Next Article in Journal
Promoting Li/MgO Catalyst with Molybdenum Oxide for Oxidative Conversion of n-Hexane
Previous Article in Journal
A Short Review on Ni Based Catalysts and Related Engineering Issues for Methane Steam Reforming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Biomass-Derived Nitrogen-Doped Porous Carbon for Highly Efficient Ambient Electro-Synthesis of NH3

1
CAS Key Laboratory of Bio-Based Materials, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, Qingdao 266101, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
3
National Engineering Lab for Textile Fiber Materials & Processing Technology (Zhejiang) Zhejiang Sci-Tech University, Hangzhou 310018, China
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(3), 353; https://doi.org/10.3390/catal10030353
Submission received: 24 February 2020 / Revised: 15 March 2020 / Accepted: 17 March 2020 / Published: 22 March 2020
(This article belongs to the Section Electrocatalysis)

Abstract

:
In this communication, we report a biomass-derived nitrogen-doped porous carbon (named as NC-800) as an electrocatalyst for the ambient conversion of N2 to NH3. The catalyst NC-800 was prepared from naturally renewable and easily available bamboo shoots, with inherently an approximate 8 wt % of N-containing components, such as the N source, in a cost-effective and environmentally benign manner. This exhibited remarkable catalytic activity with a large NH3 yield and a Faradaic efficiency as high as 16.3 μg hmg-1cat and 27.5%, respectively, at −0.35 V versus a reversible hydrogen electrode (RHE) in 0.1 M HCl solution at ambient conditions. More importantly, the catalyst NC-800 demonstrated excellent electrochemical selectivity and stability.

Graphical Abstract

1. Introduction

Ammonia (NH3), as one of the most important chemicals for industry, has been widely used as a source to synthesize fertilizer, plastics, pharmaceuticals, explosives, and so on [1]. As a high-efficiency and carbon-free energy carrier, with a hydrogen capacity of 17.8 wt % [2], NH3 is also considered as a promising candidate for the future renewable energy economy. The ever-increasing demand for NH3 drives more and more attention toward artificial N2 fixation. However, the current industrial NH3 synthesis still dominantly depends on the energy-intensive Haber–Bosch process, leading to massive fossil fuel consumption and serious greenhouse gas emissions [3,4]. Therefore, it is highly desirable to develop more efficient and energy-saving synthetic routes for NH3 production.
Over the past few decades, great efforts have been devoted to the sustainable synthesis of NH3 from nitrogen, including biomimetic catalysis, photo catalysis, and photo(electro)catalysis [5,6,7]. Apart from the development of synthetic strategies, the detection methods for quantification of the produced ammonia were also developed accordingly [8,9]. Among all of the synthetic routes, electrocatalytic N2 reduction to NH3, ideally operating under ambient conditions with water as the hydrogen source, has been considered as a sustainable and effective way for N2 fixation and NH3 production, especially when it is powered by renewable electricity (e.g., solar and wind) [10]. Consequently, various metal-based and metal-free catalysts have been developed for the synthesis of NH3 under mild conditions via the electrochemical N2 reduction reaction (NRR) [11,12]. However, the efficiency of NH3 production is still far from comparable to that of the Haber–Bosch process, due to the extreme difficulty in breaking the strong N≡N triple bond of N2 and the competing hydrogen evolution reaction (HER) during the electrocatalysis. Among those developed electrocatalysts, metal-free carbon-based catalysts are more intriguing because of their unique features, such as low-costs, large surface areas, mechanical flexibility with good electronic conductivity, and robust tolerance to acidic and/or alkaline conditions [13]. As a consequence, a number of carbon-based materials have been developed for NRR to NH3 production, specifically, heteroatom- (e.g., O [14], N [15], S [16], B [17], F [18], Cl [12], and P [19]) doped carbon materials, which demonstrated highly improved catalytic efficiency. Despite great progress, the current approaches for the introduction of heteroatoms into porous carbon significantly suffer from high-costs and tedious preparation procedures, with the utilization of complicated and expensive starting materials, e.g., metal organic frameworks (MOFs) [20], and metal-ligands complexes [21], which greatly limit their large scale-up for practical applications. Ideally, from both cost-effective and eco-friendly viewpoints, heteroatoms-doped porous carbon materials can be directly prepared from naturally renewable and easily available materials in a facile and environmentally benign manner, and more importantly can enable a highly efficient conversion of N2 to NH3 via the ambient NRR process, which are perfect candidates and will be the ultimate goal for chemists [14,22].
Recently, we have developed a facile and environmentally friendly method for the preparation of N-doped porous carbon with a large surface area, and hierarchical pores from naturally available and renewable biomass–bamboo shoots as the starting material [23]. In this process, the approximate 8 wt % of N-containing compositions, including the proteins and amino acids in bamboo shoots, was effectively decomposed to N atoms and incorporated into carbon lattices via pyrolysis under high temperature. In this work, we report such a biomass-derived N-doped porous carbon can electrochemically convert N2 into NH3 at room temperature and with atmospheric pressure in an acidic aqueous solution. In 0.1 M HCl, the rate of NH3 production and Faradaic efficiency (FE) are as high as 16.3 μg h−1mg−1cat and 27.5%, respectively, at −0.35 V versus a reversible hydrogen electrode (RHE). To the best of our knowledge, these performance metrics are unprecedented and the highest value for the electrocatalytic production of NH3 from N2 under ambient conditions at such a lower potential, compared with those previous results catalyzed by metal-free carbon-based materials [12,24].

2. Results and Discussion

The biomass-derived nitrogen-doped porous carbon was prepared through a facile hydrothermal-pyrolysis process, as shown in Scheme 1a. Typically, fresh bamboo shoots were first cut into small pieces, dried, and ground into powder, followed by the hydrothermal carbonization process in a Teflon-inner stainless-steel autoclave, with deionized water at 180 °C to obtain the biochar, which was then filtered, washed thoroughly with deionized water and dried under vacuum at 60 °C. Finally, the obtained brown biochar was pyrolyzed at 600, 800, and 1000 °C, respectively, under an N2 atmosphere for 4 h at the heating rate of 5 °C min−1. The prepared catalysts were named as NC-T, where T represents the pyrolysis temperature. Note that, toxic or environmentally unfriendly activating agents [25,26], e.g., KOH, ZnCl2, or H3PO4, are not required for carbonization in this process, making this method cost-effective and sustainable.
Comprehensive characterizations, including XRD, Raman, BET, HR-TEM, and XPS, were performed to disclose the physical and chemical features of the NC-800. The TEM image (Scheme 1b) shows that it has a loose and interconnected structure with formation of interstitial porosity and graphitic carbon layers. HR-TEM images (Scheme 1c,d) show the well-resolved lattice spacing of 0.34 nm corresponding to the (002) plane of graphitic carbon. The High-angle annular dark-field scanning TEM (HAADF-STEM) image and elemental mapping (Scheme 1e) indicate that N atoms were homogeneously distributed on graphitic carbon. XRD pattern (Scheme 1f) reveals a graphitic carbon structure with characteristic peaks at 2θ = 23 and 43o, corresponding to (002) and (100) reflection planes, which is consistent with the HR-TEM results. The formation of the graphitic order with a certain degree of defects on the NC-800, induced by high temperature pyrolysis, was further confirmed by Raman spectrum, with observation of two characteristic D (1330 cm−1) and G bands (1590 cm−1) (Scheme 1g) [27]. The XPS for NC exhibits typical C, N, and O peaks (Figures S1, S2, and S3). N 1s XPS spectrum (Scheme 1h) reveals that three types of N species, i.e., pyridinic- (398.4 eV), pyrrolic- (399.7 eV) and graphitic-N (400.9 eV), were embedded into the graphitic carbon structure [28], and the populations of them are 25.7%, 6.7%, and 67.6%, respectively. The pyridinic-N and pyrrolic-N were considered as dangling bonds to activate the N2 molecule [22]. BET analysis (Figure S6) indicates that the NC-800 has a large specific surface area (913.7 m2g−1), and a hierarchical micro-meso-macroporous structure (Table S1). The large specific surface area and the hierarchical structure could lead to the excellent NRR performance for NC-800. The N content in the catalyst NC-800 was estimated to be 3.05 wt% by elemental analysis (Table S1). The large electronegativity of N (3.04) incorporated into the carbon lattices could induce the modification of charge density, electronic structure, and polarization charge, which lead to the strong adsorption of N2 for NC [7].
The ambient NRR performance of NC-800 was conducted in a typical H-cell, separated by a piece of Nafion membrane in an N2-saturated HCl (0.1 M) electrolyte (Figure S8). The catalyst NC-800 was deposited on carbon paper as the working electrode with Ag/AgCl (saturated KCl electrolyte) as the reference in the cathode chamber, and a platinum plate as the counter electrode in the anode chamber. During electrolysis, N2 gas was bubbled into the cathode, where H+ protons transported through the electrolyte could react with N2 adsorbed on the surface of the catalyst, to produce NH3. All potentials are recorded with respect to an RHE scale. The concentrations of the produced NH3 were spectrophotometrically determined using the indophenol blue method and ion chromatography method (Figures S9 and S15) [29], and the possible N2H4 as a by-product was evaluated using the Watt and Chrisp method (Figure S11) [30].
In order to assure the potential range of NRR for the catalyst NC-800, the linear sweep voltammetry (LSV) curves were measured in an Ar- and N2-saturated HCl solution. When the potential is more negative than −0.35 V (vs. RHE), the current density of NC-800 under N2 increases obviously in comparison with that under Ar, demonstrating that electrocatalytic N2 reduction occurs on NC-800 with an onset potential of −0.25 V (Figure 1a). The time-dependent current density curves over NC-800 from −0.25 to −0.50 V were collected over 2 h and are presented in Figure 1b. The corresponding UV/Vis absorption spectra of the electrolytes stained with indophenol blue indicator were displayed in Figure 1c. Both results show that the peak of the highest absorbance intensity is located at −0.35 V. The corresponding FEs and NH3 yields at different potentials are plotted in Figure 1d. Both FEs and NH3 yields increase at negatively applied potentials until −0.35 V, and then decrease gradually, which is ascribed to the competitive selectivity toward the hydrogen evolution reaction (HER) [12,31,32]. The highest FE (27.5%) and maximum NH3 yield of 16.3 μg h−1mg−1cat are achieved at −0.35 V vs. RHE, which is comparable and even superior to those of mostly reported results catalyzed by metal-free carbon or even metal-based catalysts. A more detailed comparison is compiled in Table S3. To ensure the reliability and accuracy of the obtained catalytic activity, the ion chromatography method was also adopted for NH3 quantification, and a NH3 yield (15.13 μg h−1mg−1cat) very close to that of the indophenol blue method was observed (Figure 1e). Moreover, N2H4 was not detected at all potentials during the NRR process (Figure S12), indicative of a 100% selectivity of NC-800 for the electrocatalytic N2 reduction to NH3. In addition, for comparison, NC-600 and NC-1000 were also used as the electrocatalysts to assess their catalytic performance at −0.35 V under identical conditions, and it turned out that NC-800 gave the best FEs and NH3 yield (Figure S13). The outstanding electrochemical NRR performance of NC-800 can be attributed to the following two reasons. Firstly, the large specific surface area makes active sites more accessible for electrochemical N2-NH3, while the abundance of micropores in the NC-800 could accelerate the electron and mass transfer process (Figure S6). Secondly, the presence of pyridinic-N on carbon could function as the active sites to facilitate the activation of N2 molecule. Therefore, NC-800 was used as the ideal catalyst for further investigation.
To verify that the detected NH3 was generated through the NRR process over NC-800, the electrolysis was carried out in either N2-saturated solution at open circuit potential or Ar-saturated solution at −0.35 V (vs. RHE) for 2 h [33]. In either case, almost no NH3 was detected (Figure 2a), strongly suggesting that the NH3 was produced exclusively from the NRR reaction catalyzed by NC-800 electrode. Furthermore, a linear correlation between mNH3 and the reaction time of electrolysis was also observed, as shown in Figure 2b. To further confirm the N source of the NH3 product during NRR process, we also performed the 15N isotopic measurements using 15N2 as the feeding gas. As shown in Figure 2e, there is a distinguishable chemical shift of a doublet pattern with a coupling constant of 1JN-H = 72 Hz for 15NH4+ NRR samples, indicating that ammonia is produced via electrochemical nitrogen reduction. In contrast, a triplet coupling with the 1JN-H = 52 Hz for 14NH4+ samples are observed when using 14N2 as the feeding gas and there is no detectable amount of ammonia in the Ar-feeding electrolyte, which further confirms that the N source of ammonia is N2 gas. Of note, to exclude any contamination of NH3 production originating from the surroundings, nitrogen and argon gas were pre-purified by passing a glass vessel containing 0.1 M NaOH (50 mL) solution before feeding into the cathode. In addition, detection of any ammonia formation from the electrolyte was also analyzed before each NRR test. Besides, the electrolytes of the tail gas trap combined with the anode chamber were also detected, and the results show that the ammonia production was negligible. Taking all results into account, we can safely conclude that NH3 did originate from the NRR catalyzed by NC-800.
The stability of an electrocatalyst is a key factor for its practical application. We subsequently investigated the electrolytic performance of the catalyst NC-800 for successive uses. As presented in Figure 2d, the NH3 yields and FEs show a slight variation after five consecutive recycling tests, indicating the high stability of NC during the N2 reduction process. Meanwhile, small attenuation in the current density of the chronoamperometric measurements was also observed after around 12 h of NRR electrolysis at −0.35 V (Figure 2c). In addition, the used catalyst NC-800 also demonstrated NH3 yield (16.28 μg h−1mg−1cat) and FEs (24.43%) for NRR electrolysis under identical conditions, which is equal to that obtained from the fresh NC-800 (Table S2). All of these results further confirm the long-term electrochemical durability of the catalyst NC-800.
Previous reports regarding the NRR process electrocatalyzed by N-doped carbon as the catalyst reveal that pyridinic and pyrrolic N moieties incorporated in a graphitic layer, together with the generated adjacent atomic vacancy (e.g., C and N), are active sites for adsorbing and activating the N2 molecule to initiate the conversion of N2 to NH3 [13,22]. Thus, we consider the present catalyst similar to those previously reported catalysts in catalytically active sites and a preferable electrochemical reaction pathway for NH3 synthesis on NC-800. To further verify this, density functional theory (DFT) calculations were conducted (Scheme S1 and Figure S17). Our theoretical calculations reveal that two adjacent pyridinic-N vacancy species in the graphitic carbon structure can form a strong interaction with the molecule N2 in a very stable structure, which can be hydrogenated by the proton H+, followed by the process N-N→N-NH→HN-NH→HN-NH2→HN-NH3 to generate NH3 molecule and simultaneously create the N vacancies for further N2 adsorption and activation. Such a finding is in line with previous reports [13,22,34].

3. Conclusions

In summary, we report a hierarchically porous N-doped carbon (NC-800), derived from N-containing bamboo shoots, as a catalyst for the ambient electrochemical conversion of N2 to NH3. A large NH3 yield and a Faradaic efficiency as high as 16.3 μgh−1mg−1cat and 27.5%, respectively, were achieved at −0.35 V versus a reversible hydrogen electrode (RHE) in 0.1 M HCl solution at ambient conditions. More importantly, the NC-800 also exhibited excellent electrochemical selectivity and stability. Our work opens a new avenue to develop various metal-based catalysts for electrocatalysis using the hierarchically porous N-doped carbon.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/3/353/s1, Figure S1. N 1s XPS spectra for the catalysts NC-600, 800, 1000. Figure S2. O 1s XPS spectra for the catalysts NC-600, 800, 1000. Figure S3. C 1s XPS spectra for the catalysts NC-600, 800, 1000. Figure S4. Raman spectra for the catalysts NC-600, 800, 1000. Figure S5. XRD pattern for the catalysts NC-600, 800, 1000. Figure S6. N2 adsorption-desorption isotherms and the pore size distributions of the prepared catalysts. Figure S7. N2-TPD profiles of the NC-600, NC-800 and NC-1000 catalysts. Figure S8. The photograph of H-cell with three-electrode cell and the working electrode connections of NRR process. Figure S9. (a) UV-vis curves of indophenol assays with NH4+ ions after incubated for 2 h at HCl solution. (b) Concentration-absorbance curve of NH4+ ions with a serious of standard concentration at HCl solutions. Figure S10. UV-vis absorption spectra of the electrolytes stained with an indophenol indicator after NRR electrolysis. Figure S11. (a) UV-vis curves of Watt and Chrisp assays with N2H4 after incubated for 2 h at HCl solution (b) Concentration-absorbance curve of N2H4 solution with a serious of standard concentration at HCl solution. Figure S12. Determination of the formation of by-product N2H4 over the NC-800 at different potential (vs. RHE). Figure S13. The NH3 yield and FEs over the catalysts NC-600, 800 and 1000 at −0.35 eV (vs. RHE). Figure S14. The NH3 yield of the cathode chamber, tail gas trap, and anode chamber over the catalysts NC-800 at −0.35 V. Figure S15. (a) Ion chromatography of a suit of NH4+ at HCl solution. (b) Concentration-area curve of NH4+ ions with a serious of standard concentration at HCl solutions. Figure S16. UV-vis absorption spectra of the electrolytes stained with an indophenol indicator after NRR electrolysis. Figure S17. The Mars-van-Krevelen mechanism for electrochemical nitrogen reduction reaction to ammonia on NC-800. The free energy profiles and the structures of intermediates are shown in the reaction path. Scheme S1. a) Side view, and b) top view of the optimized pyridinic-N doped graphitic porous carbon structures along the hydrogenation process. Table S1. Chemical composition and texture properties of the catalyst NC-T. Table S2. Chemical composition and NH3 yield and FEs of the catalyst fresh- and used-NC-800.

Author Contributions

Conceptualization, Q.L., X.C. and Y.Y.; Investigation, Q.L.; Methodology, Q.L.; Project administration, X.C. and Y.Y.; Supervision, Y.Y.; Writing-original draft, Q.L. and Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to acknowledge the financial support from the Key Technology R&D Program of Shandong Province (2019GGX102075), Open Projects of State Key Laboratory of Physical Chemistry of the Solid Surface (Xiamen University) (No. 201808), and the 13th-Five Key Project of the Chinese Academy of Sciences (Grant No. Y7720519KL). Y.Y also acknowledges the support from the Royal Society (UK) for a Newton Advanced Fellowship (NAF\R2\180695).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Robert, F.S. Chemistry. New recipe produces ammonia from air, water, and sunlight. Science 2014, 345, 610. [Google Scholar]
  2. Wang, W.T.; Herreros, J.M.; Tsolakis, A.; York, A.P.E. Ammonia as hydrogen carrier for transportation; investigation of the ammonia exhaust gas fuel reforming. Int. J. Hydrogen Energy 2013, 38, 9907–9917. [Google Scholar] [CrossRef] [Green Version]
  3. Van der Ham, C.J.M.; Koper, M.T.M.; Hetterscheid, D.G.H. Challenges in reduction of dinitrogen by proton and electron transfer. Chem. Soc. Rev. 2014, 43, 5183–5191. [Google Scholar] [CrossRef] [PubMed]
  4. Kandemir, T.; Schuster, M.E.; Senyshyn, A.; Behrens, M.; Schlögl, R. The Haber–Bosch process revisited: On the real structure and stability of “ammonia iron” under working conditions. Angew. Chem. Int. Ed. 2013, 52, 12723–12726. [Google Scholar] [CrossRef]
  5. Broda, H.; Tuczek, F. Catalytic ammonia synthesis in homogeneous solution—Biomimetic at last? Angew. Chem. Int. Ed. 2014, 53, 632–634. [Google Scholar] [CrossRef]
  6. Chen, X.Z.; Li, N.; Kong, Z.Z.; Ong, W.J.; Zhao, X.J. Photocatalytic fixation of nitrogen to ammonia: State-of-the-art advancements and future prospects. Mater. Horiz. 2018, 5, 9–27. [Google Scholar] [CrossRef]
  7. Chen, X.R.; Guo, Y.T.; Du, X.C.; Zeng, Y.S.; Chu, J.W.; Gong, C.H.; Huang, J.W.; Fan, C.; Wang, X.F.; Xiong, J. Atomic structure modification for electrochemical nitrogen reduction to ammonia. Adv. Energy Mater. 2019, 10, 1903172. [Google Scholar] [CrossRef]
  8. Zhao, Y.X.; Shi, R.; Bian, X.A.; Zhou, C.; Zhao, Y.F.; Zhang, S.; Wu, F.; Waterhouse, G.I.N.; Wu, L.Z.; Tung, C.H.; et al. Ammonia detection methods in photocatalytic and electrocatalytic experiments: How to improve the reliability of NH3 production rates? Adv. Sci. 2019, 6, 1802109. [Google Scholar] [CrossRef] [Green Version]
  9. Andersen, S.Z.; Čolić, V.; Yang, S.G.; Schwalbe, J.A.; Nielander, A.C.; MacEnaney, J.M.; Rasmussen, K.E.; Baker, J.G.; Singh, A.R.; Rohr, B.A.; et al. A rigorous electrochemical ammonia synthesis protocol with quantitative isotope measurements. Nature 2019, 570, 504–508. [Google Scholar] [CrossRef] [Green Version]
  10. Suryanto, B.H.R.; Du, H.L.; Wang, D.B.; Chen, J.; Simonov, A.N.; MacFarlane, D.R. Challenges and prospects in the catalysis of electroreduction of nitrogen to ammonia. Nat. Catal. 2019, 2, 290–296. [Google Scholar] [CrossRef] [Green Version]
  11. Xu, B.; Liu, Z.C.; Qiu, W.B.; Liu, Q.; Sun, X.P.; Cui, G.W.; Wu, Y.P.; Xiong, X.L. La2O3 nanoplate: An efficient electrocatalyst for artificial N2 fixation to NH3 with excellent selectivity at ambient condition. Electrochim. Acta 2019, 298, 106–111. [Google Scholar] [CrossRef]
  12. Zang, W.J.; Yang, T.; Zou, H.Y.; Xi, S.B.; Zhang, H.; Liu, X.M.; Kou, Z.K.; Du, Y.H.; Feng, Y.P.; Shen, L.; et al. Copper single atoms anchored in porous nitrogen-doped carbon as efficient pH-Universal catalysts for the nitrogen reduction reaction. ACS Catal. 2019, 9, 10166–10173. [Google Scholar] [CrossRef]
  13. Zhao, S.L.; Lu, X.Y.; Wang, L.Z.; Gale, J.L.; Amal, R. Carbon-based metal-free catalysts for electrocatalytic reduction of nitrogen for synthesis of ammonia at ambient conditions. Adv. Mater. 2019, 31, 1805367. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, T.T.; Li, P.P.; Wang, H.B.; Zhao, R.B.; Zhou, Q.; Kong, W.H.; Liu, M.L.; Zhang, Y.Y.; Sun, X.P.; Gong, F.F. Biomass-derived oxygen-doped hollow carbon microtubes for electrocatalytic N2-to-NH3 fixation under ambient conditions. Chem. Commun. 2019, 55, 2684–2687. [Google Scholar] [CrossRef]
  15. Kong, Y.; Li, Y.; Yang, B.; Li, Z.J.; Yao, Y.; Lu, J.G.; Lei, L.C.; Wen, Z.H.; Shao, M.H.; Hou, Y. Boron and nitrogen co-doped porous carbon nanofibers as metal-free electrocatalysts for highly efficient ammonia electrosynthesis. J. Mater. Chem. A 2019, 7, 26272–26278. [Google Scholar] [CrossRef]
  16. Xia, L.; Yang, J.J.; Wang, H.B.; Zhao, R.B.; Chen, H.Y.; Fang, W.H.; Asiri, A.M.; Xie, F.Y.; Cui, G.L.; Sun, X.P. Sulfur-doped graphene for efficient electrocatalytic N2-to-NH3 fixation. Chem. Commun. 2019, 55, 3371–3374. [Google Scholar] [CrossRef]
  17. Yu, X.M.; Han, P.; Wei, Z.Z.; Huang, L.S.; Gu, Z.X.; Peng, S.J.; Ma, J.M.; Zheng, G.F. Boron-doped graphene for electrocatalytic N2 reduction. Joule 2018, 2, 1610–1622. [Google Scholar] [CrossRef] [Green Version]
  18. Yuan, D.; Wei, Z.X.; Han, P.; Yang, C.; Huang, L.S.; Gu, Z.X.; Ding, Y.; Ma, J.M.; Zheng, G.F. Electron distribution tuning of fluorine-doped carbon for ammonia electrosynthesis. J. Mater. Chem. A 2019, 7, 16979–16983. [Google Scholar] [CrossRef]
  19. Song, P.F.; Wang, H.; Kang, L.; Rang, B.C.; Song, H.H.; Wang, R.M. Electrochemical nitrogen reduction to ammonia at ambient conditions on nitrogen and phosphorus co-doped porous carbon. Chem. Commun. 2019, 55, 687–690. [Google Scholar] [CrossRef]
  20. Wang, T.T.; Zhao, Q.D.; Fu, Y.Y.; Lei, C.J.; Yang, B.; Li, Z.J.; Lei, L.C.; Wu, G.; Hou, Y. Carbon-Rich nonprecious metal single atom electrocatalysts for CO2 reduction and hydrogen evolution. Small Methods 2019, 3, 1900210. [Google Scholar] [CrossRef]
  21. Signarvic, R.S.; Degrado, W.F. Metal-binding dependent disruption of membranes by designed helices. J. Am. Chem. Soc. 2009, 131, 3377–3384. [Google Scholar] [CrossRef] [Green Version]
  22. Zhao, C.J.; Zhang, S.B.; Hu, M.M.; Zhang, X.; Liu, Y.Y.; Li, W.Y.; Chen, C.; Wang, G.Z.; Zhang, H.M.; Zhao, H.J. Ambient electrosynthesis of ammonia on a biomass-derived nitrogen-doped porous carbon electrocatalyst: Contribution of pyridinic nitrogen. ACS Energy Lett. 2019, 4, 377–383. [Google Scholar] [CrossRef]
  23. Ji, G.J.; Duan, Y.N.; Zhang, S.C.; Fei, B.H.; Chen, X.F.; Yang, Y. Selective semihydrogenation of alkynes catalyzed by Pd nanoparticles immobilized on heteroatom-doped hierarchical porous carbon derived from bamboo shoots. ChemSusChem 2017, 10, 3427–3434. [Google Scholar] [CrossRef]
  24. Chen, C.; Yang, D.F.; Wang, Y.; Zhou, Y.Y.; Zou, Y.Q.; Li, Y.F.; Wang, S.Y. B-N Pairs enriched defective carbon nanosheets for ammonia synthesis with high efficiency. Small 2019, 15, 1805029. [Google Scholar] [CrossRef]
  25. Deng, Q.F.; Liu, L.; Lin, X.Z.; Du, G.; Liu, Y.; Yuan, Z.Y. Synthesis and CO2 capture properties of mesoporous carbon nitride materials. Chem. Eng. J. 2012, 203, 63–70. [Google Scholar] [CrossRef]
  26. Fan, X.; Zhang, L.; Zhang, G.; Shu, Z.; Shi, J. Chitosan derived nitrogen-doped microporous carbons for high performance CO2 capture. Carbon 2013, 61, 423–430. [Google Scholar] [CrossRef]
  27. Song, T.; Ren, P.; Duan, Y.N.; Wang, Z.Z.; Chen, X.F.; Yang, Y. Cobalt nanocomposites on N-doped hierarchical porous carbon for highly selective formation of anilines and imines from nitroarenes. Green Chem. 2018, 20, 4629–4637. [Google Scholar] [CrossRef]
  28. Wang, X.Y.; Zhao, Q.D.; Yang, B.; Li, Z.J.; Bo, Z.; Lam, K.H.; Adli, N.M.; Lei, L.C.; Wen, Z.H.; Wu, G.; et al. Emerging nanostructured carbon-based non-precious metal electrocatalysts for selective electrochemical CO2 reduction to CO. J. Mater. Chem. A 2019, 7, 25191–25202. [Google Scholar] [CrossRef]
  29. Shi, M.M.; Bao, D.; Wulan, B.R.; Li, Y.H.; Zhang, Y.F.; Yan, J.M.; Jiang, Q. Au Sub-Nanoclusters on TiO2 toward highly efficient and selective electrocatalyst for N2 conversion to NH3 at ambient conditions. Adv. Mater. 2017, 29, 1606550. [Google Scholar] [CrossRef]
  30. Watt, G.W.; Chrisp, J.D. Spectrophotometric method for determination of hydrazine. Anal. Chem. 1952, 24, 2006–2008. [Google Scholar] [CrossRef]
  31. Zhou, F.L.; Azofra, L.M.; Ali, M.; Kar, M.; Simonov, A.N.; Mcdonnell-Worth, C.; Sun, C.H.; Zhang, X.Y.; Macfarlane, D.R. Electro-synthesis of ammonia from nitrogen at ambient temperature and pressure in ionic liquids. Energy Environ. Sci. 2017, 10, 2516–2520. [Google Scholar] [CrossRef]
  32. Li, Y.; Zhang, Q.; Li, C.; Fan, H.N.; Luo, Q.B.; Liu, H.K.; Dou, S.X. Atomically dispersed metal dimer species with selective catalytic activity for nitrogen electrochemical reduction. J. Mater. Chem. A 2019, 7, 22242–22247. [Google Scholar] [CrossRef]
  33. Tang, C.; Qiao, S.Z. How to explore ambient electrocatalytic nitrogen reduction reliably and insightfully. Chem. Soc. Rev. 2019, 48, 3166–3180. [Google Scholar] [CrossRef]
  34. Wang, Z.H.; Hu, X.; Liu, Z.Z.; Zou, G.J.; Wang, G.N.; Zhang, K. Recent developments in polymeric carbon nitride-derived photocatalysts and electrocatalysts for nitrogen fixation. ACS Catal. 2019, 9, 10260–10278. [Google Scholar] [CrossRef]
Scheme 1. (a) Schematic illustration of the fabrication process of bamboo shoots-derived NC-800. (b), (c), and (d) HRTEM images of the catalyst NC-800. (e) HAADF-HRTEM image of the NC-800 and corresponding elemental mapping of the NC-800. XRD pattern (f), Raman spectra (g), and N 1s XPS spectra (h) of the catalyst NC-800.
Scheme 1. (a) Schematic illustration of the fabrication process of bamboo shoots-derived NC-800. (b), (c), and (d) HRTEM images of the catalyst NC-800. (e) HAADF-HRTEM image of the NC-800 and corresponding elemental mapping of the NC-800. XRD pattern (f), Raman spectra (g), and N 1s XPS spectra (h) of the catalyst NC-800.
Catalysts 10 00353 sch001
Figure 1. (a) Linear sweep voltammetry (LSV) curves for the catalyst NC-800 in Ar- and N2-saturated 0.1 M HCl. (b) Time-dependent current density curves of NC at a suit of potentials for 2 h in N2-saturated 0.1 M HCl. (c) UV-Vis absorption spectra of the electrolyte colored with indophenol indicator for 2 h. (d) The obtained NH3 yield rates and Faradaic efficiency (FEs) for NC-800 at different potentials. (e) The obtained NH3 yield rates for NC under various potentials determined by ion chromatography.
Figure 1. (a) Linear sweep voltammetry (LSV) curves for the catalyst NC-800 in Ar- and N2-saturated 0.1 M HCl. (b) Time-dependent current density curves of NC at a suit of potentials for 2 h in N2-saturated 0.1 M HCl. (c) UV-Vis absorption spectra of the electrolyte colored with indophenol indicator for 2 h. (d) The obtained NH3 yield rates and Faradaic efficiency (FEs) for NC-800 at different potentials. (e) The obtained NH3 yield rates for NC under various potentials determined by ion chromatography.
Catalysts 10 00353 g001
Figure 2. (a) The NH3 production for NC under various conditions: at −0.35 V in N2, open circuit in N2, and −0.35 V in Ar for 2 h. (b) The curve of NH3 production versus reaction time at −0.35 V. (c) The time-dependent current density curve for NC at −0.35 V in N2 for 12 h (inset, NH3 yield rates and FEs for NC before and after 12 h at −0.35 V). (d) The NH3 yield rates and FEs for NC at −0.35 V over 5 cycles. (e) 1H NMR spectra of the electrolytes after NRR electrolysis on NC for 2 h using Ar gas, 14N2 gas, 15N2 gas as the feeding gas.
Figure 2. (a) The NH3 production for NC under various conditions: at −0.35 V in N2, open circuit in N2, and −0.35 V in Ar for 2 h. (b) The curve of NH3 production versus reaction time at −0.35 V. (c) The time-dependent current density curve for NC at −0.35 V in N2 for 12 h (inset, NH3 yield rates and FEs for NC before and after 12 h at −0.35 V). (d) The NH3 yield rates and FEs for NC at −0.35 V over 5 cycles. (e) 1H NMR spectra of the electrolytes after NRR electrolysis on NC for 2 h using Ar gas, 14N2 gas, 15N2 gas as the feeding gas.
Catalysts 10 00353 g002

Share and Cite

MDPI and ACS Style

Li, Q.; Chen, X.; Yang, Y. Biomass-Derived Nitrogen-Doped Porous Carbon for Highly Efficient Ambient Electro-Synthesis of NH3. Catalysts 2020, 10, 353. https://doi.org/10.3390/catal10030353

AMA Style

Li Q, Chen X, Yang Y. Biomass-Derived Nitrogen-Doped Porous Carbon for Highly Efficient Ambient Electro-Synthesis of NH3. Catalysts. 2020; 10(3):353. https://doi.org/10.3390/catal10030353

Chicago/Turabian Style

Li, Qinglin, Xiufang Chen, and Yong Yang. 2020. "Biomass-Derived Nitrogen-Doped Porous Carbon for Highly Efficient Ambient Electro-Synthesis of NH3" Catalysts 10, no. 3: 353. https://doi.org/10.3390/catal10030353

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop