the H+ adsorption is a competing factor leading to the HER over NO3RR (Ru— and Rh—N—C) and when there is poor NO3− activation in contrast with significant NO2− activation (Co—N—C). This work enables the design of tandem NO3RR systems and nitrate-containing systems composed of either multi-metallic M-N—C catalysts or extended catalytic surfaces supported by synergistic M-N—C supports.
The present invention has been described as electrocatalytic conversion of nitrates and nitrites to ammonia. 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.
Claims
1. A method for the electrocatalytic conversion of nitrates and nitrites to ammonia, comprising:
- providing an electrochemical H-cell, comprising a working electrode comprising an atomically dispersed transition metal-nitrogen-carbon (M-N—C) catalyst, a counter electrode, a hydrogen-permeable membrane between the working electrode and the counter electrode, and an electrolyte comprising an aqueous solution of nitrate and/or nitrite; and
- applying a cathodic potential between the working electrode and the counter electrode sufficient to electrocatalytically reduce the nitrate and/or nitrite to ammonia at the working electrode.
2. The method of claim 1, wherein the transition metal comprises Fe, Mo, Cr, Mn, Co, Ni, Cu, Ru, Rh, Pd, W, La, or Ce, or a combination thereof.
3. The method of claim 1, wherein the M-N—C catalyst comprises a monometallic catalyst.
4. The method of claim 1, wherein the M-N—C catalyst comprises a bimetallic catalyst.
5. The method of claim 1, wherein the nitrate comprises potassium nitrate.
6. The method of claim 1, wherein the nitrite comprises potassium nitrite.
7. The method of claim 1, wherein the electrolyte comprises a concentration of greater than 0.01 M nitrate or nitrite.
8. The method of claim 1, wherein the electrolyte is alkaline.
9. The method of claim 1, wherein the electrolyte is neutral or acidic.
10. The method of claim 1, wherein the cathodic potential is more negative than −0.1 V.
Type: Application
Filed: Mar 28, 2023
Publication Date: Oct 3, 2024
Inventors: Plamen Atanassov (Irvine, CA), Eamonn Murphy (Irvine, CA), Yuanchao Liu (Irvine, CA), Erik David Spoerke (Albuquerque, NM)
Application Number: 18/127,111