Wuyong
Zhang
,
Yingjie
Wen
,
Haocheng
Chen
,
Minli
Wang
,
Caihan
Zhu
,
Yunan
Wang
* and
Zhiyi
Lu
*
Key Laboratory of Advanced Fuel Cells and Electrolyzers Technology of Zhejiang Province, Chinese Academy of Sciences, Ningbo, Zhejiang 315201, P. R. China. E-mail: luzhiyi@nimte.ac.cn; wangyunan@nimte.ac.cn
First published on 22nd June 2024
Developing high-efficiency electrocatalysts for nitrate-to-ammonia transformation holds significant promise for the production of ammonia, a crucial component in agricultural fertilizers and as a carbon-free energy carrier. In this study, we propose a viable strategy involving sulfur doping to modulate both the microstructure and electronic properties of CoSe2 for nitrate reduction. This approach remarkably enhances the conversion of nitrate to ammonia by effectively regulating the adsorption capability of nitrogenous intermediates. Specifically, sulfur-doped CoSe2 nanowires (S-CoSe2 NWs) exhibit a peak faradaic efficiency of 93.1% at −0.6 V vs. RHE and achieve the highest NH3 yield rate of 11.6 mg h−1 cm−2. Mechanistic investigations reveal that sulfur doping facilitates the creation of highly charged active sites, which enhance the adsorption of nitrite and subsequent hydrogenation, leading to improved selectivity towards ammonia production.
New conceptsIn this study, we demonstrate a new concept that regulates the nanodimensions of transition-metal dichalcogenides (TMDs) by introducing sulfur. Our sulfur doping method differs from the normal electronic regulation of materials and further modulates the morphology of the CoSe2 materials from 2D to 1D by changing the supersaturation and ionization of the synthetic system. The 1D ternary S-CoSe2 NWs can drive nitrate electroreduction to ammonia (NARR) with high activity and selectivity due to their advanced one-dimensional structure and high-charge active center. This is the first time that a morphology and electron-control method has been proposed to provide an advanced structure for NARR in TMDs. The significance of an advanced microstructure and electronic structure in NARR has been proven, and we are sure this is a highly innovative approach from the perspective of materials chemistry and from electrochemistry, which opens up an interesting field of research. |
The primary obstacle to advancing NRR lies in the inertness of nitrogen (941 kJ mol−1 for NN) and its low water solubility (20 mg L−1).15 In contrast, nitrate (NO3−) boasts a more readily activated chemical bond (204 kJ mol−1 for N–O) and greater solubility, facilitating nitrate reduction reactions (NARRs) at the solid–liquid interface with a lower energy barrier.16 Moreover, nitrate is abundant in the Earth's environment and is extensively used as a nitrogen-containing fertilizer.17 Also, precisely because of this, its widespread use has led to pervasive nitrogen contamination, significantly impacting biodiversity.18 Additionally, many individuals are exposed to water environments exceeding regulated nitrate limits (50 mg L−1, WHO, 2004), potentially leading to health issues such as non-Hodgkin's lymphoma and methemoglobinemia.19–21 Therefore, reversing nitrate conversion is of substantial importance for both industrial ammonia production and environmental remediation.22
In addition to the aforementioned merits, the NARR process generates undesirable byproducts such as dinitrogen, nitrite, and nitric oxide due to multiple electron–proton transfer processes.23–25 The formation of these byproducts depends on the nitrate concentration and electrolyte pH, highlighting the need for novel materials with abundant active sites to facilitate efficient ammonia production.26–28 While traditional efficient NARR processes often rely on noble metals like Ru-based materials, their high cost and natural scarcity pose significant barriers to widespread applications.29–32 In contrast, transition-metal dichalcogenides (TMDs) such as CoSe2 and CoS2 have emerged as compelling candidates due to their abundance and cost-efficient production processes.33–35 Moreover, given that electrochemical reactions predominantly occur at material surfaces or interfaces, the high specific surface area of TMDs, akin to MoS2 for hydrogen evolution, holds promise for electrochemical energy conversion.36–38 However, the common 2D TMDs suffer from aggregation or restacking that severely limits their potential performance, necessitating the construction of advanced structures from the molecular level to enhance the intrinsic activity of TMDs.39,40 Furthermore, the application of TMDs in the NARR has been scarcely reported to date, highlighting the untapped potential in this domain.41,42
Herein, we present a novel ternary TMD, S-CoSe2 nanowire, designed and synthesized for the NARR. By introducing sulfur to modulate the ionization and supersaturation of the synthetic system, the microstructure of the original binary CoSe2 transitions from 2D to 1D, coinciding with the generation of high-charge cobalt centers. The resulting 1D S-CoSe2 nanowire with an advanced structure exhibits pronounced nitrate-to-ammonia activity and selectivity, achieving a peak faradaic efficiency of 93.1% and the highest NH3 yield rate of 11.6 mg h−1 cm−2. Further theoretical studies suggest that the high-charge cobalt center enhances the adsorption of NARR intermediates, such as nitrite, for further hydrogenation.
The fabrication process of the S-CoSe2 NWs was also further investigated. First, the formation of the S-CoSe2 NWs highly relies on the reaction temperature and time. Sufficient temperature is required to generate enough solvent pressure to drive the formation of the desired morphology. By regulating the reaction temperature from 120 °C to 190 °C, it was observed that temperatures below 170 °C resulted in the formation of unexpected bulk materials, and the materials obtained individual fine nanowires at temperatures over 170 °C (Fig. S7, ESI†). With an increase of the reaction time, the material is gradually stretched to obtain a flexible and smooth nanowire morphology (Fig. S8, ESI†). Furthermore, it was found that the addition of sulfur was pivotal to forming a unique nanowire morphology. When sulfur is absent in the synthesis process, a totally different nanosheet-like material (CoSe2 NSs) can be obtained (Fig. S9, ESI†). With the regulation of sulfur usage, it can be clearly observed that the material has an obvious evolutionary morphology change from nanosheets to nanowires (Fig. 2b–f and Fig. S10, ESI†). Meanwhile, the XRD patterns of CoSe2 also broadened and weakened, indicating that the crystallinity of the material was poor. The slightly negative shift of the XRD peaks indicates the lattice expansion of the as-prepared materials with sulfur doping (Fig. S11, ESI†). This intriguing phenomenon drives us to elucidate the intrinsic formation mechanism of S-CoSe NWs. Ethylenediamine (EDA) was employed as the solvent and sequestrant in this system. Co2+, Se, S can coordinate with EDA to form a series of complexes [Co(EDA)2]2+, [Sex-EDA]− and [Sx-EDA]−, respectively.48 The addition of sulfur can replace the counterion for [Co(EDA)2]2+, which results in a blue shift in the UV-vis spectra (Fig. S12, ESI†). The distinct precursor complex also has a different supersaturation (S) of the solution and electrolytic dissociation (α) of the reactants.49S can influence the surface engineering of the resulting materials, and α can determine whether the growth of materials is isotropic or anisotropic.50,51 There is no doubt that EDA is a weak electrolyte with low α (10−4.07), which results in the anisotropic growth of CoSe2 to be 2D nanosheets or a 1D nanowires, as the EDA can provide a constant and equilibrium-driven force to support the orientated growth. Moreover, the new counterion of [Sx-EDA]− in the system significantly decreases the S of the final products, which further results in a small, saturated region with a slow growth rate to obtain regularly shaped S-CoSe2 NWs (Fig. 2a). Besides the morphology regulation, sulfur doping also affects the electronic states of the resulting materials. With more electronegative sulfur doping, the oxidation state of cobalt in the S-CoSe2 NWs is significantly different from that of CoSe2 NSs, which can be observed from the positive binding energy shift from Co 2p X-ray photoelectron spectroscopy (Fig. S13, ESI†). The local structure and coordination information of the CoSe2 materials were further studied by X-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) spectroscopy. The main adsorption Co K-edge XANES peak shifts to higher energy after sulfur doping, which further consolidates the higher oxidation state of the S-CoSe2 NWs (Fig. 2g). Moreover, CoSe2 NSs and S-CoSe2 NWs both show a pre-edge peak at around 7709.8 eV for the 1s to eg intra-atomic transition in cobalt, which indicates the sulfur doping has a negligible influence on the octahedral configuration of CoSe2.52 However, it is worth mentioning that the intensity of the pre-edge peak has a perceptible decrease with sulfur incorporating, indicating the enhancive eg filling of cobalt (inset in Fig. 2g).52 Further extended X-ray absorption fine structure (EXAFS, Fig. 2h) spectroscopy discloses an obvious decrease of the Co–Se scattering path due to the shorter bond distance of Co–S than that of Co–Se, and the emergent Co–S scattering path in S-CoSe2 NWs is clearly observed in wavelet transform analysis with the high-k direction extension (Fig. 2i and j). The corresponding EXAFS fitting also implies a monotonic decrease of the coordination number (CN) of Co–Se and increase of the CN of Co–S (Fig. S14 and Table S2, ESI†). Generally, the mechanistic investigation and spectroscopic studies on CoSe2 materials reveal that sulfur doping can readily regulate not only the microstructure from 2D to 1D but also controllably tailor the electronic structure and local chemical environment of CoSe2.
Based on the above analysis, CoSe2 NSs and S-CoSe2 NWs with distinct microstructures and electronic structures could serve as promising research platforms to explore the structure–property relationship of nitrate reduction in TMDs, and further unveil the fundamental mechanisms involved in the NARR. As doping strategy can effectively regulate the catalytic activity, and sulfur can combine with hydrogen with the p-orbital to form active hydrogen, which is considered an important species to combine with NARR intermediates to modulate the reaction behavior.53–57 Based on these considerations, the self-supported working electrode of the in situ growth S-CoSe2 NWs on 1 × 1 cm2 carbon cloth (S-CoSe2 NWs@CC, Fig. S15, ESI†) was employed for the NARR test, and CoSe2 NSs@CC was used as a reference. The electrodes were further dwelled in an H-cell filled with 0.5 M Na2SO4 and 0.5 M KNO3 as electrolytes. Linear sweep voltammetry (LSV) was initially conducted on the utilized catalysts, while both CoSe2 NSs@CC and S-CoSe2 NWs@CC exhibited pronounced current responses in the presence of NO3−. However, S-CoSe2 NWs@CC had a more positive onset potential and higher current response (Fig. 3a), which can be attributed to the higher intrinsic activity and advanced porous structure with a higher BET surface area(Fig. S16, ESI†).58 The ammonia concentration of the electrolytes for the above catalysts was investigated after chronoamperometry at −0.5 V vs. RHE using the indophenol blue method. Apparently, S-CoSe2 NWs@CC can produce more ammonia under the same potential compared to CoSe2 NSs@CC and bare CC (Fig. 3b), as indicated by the sharp peak of indophenol blue centered at 655.5 nm in the UV spectra. The NH3 yield and the corresponding FE of the electrodes at all given potentials are further calculated from the standard curve recorded at different concentrations of NH4+–N (Fig. S17, ESI†). S-CoSe2 NWs@CC can produce increasing amounts of ammonia with the cathodic potential boosting to be more negative, reaching the highest faradaic efficiency (FE) of 93.1% at −0.6 V vs. RHE and the highest rate of 11.6 mg h−1 cm−2 at −0.8 V vs. RHE (Fig. 3c). This can also be confirmed by the corresponding UV-vis spectra of the electrolytes after coloring (Fig. S18, ESI†). It is worth mentioning that the FE remains consistently above 90% at high potentials, indicating the remarkable selectivity exhibited by the S-CoSe2 NWs. The ammonia production of the S-CoSe2 NWs was additionally quantified using 1H nuclear magnetic resonance (NMR) spectroscopy with maleic acid as an internal standard (Fig. S19, ESI†). An FE of 91.5% at −0.6 V vs. RHE was calculated, which aligns well with the results obtained from the UV-vis analysis, further corroborating the high selectivity towards ammonia (Fig. S20, ESI†). The source of nitrogen in the produced ammonia was unequivocally confirmed by 1H NMR experiments employing 15N isotopic labeling. As 15N is a spin-½ nucleus that results in two symmetric signals for 1H-15N, in contrast to the three symmetric signals observed for 1H–14N interactions, which involve a spin-1 nucleus.59 With standard ammonium salt as a reference, the electrolyte after NARR with 15NO3− exhibited distinct double peaks, while triple peaks were observed with 14NO3− as the nitrogen source (Fig. S21, ESI†). The NARR byproducts of S-CoSe2 NWs and CoSe2 NSs, nitrite (NO2−) and the competitive HER product of H2 are also detected, and corresponding FEs are calculated (Fig. 3d and Fig. S22, ESI†). In the case of the S-CoSe2 NWs, as the potential is progressively increased to the positive range, the FEs of nitrite gradually decrease, falling below 5%, with ammonia becoming the dominant product. Conversely, for the CoSe2 NSs, a similar trend is observed with nitrite, albeit with a higher FE share. The superior selectivity towards ammonia exhibited by the S-CoSe2 NWs further validates the role of sulfur doping in providing higher charge active centers, which enhances the adsorption of intermediates. We further investigated the NARR performance of S-CoSe2 nanomaterials with different sulfur doping contents ranging from 0.86% to 4.3% (Fig. S23, ESI†). The FE collections clearly show an increase in the FE for the target ammonia, accompanied by a decrease in the FE for the byproduct nitrite, which strongly confirms the pivotal role of sulfur in enhancing the NARR performance. Moreover, bulk CoSe2 was also synthesized as a reference (Fig. S24, ESI†). It exhibited inferior activity and undesirable selectivity towards ammonia, underscoring the superiority of the nanostructured S-CoSe2 NWs (Fig. S25, ESI†). Besides activity, the stability of the S-CoSe2 NWs is also evaluated using a long-term chronopotentiometry test in a self-organized device (Fig. S26, ESI†). We fabricated this flow cell device by coupling the S-CoSe2 NWs@CC as the cathode and NiFe LDH@ Ni foam as the anode, which was separated by a Nafion membrane (Fig. S27, ESI†), under a high current of 200 mA cm−2; this device could continuously produce ammonia over 500 h without an obvious surge in the cell voltage. The structural integrity and electronic structure of the S-CoSe2 NWs were further examined using TEM and XPS after a long-term reaction. The analyses revealed that S-CoSe2 NWs nearly retained their original microstructures and electronic configuration, underscoring their robustness (Fig. S28, ESI†). After the operation, the ammonia produced can be further fixed in the solid form of struvite ((NH4)MgPO4·6H2O) by adding a coprecipitator of Mg2+ and PO43−, which can be further used as fertilizer. Using this struvite-precipitation method, we obtained over 150 g of struvite with high purity (Fig. S29, ESI†). It also provides a promising way to capture nitrite and phosphate in eutrophicated water and is further convert to a value-added fertilizer.
This superior performance motivated us to investigate the reaction mechanism and uncover the reason for the distinct selectivity difference between CoSe2 NSs and S-CoSe2 NWs. First, in situ electrochemical spectroscopy was employed to identify the nitrogen-containing intermediates during NARR. Attenuated total reflection Fourier-transform infrared spectroscopy (ATR-FTIR, Fig. S30, ESI†) was performed to detect the species adsorbed on the electrode surface, and the reference spectrum obtained under open circuit potential was collected as the background. Then, the potential in the operating range was stepped from −0.3 V to −0.7 V vs. RHE. With increasing potential, several downward peaks emerged that were independent of the working potential, indicating the generation of NARR intermediates. The downward bands at around 1150 cm−1 and 1210 cm−1 can be, respectively, attributed to the –N–O– stretching vibration of NH2OH and the N–O antisymmetric stretching vibration of NO2−, which are both important intermediates during NARR. Furthermore, the upward band at around 1350 cm−1, assigned to the asymmetric stretching vibration of NO3−, indicates the consumption of nitrate at the electrode interface.60 For CoSe2 NSs, the FTIR spectra collected were similar to those of S-CoSe2 NWs, indicating the analogous NARR route on both catalysts. However, the adsorption band center of NO2− on the CoSe2 NSs has an obvious red shift to 1224 cm−1, indicating the adsorption attenuation of NO2− on the materials, which could be difficult for further hydrogenation (Fig. S31, ESI†). Besides ATR-FTIR, in situ electrochemical Raman spectroscopy was performed to obtain adminicular information about the intermediates during the NARR. The Raman shifts at 979 cm−1 and 1046 cm−1 can be attributed to the sulfate and nitrate groups. The peaks observed at 1373, 1535, and 1598 cm−1 are associated with the symmetric N–O stretches of adsorbed NO2−, NO stretches of NOH, and H–N–H of NH2, respectively.61,62 The possible NARR intermediates were also captured by online differential electrochemical mass spectrometry (DEMS) with consecutive cycles for uncovering the integrated mechanism.63 The major mass–charge ratio (m/z) signals of NH(15), NH2(16), and NH2OH (33) were definitely captured as reaction intermediates, and NH3(17) with a strong signal intensity indicates the formation of the NARR product NH3. It should be noted that the m/z signal of 46 represents the possible byproduct of NO2−, and its weak intensity also reversely validates the high selectivity of S-CoSe2 NWs. Combining the information provided by the above electrochemical in situ technologies, the nitrate reduction pathway for the given materials can be proposed as follows: NO3− → NO3* → NO2* → NO* → NOH* → NOH* → N* → NH* → NH2* → NH3.
We further conducted density functional theory (DFT) calculations to determine the possible reaction mechanism based on the structural model of sulfur-doped CoSe2 and observed the experimental evidence. First, the (210) facet was selected as the NARR surface due to its lower surface energy than that of the exposed (311) facet (Fig. S32, ESI†). As NARR is a multiple electron–proton transfer process with many intermediates and adsorption models on active sites, the most possible pathway is illustrated (Fig. 4d).64 In terms of thermodynamics, the nitrate adsorption on sulfur-doped CoSe2 is unfavorable with an uphill free energy of 0.47 eV during the entire catalytic processes, which is lower than that on CoSe2 (0.55 eV, Fig. S33, ESI†). The first hydrogenation on the *NO3 and *NO2 (*NO3 to *NO2OH, *NO2 to *NOOH) without electron transfer shows uphill free energies of 0.08 eV and 0.17 eV, respectively. The rate-determining step (RDS) on S-CoSe2 NWs appeared on the hydrogenation of *NO (*NO to *NOH) with a free energy increase of 1.21 eV, and CoSe2 shows the same RDS but with a lower free energy increase of 1.08 eV. This result contradicts the fact that S-CoSe2 NWs with lower onset potential and overpotential have higher apparent activity. This inconformity could be explained by the strong competition of H* and nitrogen-containing species produced during NARR. Sulfur has a higher *H adsorption capability, which can promote hydrogenation of NARR; however, excess *H accumulation will also lead to the decrease of NARR selectivity, and the adjacent active sites of cobalt can in situ utilize the produced *H to balance the kinetics of production and utilization of *H. Moreover, as mentioned before, the structural advantage of S-CoSe2 NWs with a higher BET surface area and more abundant active sites cannot be ignored; it also contributes significantly towards apparent activity. The reduction steps of *NOH to NH3* on sulfur-doped CoSe2 are nearly downhill in free energy, which contributes to the remarkable catalytic performance of S-CoSe2 NWs. CoSe2 follows similar reaction steps and trends with sulfur-doped CoSe2, but the selectivity towards NO2− between them is distinct at all given potentials. This could be explained by the difference in the free energy between the hydrogenation and desorption of *NO2; this value is 0.22 eV for CoSe2, which is higher than that of sulfur-doped CoSe2 with a high-charge active center (0.12 eV), indicating a higher tendency of hydrogenation on the latter, which also reflects the better ammonia selectivity (Fig. S34, ESI†). The electronic structure analysis showed the higher charge state of cobalt on the S-CoSe2 NWs. Combined with the above DFT calculation results, the high activity and selectivity of ammonia can be attributed to the high-charge active sites.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4mh00593g |
This journal is © The Royal Society of Chemistry 2024 |