Abstract
Activation of high-energy triple-bonds of N2 is the most significant bottleneck of ammonia synthesis under ambient conditions. Here, by importing cobalt single clusters as strong electron-donating promoter into the catalyst, the rate-determining step of ammonia synthesis is altered to the subsequent proton addition so that the barrier of N2 dissociation can be successfully overcome. As revealed by density functional theory calculations, the N2 dissociation becomes exothermic over the cobalt single cluster upon the strong electron backdonation from metal to the N2 antibonding orbitals. The energy barrier of the positively shifted rate-determining step is also greatly reduced. At the same time, advanced sampling molecular dynamics simulations indicate a barrier-less process of the N2 approaching the active sites that greatly facilitates the mass transfer. With suitable thermodynamic and dynamic property, a high ammonia yield rate of 76.2 μg h–1 mg
and superior Faradaic efficiency of 52.9% were simultaneously achieved.
Keywords: nitrogen reduction reaction, rate-determining step, cobalt single cluster, mass transfer, molecular dynamics simulations
The most significant bottleneck of ambient ammonia synthesis, namely nitrogen activation, has been successfully overcome by altering the rate-determining step over cobalt single clusters-contained electrocatalyst, and superior nitrogen reduction performance was achieved.
INTRODUCTION
Ammonia (NH3) is widely considered as a critical chemical whether in agriculture or transportation [1,2], since it is the main ingredient for fertilizer production and a carbon-free energy storage intermediate with high-energy density [3]. Although an infinite nitrogen (N2) source from the atmosphere can be easily obtained, large scale ammonia production is hindered by the chemical stability of the N≡N bond (bond energy: 940.95 kJ mol−1) [4,5]. To date, the traditional Haber-Bosch process using transition metal as catalyst under drastic reaction conditions still dominates the industrial market of NH3 synthesis [6–8]. However, this typical strategy can only reach a relatively low conversion ratio (∼15%) and consumes nearly 5% of the world's natural gas [9,10]. The use of fossil fuels, at the same time, accounts for large quantities of CO2 generation into the atmosphere [11]. Therefore, a clean and sustainable strategy for NH3 production is urgently demanded for both the global population and energy.
The electrocatalytic N2 reduction reaction (NRR), using protons from water as the hydrogen source and powered by renewable electricity sources, is an alternative method to achieving N2 fixation under ambient conditions [12–16]. Theoretically, common mechanisms of the NRR start with N2 chemisorption on the catalyst’s surface, followed by the cleavage of N≡N bond and consecutive proton addition to form NH3 [17]. Yet, strong bonding energy, high ionization potential, broad HOMO-LUMO gap as well as poor electron affinity of N2 do not favor any electron transfer process, and the N2 activation process is thus commonly considered as the rate-determining step [18,19]. A catalyst which features active sites with suitable energy and symmetry of orbitals is able to bind with N2 molecules through accepting electron density from, and backdonating to, N2 [20]. The backdonation, known as the π backbonding, strengthens the catalyst–nitrogen bond, weakens the N≡N bond, and thus contributes to lowering the energy barrier of N2 activation and positively shifting the rate-determining step [19]. This process could be enhanced by strong electron-donating ability, which enables the smooth electron transfer from the active site to the N2 antibonding π-orbitals, termed the electronic promoting effect [21–23], and further benefits the eventual N2 dissociation. Unfortunately, only very few catalysts reported to date can efficiently reduce the nitrogen activation barrier, leaving the ammonia production rate and Faradaic efficiency in low level [23–25]. Hence, searching for highly active catalysts that could alter the rate-determining step of electrochemical ammonia synthesis is still a challenging goal.
Herein, we successfully demonstrate that deliberately introducing cobalt single clusters as electron-donating promoter in nitrogen-doped carbon alters the rate-determining step of ammonia synthesis from N2 cleavage to proton addition (Fig. 1a). An excellent ammonia yield rate (76.2 μg h–1 mg
) and a superior Faradaic efficiency (52.9%) were simultaneously obtained under ambient conditions. Isotopic labeling experiments and control experiments are combined to confirm that all the NH3 produced is from the N2 electrochemical reduction. Also, the catalyst is steady enough to suffer consecutive electrolysis recycle with negligible attenuation in the NRR activity and selectivity. Density functional theory (DFT) calculations reveal that N2 activation is transferred into a strong
Figure 1.
(a) Schematic illustration for the mechanism of enhanced NRR activity by introducing Co single cluster in nitrogen-doped carbon. The cyan, red, purple and gray spheres represent C, N, Co and H atoms, respectively. (b) Dark-field TEM image of CoSC-N-C showing highly dispersed Co single clusters in the material and (c) corresponding element maps showing the distribution of Co (blue), N (orange) and C (red). (d) Co K-edge X-ray absorption near-edge structure (XANES) and (e) Fourier-transformed (FT) k3-weighted extended X-ray absorption fine structure (EXAFS) spectra of CoSC-N-C, CoNP-N-C and Co foil.
thermodynamically exothermic process on cobalt single clusters, so that it is no longer the rate-determining step of ammonia synthesis. Instead, only small energy barriers exist upon NHx formation in the whole nitrogen fixation process. Also, the approaching process of N2 molecules towards the single cluster sites is confirmed to be barrier-less by molecular dynamics (MD) simulations, which greatly favors the whole nitrogen reduction process. Altering the rate-determining step of nitrogen reduction effectively leads to a desirable NRR performance, and thus provides a powerful guidance for future design of catalysts.
RESULTS AND DISCUSSION
Characterization of the CoSC-N-C catalyst
The cobalt single clusters dispersed in nitrogen-doped carbon (CoSC-N-C) were fabricated and carefully characterized. The transmission electron microscopy (TEM) image (Supplementary Fig. 1) and dark-field TEM image (Fig. 1b) of CoSC-N-C clearly show highly dispersed cobalt single clusters in the catalyst, with corresponding element mapping (Fig. 1c) demonstrating the uniform distribution of Co superimposing with C and N. Using aberration corrected high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) with sub angstrom resolution, the clusters are observed to be of small size with an average diameter of approximately 0.5 nm (Supplementary Fig. 2). The overall Co content in the CoSC-N-C is about 3.15 wt%, as determined by inductively coupled plasma optical emission spectrometry (ICP-OES) analysis. To highlight the specific role of cobalt single clusters, counterparts with cobalt nanoparticles (CoNP-N-C) and without metal (N-C) were also synthesized by replacing the Co2+/Zn2+ mixture with single Co2+ and Zn2+, respectively, under otherwise identical conditions (Supplementary Figs 3 and 4). As shown in the X-ray powder diffraction (XRD) patterns (Supplementary Fig. 5), CoNP-N-C exhibits distinct metallic cobalt diffraction, whereas the cobalt single clusters in CoSC-N-C exist as amorphous phase. The increased content of Co accounts for the improved graphitization degree, as evaluated by the Raman spectra (Supplementary Fig. 6).
To confirm the chemical state of Co species in different samples, X-ray absorption fine structure (XAFS) measurements were conducted with Co foil as reference. The Co K-edge X-ray absorption near-edge structure (XANES) shows that the absorption edge of CoSC-N-C exhibits a positive shift compared with that of Co foil, reflecting that the average valence state of Co atoms is at positive level and is higher than that in CoNP-N-C (Fig. 1d). The Fourier-transformed (FT) k3-weighted extended X-ray absorption fine structure (EXAFS) spectrum of the CoSC-N-C (Fig. 1e) shows two main peaks at about 1.5 Å and 2.2 Å, attributing to Co-N and Co-Co, respectively. In great contrast, only a strong Co-Co coordination is detected in CoNP-N-C, so that the Co atoms are present as nanoparticles in the carbon framework. The surface chemistry of different samples was further investigated by X-ray photoelectron spectroscopy (XPS, Supplementary Fig. 7). The coexistence of four different N species in different samples, namely pyridinic-N, pyrrolic-N, graphitic-N and N-oxides, was confirmed by the high-resolution N 1s spectra (Supplementary Fig. 8) [26,27]. The high-resolution Co 2p spectra (Supplementary Fig. 9) show higher valence states of Co in CoSC-N-C compared with CoNP-N-C [28,29], indicating the coordination of cobalt and nitrogen, which is in accordance with the XAFS responses. The XPS results and the XAFS responses are combined to confirm the existence of nitrogen-stabilized cobalt single clusters in CoSC-N-C. In addition, CoSC-N-C possesses a large surface area of 287.1 m2 g−1, as determined by the Brunauer-Emmett-Teller (BET) method (Supplementary Fig. 10).
Electroreduction of N2 to NH3 on the CoSC-N-C catalyst
To evaluate the NRR activity of different samples, linear sweep voltammetry (LSV) curves were first measured (Supplementary Fig. 11). The current density difference between Ar and N2 clearly affirms the contribution from the nitrogen reduction. Without cobalt sites, N-C exhibits the weakest NRR and hydrogen evolution reaction (HER) activity. While, compared with CoNP-N-C, the bigger current density gap between Ar and N2 and lower HER current density of CoSC-N-C indicate that the introduction of cobalt single clusters leads to much more improved NRR activty and selectivity than those of the cobalt nanoparticles. Then, a quantified study of the NRR ability for different samples was carried out via chronoamperometry measurement using an H-shape electrochemical cell [30]. Here, the most rigorous experimental protocol was followed for reliable proof of the NRR performance [31]. The nitrogen gas was sufficiently purified before use to avoid the possible existence of NH3 and NOx. Possible reduction products including NH3 and N2H4 were both tested (Supplementary Figs 12–15), whereas only NH3 was detected in this work (Supplementary Fig. 16). Before the NRR test, several control experiments were first carried out to ensure that no contamination was present in the feeding gas or the equipment, and the carbon paper (CP) as current collector of the cathode did not have NRR activity, so that NH3 could only be produced by N2 reduction in the presence of catalyst (Fig. 2a). Detailed comparison of the NH3 yield rates and corresponding Faradaic efficiencies of CoSC-N-C, CoNP-N-C and N-C under various applied potentials is displayed in Fig. 2b and c. Clearly, pristine N-C exhibits only negligible NRR performance. After incorporating cobalt single clusters, the CoSC-N-C, with an optimized loading of 0.5 mg cm–2 (Supplementary Fig. 17), exhibits the highest NH3 yield rate of 76.2 μg h–1 mg
at −0.2 V versus reversible hydrogen electrode (vs. RHE), far exceeding the peak value of CoNP-N-C (22.5 μg h−1 mg
) realized at an even more negative potential (−0.3 V vs. RHE). Notably, the maximal NRR Faradaic efficiency of CoSC-N-C is also obtained at −0.2 V vs. RHE (52.9%), which is more than one order of magnitude higher than the counterparts. Detailed chronoamperometry responses are shown in Supplementary Fig. 18. Also, the electrochemical active surface area (EASA) of each sample was determined by double-layer capacitance (Cdl, Supplementary Fig. 19). Expectedly, the surface-area-normalized ammonia production rate of CoSC-N-C shows an obvious advantage over other samples (Supplementary Fig. 20). The NRR activity and selectivity of CoSC-N-C remarkably stands at the top level of reported catalysts under mild conditions (Supplementary Table 1), and is thus of great significance in energy utilization. Based on the nitrogen temperature-programmed desorption (N2-TPD) spectra (Supplementary Fig. 21), CoSC-N-C exhibits the strongest N2 adsorption ability, confirming that cobalt single clusters are superior nitrogen-adsorption active sites compared with cobalt nanoparticles or nitrogen-doped carbon [32]. The gradual decrease of the NRR Faradaic efficiency of CoSC-N-C as the applied potential becomes more negative than −0.2 V vs. RHE is mainly due to the increased HER [33], as evaluated by the gas chromatography (GC) responses (Fig. 2d and Supplementary Fig. 22). In terms of selectivity, CoSC-N-C shows a much smaller proportion of the competing HER compared with CoNP-N-C and N-C, and is more inclined to proceed NRR (Supplementary Fig. 23). Based on the above results, CoSC-N-C has a prominent advantage in nitrogen reduction, especially at −0.2 V vs. RHE (Fig. 2e and Supplementary Table 2).
Figure 2.
(a) The UV-vis absorption spectra of the electrolytes after electrolysis under different conditions. (b) NH3 yield rates and (c) corresponding Faradaic efficiencies at each given potential of CoSC-N-C, CoNP-N-C and N-C. (d) H2 yield of CoSC-N-C, CoNP-N-C and N-C at different potentials. The error bars correspond to the standard deviations of the obtained data over three separately conducted electrochemical measurements under the same conditions. (e) Selectivity of NRR and HER at −0.2 V vs. RHE of different samples.
Careful examination of the N source for produced NH3 is helpful in getting an in-depth understanding of the catalyzing mechanism. Thus, isotope-labeling experiments were systematically conducted [34–37]. The 15N2 gas was also sufficiently purified before use. As shown in the ultraviolet-visible (UV–vis) absorption spectra (Fig. 3a), no NH3 can be detected when 15N2 is fed unless an electrocatalytic potential was applied to the CoSC-N-C working electrode. Then, the produced NH3 was distinguished using 1H nuclear magnetic resonance (1H NMR) spectra (Fig. 3b). After a continuous electrolysis under −0.2 V vs. RHE using 15N2 as feeding gas, only a doublet signal (∼73 Hz) representing 15NH4+ is detected in the spectra instead of the triplet signal (∼52 Hz) of 14NH4+. When Ar was used for electrolysis, no NH3 signal was detected in the NMR spectrum, indicating the obtained ammonia is totally from the N2 electroreduction process. For accuracy of the performance data, NMR method was also employed for ammonia quantification (Fig. 3c and Supplementary Fig. 24). The NH3 yield rate and Faradaic efficiency obtained by quantitative 1H NMR and colorimetric method using both 14N2 and 15N2 exhibit good consistency (Fig. 3d), thus demonstrating the reliability of the experimental results.
Figure 3.

(a) The UV-vis absorption spectra of the electrolytes after electrolysis using 15N2 as feeding gas under different conditions. (b) 1H nuclear magnetic resonance (NMR) spectra of both 14NH4+ and 15NH4+ produced from the NRR using 14N2 and 15N2 as feeding gas, respectively. (c) 1H NMR spectra of the NRR product at −0.2 V vs. RHE using NMR quantification method. (d) Comparison of NH3 yield rate and Faradaic efficiency for NRR at −0.2 V vs. RHE using different quantification methods. The error bars correspond to the standard deviations of the obtained data over three separately conducted electrochemical measurements under the same conditions.
Stability is another vital parameter for an electrocatalyst. Thus, the durability of CoSC-N-C was consecutively tested by electrolyzing at a constant potential of −0.2 V vs. RHE for 10 cycles. The CoSC-N-C catalyst can keep the superior Faradaic efficiency of NH3 production unchanged even after 10 cycles of continuous electrolysis (Supplementary Fig. 25), manifesting its broad prospect for practical applications. Simultaneously, TEM image, corresponding element maps and Raman spectra (Supplementary Figs 26 and 27) exhibit no variation in the morphology and structural properties of CoSC-N-C after the NRR process. Its chemical state is also well maintained as confirmed by the high-resolution XPS spectra analyses (Supplementary Fig. 28), demonstrating that the CoSC-N-C is robust enough for long-term NRR electrocatalysis.
Computational studies
Computational studies on both thermodynamics and dynamics were carried out to investigate the mechanism of ammonia synthesis over the CoSC-N-C catalyst. The thermodynamic process of different models was first studied by DFT calculations. As confirmed by catalyst characterizations, the cobalt in CoSC-N-C mainly exists as single clusters. Accordingly, several possible models of cobalt single cluster with different numbers of cobalt atoms on nitrogen-doped carbon (Cox-N4/C, x = 2 to 6) were systematically proposed, and pure nitrogen-doped carbon (N4/C) was also calculated for comparison (Supplementary Fig. 29). For N4/C, the nitrogen adsorption is a strong endothermic process with a step-by-step uphill trend of Gibbs free energy (Supplementary Fig. 30). The introduction of cobalt single clusters is able to turn the nitrogen adsorption into exothermic process without energy barriers, suggesting that the cobalt single clusters are superior nitrogen-adsorption active sites. Notably, all of the structures with cobalt single clusters are able to alter the rate-determining step to the subsequent nitrogen hydrogenation with relatively small energy barriers (Supplementary Figs 31 and 32). Taking Co4-N4/C for example, as clearly shown in Fig. 4a, starting from a favored side-on mode N2 adsorption, the preferred NRR approach of Co4-N4/C is verified to be the associative alternating pathway instead of the associative distal pathway (Supplementary Fig. 33a). When chemically adsorbed on Co cluster, N2 is spontaneously activated and experiences a significant weakening of the N≡N bond due to the strong electron backdonation from the metal to the N2 antibonding orbitals, and the N2 dissociation becomes an exothermic process over the cobalt single cluster. In addition, the energy released from the N2 adsorption step greatly benefits the following N2 cleavage, according to the ‘hot atom’ mechanism [19,38]. Thus, the rate-determining step has been successfully shifted from the usual N2 activation to the subsequent hydrogenation with only a small energy barrier of 0.85 eV. In great contrast, without the Co single cluster, the N4/C model not only suffers a severely endothermic N2 adsorption process, but also possesses high NRR rate-limiting barriers of 2.04 eV and 1.84 eV for the alternating and distal pathways, respectively, indicating a weak activity towards NRR (Fig. 4b and Supplementary Fig. 33b). On the other hand, N2 has priority in the adsorption competition with *H on Co4-N4/C. As shown in Fig. 5a and b, the N2 chemisorption on Co4-N4/C is strongly exothermic (−0.82 eV) with a step-by-step downhill trend of the Gibbs free energy, whereas the H chemisorption suffers an extremely high energy barrier of 2.53 eV due to the water dissociation process. Even if *H is adsorbed, its desorption to form H2 is still an endothermic reaction (Supplementary Fig. 34), which is also beneficial to promoting the NRR Faradaic efficiency.
Figure 4.
Free energy diagram and models represent the corresponding adsorbates on (a) Co4-N4/C and (b) N4/C through the associative alternating pathway. The cyan, red, purple and gray spheres represent C, N, Co and H atoms, respectively.
Figure 5.
(a) Calculated free energy and (b) computational models of the transition state (TS) of hydrogen and nitrogen chemisorption on Co4-N4/C model. The cyan, red, purple, yellow and gray spheres represent C, N, Co, O and H atoms, respectively. (c) PMF for N2 adsorption on the Co4-N4 model in 0.1 M KOH; inset: complete histograms of all window umbrella sampling statistics used for calculation of the N2 adsorption on the Co4-N4 model. (d) Major part of the MD simulation snapshots at 1.50 nm, 0.69 nm and 0.42 nm, with N2 (red spheres), H2O (yellow and gray sticks), OH− (yellow and gray spheres) and K+ (wine red spheres).
From a dynamic point of view, the N2 approaching process towards Co4-N4/C was then explored by advanced sampling MD simulations. The system was set up by randomly placing 6 K+, 6 OH− and 3000 water molecules in the simulation box, with a Co4-N4/C model fixed perpendicular to the z-axis at the center of the simulation system. The N2 molecule initially located at 1.5 nm above the geometric center of the upper three Co atoms. Then, the N2 molecule was pulled towards the Co4-N4/C along the z-axis at a rate of 0.01 nm ps–1 under a harmonic force constant to generate configurations for umbrella-sampling (Fig. 5c). The potential of the mean force (PMF) as a function of distance and the corresponding snapshots (Fig. 5d and Supplementary Fig. 35) illustrate that the approaching process of N2 molecule is nearly barrier-less with only a small energy hill at approximately 0.52 nm. This clearly indicates that the cobalt single cluster as active site is accessible to the N2 molecule, which can greatly promote the following N2 chemisorption, and is thus contributory to the entire NRR process. The above computational conclusions well explain the experimental
results, and further highlight the contribution of the cobalt single cluster as an important promoter for ambient N2 fixation.
CONCLUSION
In summary, we have successfully altered the rate-determining step of ambient NH3 synthesis by deliberate introduction of cobalt single clusters as electron-donating promoter in nitrogen-doped carbon, and achieved outstanding ammonia yield rate of 76.2 μg h–1 mg
and superior Faradaic efficiency of 52.9%. The CoSC-N-C is steady enough for long-term NRR electrocatalysis with negligible decay of the amazingly high Faradaic efficiency through consecutive electrolysis recycle. The 15N isotopic labeling experiments and control experiments are combined to confirm that the N source of N2-to-NH3 conversion is completely from the feeding gas instead of any activated N species in the catalyst. First-principles simulations demonstrate a strong exothermic process of N2 chemisorption on the cobalt single cluster site, which greatly promotes the following N2 dissociation. The N2 activation, therefore, becomes exothermic, and the rate-determining step has been successfully altered to the subsequent nitrogen hydrogenation. A smooth N2 approaching process with almost no energy hill towards the catalyst paves the way for the N2 mass transfer, as confirmed by the MD simulations. Our work overcomes the obstacle of ambient ammonia synthesis, and contributes to the guiding of future catalyst design for sustainable NRR systems.
Supplementary Material
Contributor Information
Sisi Liu, College of Energy, Key Laboratory of Advanced Carbon Materials and Wearable Energy Technologies of Jiangsu Province, Soochow University, Suzhou 215006, China.
Mengfan Wang, College of Energy, Key Laboratory of Advanced Carbon Materials and Wearable Energy Technologies of Jiangsu Province, Soochow University, Suzhou 215006, China.
Haoqing Ji, College of Energy, Key Laboratory of Advanced Carbon Materials and Wearable Energy Technologies of Jiangsu Province, Soochow University, Suzhou 215006, China.
Xiaowei Shen, College of Energy, Key Laboratory of Advanced Carbon Materials and Wearable Energy Technologies of Jiangsu Province, Soochow University, Suzhou 215006, China.
Chenglin Yan, College of Energy, Key Laboratory of Advanced Carbon Materials and Wearable Energy Technologies of Jiangsu Province, Soochow University, Suzhou 215006, China.
Tao Qian, College of Energy, Key Laboratory of Advanced Carbon Materials and Wearable Energy Technologies of Jiangsu Province, Soochow University, Suzhou 215006, China.
FUNDING
This work was supported by the National Natural Science Foundation of China (21703149, 51622208, 51872193 and 5192500409) and the Natural Science Foundation of Jiangsu Province (BK20190827 and BK20181168).
AUTHOR CONTRIBUTIONS
T.Q. conceived and designed this work. C.Y. supervised the project. S.L. prepared materials and performed the material characterizations and electrochemical measurements. S.L. and M.W. participated in analyzing the experimental results. H.J. conducted the theoretical calculations. X.S. offered help in the material characterizations. S.L. wrote this paper with feedback from the other authors. All authors discussed the results and commented on the manuscript.
Conflict of interest statement. None declared.
REFERENCES
- 1. Gruber N, Galloway JN. An earth-system perspective of the global nitrogen cycle. Nature 2018; 451: 293–6. [DOI] [PubMed] [Google Scholar]
- 2. Li X, Li T, Ma Yet al. Boosted electrocatalytic N2 reduction to NH3 by defect-rich MoS2 nanoflower. Adv Energy Mater 2018; 8: 1801357. [Google Scholar]
- 3. Ali M, Zhou F, Chen Ket al. Nanostructured photoelectrochemical solar cell for nitrogen reduction using plasmon-enhanced black silicon. Nat Commun 2016; 7: 11335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. van der Ham CJM, Koper MTM, Hetterscheid DGH. Challenges in reduction of dinitrogen by proton and electron transfer. Chem Soc Rev 2014; 43: 5183–91. [DOI] [PubMed] [Google Scholar]
- 5. Shi M-M, Bao D, Wulan B-Ret al. Au sub-nanoclusters on TiO2 toward highly efficient and selective electrocatalyst for N2 conversion to NH3 at ambient conditions. Adv Mater 2017; 29: 1606550. [DOI] [PubMed] [Google Scholar]
- 6. Licht S, Cui B, Wang Bet al. Ammonia synthesis by N2 and steam electrolysis in molten hydroxide suspensions of nanoscale Fe2O3. Science 2014; 345: 637–40. [DOI] [PubMed] [Google Scholar]
- 7. Wang M, Liu S, Qian Tet al. Over 56.55% Faradaic efficiency of ambient ammonia synthesis enabled by positively shifting the reaction potential. Nat Commun 2019; 10: 341. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Wang L, Xia M, Wang Het al. Greening ammonia toward the solar ammonia refinery. Joule 2018; 2:1055–74. [Google Scholar]
- 9. Wang J, Yu L, Hu Let al. Ambient ammonia synthesis via palladium-catalyzed electrohydrogenation of dinitrogen at low overpotential. Nat Commun 2018; 9: 1795. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Li SJ, Bao D, Shi M-Met al. Amorphizing of Au nanoparticles by CeOx-RGO hybrid support towards highly efficient electrocatalyst for N2 reduction under ambient conditions. Adv Mater 2017; 29: 1700001. [DOI] [PubMed] [Google Scholar]
- 11. Zhang L, Ji X, Ren Xet al. Electrochemical ammonia synthesis via nitrogen reduction reaction on a MoS2 catalyst: theoretical and experimental studies. Adv Mater 2019; 30: 1800191. [DOI] [PubMed] [Google Scholar]
- 12. Seh ZW, Kibsgaard J, Dickens CFet al. Combining theory and experiment in electrocatalysis: insights into materials design. Science 2017; 355: eaad4998. [DOI] [PubMed] [Google Scholar]
- 13. Lv C, Yan C, Chen Get al. An amorphous noble-metal-free electrocatalyst enables N2 fixation under ambient conditions. Angew Chem Int Ed 2018; 57: 6073–6. [DOI] [PubMed] [Google Scholar]
- 14. Liu S, Wang M, Qian Tet al. Facilitating nitrogen accessibility to boron-rich covalent organic frameworks via electrochemical excitation for efficient nitrogen fixation. Nat Commun 2019; 10: 3898. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Chen G-F, Ren S, Zhang Let al. Advances in electrocatalytic N2 reduction-strategies to tackle the selectivity challenge. Small Methods 2019; 3: 1800337. [Google Scholar]
- 16. Qiu W, Xie XY, Qiu Jet al. High-performance artificial nitrogen fixation at ambient conditions using a metal-free electrocatalyst. Nat Commun 2018; 9: 3485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Guo C, Ran J, Vasileff Aet al. Rational design of electrocatalysts and photo(electro)catalysts for nitrogen reduction to ammonia (NH3) under ambient conditions. Energy Environ Sci 2018; 11: 45–56. [Google Scholar]
- 18. Cui X, Tang C, Zhang Q. A review of electrocatalytic reduction of dinitrogen to ammonia under ambient conditions. Adv Energy Mater 2018; 8: 1800369. [Google Scholar]
- 19. Gong Y, Wu J, Kitano Met al. Ternary intermetallic LaCoSi as a catalyst for N2 activation. Nat Catal 2018; 1: 178–85. [Google Scholar]
- 20. Légaré M-A, Bélanger-Chabot G, Dewhurst RDet al. Nitrogen fixation and reduction at boron. Science 2018; 359: 896–900. [DOI] [PubMed] [Google Scholar]
- 21. Ma X-L, Liu J-C, Xiao Het al. Surface single-cluster catalyst for N2-to-NH3 thermal conversion. J Am Chem Soc 2018; 140: 46–9. [DOI] [PubMed] [Google Scholar]
- 22. Kitano M, Inoue Y, Yamazaki Yet al. Ammonia synthesis using a stable electride as an electron donor and reversible hydrogen store. Nat Chem 2012; 4: 934–40. [DOI] [PubMed] [Google Scholar]
- 23. Kitano M, Kanbara S, Inoue Yet al. Electride support boosts nitrogen dissociation over ruthenium catalyst and shifts the bottleneck in ammonia synthesis. Nat Commun 2015; 6: 6731. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Wang P, Chang F, Gao Wet al. Breaking scaling relations to achieve low-temperature ammonia synthesis through LiH-mediated nitrogen transfer and hydrogenation. Nat Chem 2017; 9: 64–70. [DOI] [PubMed] [Google Scholar]
- 25. Inoue Y, Kitano M, Kishida Ket al. Efficient and stable ammonia synthesis by self-organized flat Ru nanoparticles on calcium amide. ACS Catal 2016; 6: 7577–84. [Google Scholar]
- 26. Fei H, Dong J, Arellano-Jiménez Aet al. Atomic cobalt on nitrogen-doped graphene for hydrogen generation. Nat Commun 2015; 6: 8668. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Fu J, Hassan FM, Li Jet al. Flexible rechargeable zinc-air batteries through morphological emulation of human hair array. Adv Mater 2016; 28: 6421–8. [DOI] [PubMed] [Google Scholar]
- 28. Jin H, Wang J, Su Det al. In situ cobalt−cobalt oxide/N-doped carbon hybrids as superior bifunctional electrocatalysts for hydrogen and oxygen evolution. J Am Chem Soc 2015; 137: 2688–94. [DOI] [PubMed] [Google Scholar]
- 29. Guan C, Sumboja A, Zang Wet al. Decorating Co/CoNx nanoparticles in nitrogen-doped carbon nanoarrays for flexible and rechargeable zinc-air batteries. Energy Storage Mater 2019; 16: 243–50. [Google Scholar]
- 30. Bao D, Zhang Q, Meng F-Let al. Electrochemical reduction of N2 under ambient conditions for artificial N2 fixation and renewable energy storage using N2/NH3 cycle. Adv Mater 2017; 29: 1604799. [DOI] [PubMed] [Google Scholar]
- 31. Andersen SZ, Čolić V, Yang Set al. A rigorous electrochemical ammonia synthesis protocol with quantitative isotope measurements. Nature 2019; 570: 504–8. [DOI] [PubMed] [Google Scholar]
- 32. Zhang L, Ding L-X, Chen G-Fet al. Ammonia synthesis under ambient conditions: selective electroreduction of dinitrogen to ammonia on black phosphorus nanosheets. Angew Chem Int Ed 2019; 58: 2612–6. [DOI] [PubMed] [Google Scholar]
- 33. Chen H, Ding L-X, Chen G-Fet al. Molybdenum carbide nanodots enable efficient electrocatalytic nitrogen fixation under ambient conditions. Adv Mater 2018; 30: 1803694. [DOI] [PubMed] [Google Scholar]
- 34. Chen H, Cui P, Wang Fet al. High efficiency electrochemical nitrogen fixation achieved on a low-pressure reaction system by changing chemical equilibrium. Angew Chem Int Ed 2019; 58: 15541–7. [DOI] [PubMed] [Google Scholar]
- 35. Luo Y, Chen G-F, Ding Let al. Efficient electrocatalytic N2 fixation with MXene under ambient conditions. Joule 2019; 3: 279–89. [Google Scholar]
- 36. Chen G-F, Cao X, Wu Set al. Ammonia electrosynthesis with high selectivity under ambient conditions via a Li+ incorporation strategy. J Am Chem Soc 2017; 139: 9771–4. [DOI] [PubMed] [Google Scholar]
- 37. Zhou F, Azofra LM, Ali Met al. Electro-synthesis of ammonia from nitrogen at ambient temperature and pressure in ionic liquids. Energy Environ Sci 2017; 10: 2516–20. [Google Scholar]
- 38. Wang J, Zhang L, Zeng Qet al. Adsorption of atomic and molecular oxygen on 3C-SiC(111) and (111) surfaces: a first-principles study. Phys Rev B 2009; 79: 125304. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




