Hao
Huang
and
Kaiying
Wang
*
Department of Microsystems, University of South-Eastern Norway, Borre 3184, Norway. E-mail: Kaiying.Wang@usn.no
First published on 16th October 2023
With their abundant metal sites, ordered porous structure and great conductivity, conductive metal–organic frameworks display many excellent single-atom electrocatalytic activities, superior to those of conventional inorganic nanostructures. However, their electrochemical application is greatly limited by the fragility of coordinated frameworks. Here, we describe a metal-covalent organic framework (MCOF) strategy to construct a nitrate reduction (NRA) catalyst using M3·HATN as the subgroup. Assisted by a salt-template, M-HATN-COFs with abundant metal sites (M at% ≈ 12.5%) are achieved by a one-step coordination–condensation approach. More importantly, the M-HATN-COFs provide reasonable platforms for studying the metal-atom catalytic mechanism, surpassing that of current inorganic structures. The Mo-HATN-COFs exhibit outstanding electrocatalytic properties with a high ammonia yield rate (8.52 mg h−1 cm−2), FE (91.3%) and stability for the NRA reaction. As the first work on MCOFs for electrochemical NRA reactions, the M-HATN-COF strategy will innovate the design concept of next-generation catalysts and the catalytic mechanism of single-metal atoms.
Currently, the worldwide NH3 production industry relies on the Haber–Bosch method, which involves the reaction between hydrogen and nitrogen under harsh conditions (high pressure, ≥200 atm and high temperature, ≥400 °C).8,9 However, this method results in extreme energy consumption (≈38 GJ tNH3−1) and carbon emission (≈2.9 tCO2 per tNH3).10 In response to the growing demand for NH3, there is a need to expand the production scale and improve technology. As an attractive alternative, electrochemical nitrogen reduction offers a strategy that uses nitrogen and water as sources for NH3 synthesis under mild conditions.11,12 However, direct dinitrogen electroreduction faces several challenges: dinitrogen has an ultrahigh energy barrier to break the inert NN bond of 941 kJ mol−1 and it exhibits extremely low solubility in water (17.1 mg kg−1water).13–15 Additionally, as a nonpolar molecule, dinitrogen shows weak absorption ability on the surface of electrodes.16 These three obstacles severely limit the utilization of direct dinitrogen electroreduction. As an alternative, nitrate (NO3−) presents unique advantages to boost the electrochemical reduction strategy for NH3 production: (1) low bond energy (204 kJ mol−1); (2) large water solubility (383 g kg−1water for potassium nitrate at 25 °C); and (3) high chemical polarity for good absorption at active sites.16–19 Furthermore, due to its long-term accumulation as a result of agricultural and industrial activities, NO3− acts as a noxious pollutant that can cause various diseases in human tissue.20,21 Therefore, the reduction of NO3− to NH3 (NRA, NO3− + 9H+ + 8e− → NH3 + 3H2O, −0.12 V vs. SHE) not only enhances the efficiency of NH3 production, but also greatly alleviates pollution.11
As an emerging technology, various electrocatalysts have been exploited for the NRA reaction, including noble metal nanoparticles, transition metal nanostructures and alloys, oxides, phosphides, etc.5,22,23 These electrocatalysts rely on inorganic metal-based nanomaterials, where the active sites for NRA are typically limited to the exposed metal atoms at the surface and/or the edge of the nanostructures.5 The internal metal atoms in these inorganic catalysts are often inert and difficult to utilize for the reduction of NO3− to NH3. This natural limitation results in very low metal-atom utilization and catalytic efficiency of current NRA catalysts. Fortunately, owing to their abundant exposed metal sites and ordered in-plane porous structure, metal–organic frameworks (MOFs) offer a promising approach for maximizing the utilization of metal atoms (in principle, up to 100%) and enhancing the efficiency of electrocatalytic reactions, similar to single-atom catalysts.24 However, the structural stability of metal linkages in MOFs is often fragile, leading to framework dissociation under applied potentials.25,26 To address this challenge, the incorporation of conjugated covalent linkages into c-MOFs to form conductive metal-covalent frameworks (c-MCOFs) could be instrumental.25 Although only a few studies have reported on the construction of electrocatalysts using c-MCOFs, most of them are based on porphyrin subgroups with very low metal content (∼1 at%).27,28 Hexaazatriphenylene (HATN, Fig. S1, ESI†) is considered as a preferred alternative to overcome this limitation because its three bidentate tetraamine moieties can coordinate with most metal atoms, forming M3·HATN with an ultra-high metal content of ∼14 at%.29 Therefore, the construction of c-MCOFs based on M3·HATN would pave the way for designing novel electrocatalysts toward the NRA reaction. The exploitation of c-MCOF structures is still in its early stage, especially in terms of c-MCOFs with high metal content for electrochemical NRA catalysts.
Herein, we propose a novel strategy for constructing NRA electrocatalysts using the c-MCOF structure with an M3·HATN subgroup. Inspired by the active enzyme nitrate reductase found in denitrifying bacteria, we utilize Mo atoms to establish HATN-based c-MCOFs, known as Mo-HATN-COFs. The Mo-HATN-COFs are achieved via a facile coordination–condensation reaction, resulting in structures with an ultra-high Mo content of 12.5 at%, an ordered in-plane porous structure with a size of 1.2 nm and excellent metallicity for good conductivity of 1.5 S cm−2. To further enhance the catalytic activity for NRA, a salt-template approach is employed to synthesize Mo-HATN-COFs with ultra-thin nanosheets (∼1.4 nm). As a result of our experiment, the Mo-HATN-COF nanosheets exhibit a remarkable ammonia yield rate of 8.52 mg h−1 cm−2 (0.50 mmol h−1 cm−2) at −0.5 V vs. RHE. The highest faradaic efficiency (FE) achieved reaches an impressive value of 91.3%. Most importantly, unlike traditional inorganic catalytic structures, c-MCOFs with an M3·HATN subgroup can serve as preferred models for building novel platforms to investigate the mechanism of metal-atom electrocatalysis. Our findings represent a significant contribution towards revealing the inherent mechanisms of and designing next-generation electrocatalysts for the NRA reaction. By pioneering this research, we hope to pave the way for future advancements in this field.
The details of physical characterization, electrochemical measurements, and density functional theory (DFT) calculations can be found in the ESI.†
![]() | ||
Fig. 1 (a) Synthetic route of various M-HATN-COFs. Partial density of states of (b) HATN-COF, (c) Mo-HATN-COF and (d) Ni-HATN-COF slabs. |
To understand the frameworks, the structure of Mo-HATN-COFs was simulated by density functional theory (DFT) calculations. After geometry optimization, the Mo-HATN-COF is a two-dimensional (2D) covalent framework, constituted by a periodic hexagonal ring of six Mo3·HATN subgroups connected through conjugated CN linkages. As with P6/mmm symmetry, the in-plane lattice length of HATN-COFs is about 16.54 Å (a = b; α: 90°, β: 90°, γ: 120°) with an ordered in-plane porous structure of ∼1.2 nm. Meanwhile, experimental data are also used to comprehend the frameworks (Fig. S3†). The X-ray powder diffractometer (XRD) pattern of Mo-HATN-COFs shows a prominent peak at 26.5°, which belongs to the (002) plane, well matched with the simulated XRD result and previous works.31 This result demonstrates that the Mo-HATN-COFs are successfully prepared through the coordination–condensation approach. In addition, the Ni-HATN-MCOF and HATN-COF slabs also show hexagonal chemical structures with similar lattice lengths of 16.48 Å and 16.68 Å, respectively (Table S1†). The introduction of metal atoms can reduce the lattice length from 16.68 Å to 16.48 Å, which indicates that the metal coordination reaction can make the HATN-COF structure more compact and may bring about a higher degree of conjugation for better conductivity. To confirm this, partial density of states (PDOS) calculations were carried out for investigating the electronic band structure of various M-HATN-MCOF slabs. As a result, the HATN-COF slab exhibits semiconductor properties with a band-gap of 1.73 eV (Fig. 1b). In contrast, the Mo-HATN-COFs and Ni-HATN slabs show metallicity without any band-gap (Fig. 1c and d). The PDOS result of the Mo-HATN-COF slab reveals that the electron distribution near the Femi level mainly originates from the electrons of the Mo atoms. Meanwhile, a similar result is observed with the Ni-HATN-MCOF slab (Fig. S4 and S5†). Comparison of the d orbitals shows that the d-center of Mo-HATN-COFs is located at −3.08 eV, which is more negative than that of Ni-HATN-COFs (−2.67 eV). This result demonstrates that the Mo sites are more inert for H+ absorption (HER process) than the Ni sites, which means that the Mo sites may be advantageous to the NRA reaction (Fig. S6†). Four-point probe technology reveals that the compressed pellet of Mo-HATN-COFs has great electrical conductivity of 1.5 S cm−2, much larger than that of the HATN-COFs without Mo sites (0.06 S cm−2). To directly prove the enhanced conductivity of Mo-HATN-COFs, we employed EIS tests on Mo-HATN-COF and HATN-COF samples (Fig. S7†). The Rct value of HATN-COFs is 521.9 Ω, which is much lower than that of non-conjugated COFs. After introducing Mo atoms into the HATN-COF structure, the Rct value drops to 117.2 Ω, indicating that the incorporation of Mo atoms can enhance the conductivity of the HATN-COF structure.
For fully tapping into the potential for NRA electrocatalysis, an ultrathin structure is also incorporated into the M-HATN-COFs. A salt-template serves in a directed strategy for achieving ultrathin M-HATN-COFs nanosheets, providing an effective route toward the growth of various few-layered inorganic nanostructures.32–34 As a result of its excellent stability, flat surface and low cost, a sodium chloride crystal can be an inert planar substrate to guide the growth of most 2D materials.35–37 In this work, we first propose that a sodium chloride crystal (Fig. S8†) serves as an agent to orientate construction of ultrathin COF nanosheets (Fig. 2a). The morphology of various M-HATN-COFs is confirmed by scanning electron microscopy (SEM). Without the NaCl salt-template, Mo-HATN-COFs show severely reunited layered structures as bulk materials (Fig. 2b). In contrast, the SEM image of template-directed Mo-HATN-COFs shows that they have an obvious ultrathin layered morphology with generous wrinkles (Fig. 2c). The interaction between these nanosheets results in not only an extensive flower-like structure, but also a large-scale hierarchical porous structure for Mo-HATN-COFs, where the size distribution of pores ranges from 5 to 100 nm (Fig. S9†). As far as we know, the creation of this hierarchical structure is a very efficient strategy for rapid mass transfer, improving the interaction between the catalytic substrate and active centers. In this work, the hierarchical mesoporous structure can rapidly transfer nitrate from the electrolyte to the Mo sites and releases ammonia once the reduction is completed, which is beneficial for accelerating the process of adsorption and desorption in the electrochemical reduction reaction. To investigate the porous structure, N2 adsorption/desorption testing was employed. The isotherm of Mo-HATN-COF nanosheets indicates a large surface area of 445.6 m2 g−1 and two types of size distribution: (1) a wide peak at about 18.7 nm, which can be attributed to the interaction with hierarchical pores; and (2) a sharp peak at 1.3 nm attributed to the ordered in-plane porous structure, the same as that for the DFT slab (Fig. S10†). In contrast, the Mo-HATN-COFs without a salt-template show a blocky-shaped nanostructure, which consists of agglomerated layered nanosheets. This comparison indicates the effective function of the salt-template-directed strategy for engineering the 2D structure of HATN-COFs. Similarly, the SEM images show that the Ni-HATN-MCOF and HATN-COF samples also have ultrathin-layered structures with hierarchical pores, indicating the universality of the salt-template-directed strategy for 2D COF materials (Fig. 2d and S11†).
In transmission electron microscopy (TEM) images, similar to the results of SEM, these M-HATN-COF nanosheets have a smooth silk-like layered structure (Fig. 3a–c) and the Mo-HATN-COF nanosheet thickness is about 1.4 nm (Fig. 3d), which is the same as those of the Ni-HATN-MCOF and HATN-COF nanosheets (Fig. S12 and S13†).38 The high-resolution TEM images show that the M-HATN-COF nanosheets exhibit good crystallinity (Fig. 3e and f). Thus, the results of morphological characterization illustrate that the salt-template-directed strategy can serve as a highly feasible method to control the growth of 2D structures on COFs. To inspect the elemental composition, energy dispersive spectroscopy (EDS) mapping was carried out on the HATN-COF nanosheets. The mapping image reveals that the Mo and N elements are present and uniformly distributed in the Mo-HATN-COF nanosheets (Fig. 3g).
Inductively coupled plasma (ICP-OES) technology was employed to analyze the actual Mo content. The ICP-OES result shows that the Mo content in the Mo-HATN-COF is 7.9 at% (40.08 wt%), which is lower than its theoretical value (12.5 at%; 52.90 wt%). In our opinion, this discrepancy in Mo-HATN-COFs may be attributed to insufficient coordination between Mo ions and 1,2,4,5-tetraaminobenzene tetrahydrochloride in the synthesis process. To further confirm the structure of Mo-HATN-COFs, XPS analysis was carried out. The XPS analysis confirms the presence of Mo, N, and C elements in Mo-HATN-COFs. The Mo XPS spectra of Mo-HATN-COFs are consistent with previously reported results (Fig. S14a†), showing two peaks at 230.9 and 234.1 eV, which are ascribed to the Mo5+ 3d5/2 and 3d3/2, respectively.39 The peak at 237.3 eV indicates the presence of an Mo6+ structure, possibly resulting from mild oxidation during the synthetic process. The XPS spectra of N 1s (Fig. S14b†) exhibit two characteristic peaks at 398.3 and 401.1 eV, attributed to NC and N–H, respectively, indicating the existence of C
N linkages.40 Moreover, the peak observed at 394.9 eV indicates the formation of an Mo–N moiety.41 For C 1s, the spectrum of the Mo-HATN-COF is deconvoluted into three carbon species of C–C (284.9 eV), C
N (286.3 eV) and C–O (288.7 eV), confirming the existence of C-based frameworks (Fig. S14c†). Thus, all experimental data and analyses manifest the achievement of M-HATN-COF nanosheets in this work.
The NH3 yields of various c-MCOF samples were analyzed using chromogenic methods on the electrolyte after electrocatalysis (Fig. S20 and 21†). On comparing with absorbance calibration curves, it was observed that the Mo-HATN-COF nanosheets exhibit an NH3 yield of 8.52 mg h−1 cm−2 (0.50 mmol h−1 cm−2), dramatically superior to that of the bulk Mo-HATN-MCOFs (2.37 mg h−1 cm−2) at −0.5 V vs. RHE (Fig. 4b). At the applied potential of −0.4 V, the TOF (turnover frequency) value of the Mo-HATN-COF, calculated from the LSV curves, is 0.714 NH3 per s per site, indicating that Mo-HATN-COFs exhibit the best intrinsic activity for ammonia production. This shows that engineering the 2D structure is an important factor for regulating the electrocatalytic behaviors of COFs. In this work, the ultrathin structure improves the NRA activity of HATN-COFs due to the increased exposure to metal sites. To investigate the source of electrocatalytic activity, the NRA behaviors of Ni-HATN-MCOF and HATN-COF nanosheets were also evaluated. The Ni-HATN-MCOF and HATN-COF samples demonstrate lower NH3 yields of 1.92 and 0.49 mg h−1 cm−2, respectively, revealing that the metal atoms are the highly active sites for NRA electrocatalysis and the activity order is Mo sites > Ni sites > N sites. The NH3 yield on Mo-HATN-COFs firstly increases from −0.2 to −0.5 V (vs. RHE); subsequently, the yield decreases with the continued increase of applied potentials (Fig. 4c). The FE is enhanced with an increase in applied potential (from −0.2 to −0.4 V vs. RHE) and reaches the maximum value of 91.3%; beyond −0.40 V (vs. RHE), the FE obviously decreases because the competitive reaction (HER) dominates at a large applied potential and the Mo sites begin to adsorb protons instead of nitrates. The activity of Mo-HATN-COFs is compared with other MOF-, COF- and Ni-based catalysts (Table S2†). It can be concluded that the Mo-HATN-COFs in this work show better electrocatalytic properties than the other MOF-, COF- and Ni-based catalysts. The bulk Mo-HATN-COF sample exhibits the best FE of 77.8% at −0.4 V (vs. RHE) (Fig. S22†). The Ni-HATN-MCOF nanosheets reach the best FE of 31.6% at −0.30 V (vs. RHE) (Fig. S23†), even lower than that of HATN-COFs (57.1% at −0.70 vs. RHE, Fig. S24†). The lower FE of Ni-HATN-MCOFs may be attributed to the better HER activity of Ni sites. Thus, the strategy of simulating enzymes is feasible; the HATN-COFs exhibit the optimal electrocatalytic activity for the NRA reaction. The selectivity for NH3 is another factor to estimate the activity of NRA catalysts, because of the nitrite by-product (NO3− + 2H+ + 2e− → NO2− + H2O, 0.93 vs. SHE). In this work, the Mo-HATN-COF nanosheets demonstrate good NH3 selectivity of ∼95.8% in the NRA reaction, indicating their high efficiency for NH3 production (Fig. 4d). For the Ni-HATN-MCOF nanosheets, the selectivity is ∼92.7%, higher than that of HATN-MCOFs (80.6%). This result indicates that the metal sites (Mo or Ni) exhibit better catalytic activity for reduction of NO3− to NH3, but some reduction reactions stop at the NO2− step on nanosheets with the non-metal sites (N sites). Finally, to investigate the effect of nitrate concentration on catalytic activity, nitrates of various concentrations (0.01, 0.02, 0.05, 0.1 and 0.5 M) were used as nitrogen sources in the electrolyte (Fig. 4e). The yield curve shows that the NH3 production becomes faster as the concentration increases. The FE slightly increases as the concentration is increased until 0.1 M NO3−. However, at higher concentrations, the FE decreases because the produced ammonia cannot leave the active sites in time.20 All electrochemical results demonstrate that HATN-COF nanosheets have the optimal catalytic activity for NRA, which originates from three factors: (1) highly active Mo sites for NRA; (2) the fully conjugated framework for electron transfer; and (3) the hierarchical and in-plane porous structure for mass transfer. Moreover, the M-HATN-MCOF nanosheets also act as ideal platforms to investigate the catalytic activity of metal atoms. In this work, we find that the activity of the Mo atom is better than that of the Ni atom and the non-metal atom (nitrogen). The scope of the MCOF platform can be extended to most metal species due to the very strong coordinating ability of the bidentate tetraamine in the HATN subgroup.
To further investigate the electrocatalytic property for NRA, several control experiments were conducted. Firstly, in the absence of nitrate, no NH3 was detected in the electrolyte (Fig. S25†). Only trace amounts of NH3 were produced during the electroreduction process under open circuit conditions. This indicates that the presence of a nitrogen source and an appropriate applied potential are necessary conditions for NH3 production. Secondly, isotopic labeling experiments were performed to verify the origin of NH3 (Fig. 4f).42,43 The 1H NMR result shows three signals assigned to the 14N atom in the electrolyte containing normal 14NO3−. In contrast, the NMR spectrum exhibits two 15N signals when labeled Na15NO3 is used. This clearly illustrates that the nitrogen element in as-prepared NH3 originates solely from the nitrate in the electrolyte. Moreover, the stability of these electrocatalysts determines their practical utility. When an applied potential of −0.5 V vs. RHE was used, the current density of Mo-HATN-COFs showed only a slight loss of 13.4% after 10 hours of electrocatalytic NRA (Fig. 5a). Furthermore, the NH3 yield rates of Mo-HATN-COFs remained high, exceeding 8.31 mg h−1 cm−2 with the FEs of at least 83.6% (Fig. 5b) in 5 consecutive electrochemical tests. SEM and XRD techniques were used to investigate the effect of the electrochemical process on the structure. The SEM image reveals a thin-layered morphology with a porous structure, which has not changed at all. This also indicates that the Mo-HATN-COFs maintain stable chemical structures in long-term electrochemical reduction reactions (Fig. 5c). More importantly, the XRD result provides direct evidence to prove the structural stability of Mo-HATN-COFs. After 10 h of electrocatalysis, the XRD pattern of Mo-HATN-COFs shows an obvious diffraction peak at 26.4° (Fig. S26†), which is similar to that of the as-prepared Mo-HATN-COF sample (26.5°). Thus, these results demonstrate excellent structural stability during the long-term electrochemical NRA process. These findings indicate the good stability of Mo-HTAN-MCOFs toward NRA. All above-mentioned results collectively highlight that M-HATN-COF nanosheets serve as exceptional catalytic models for efficient NRA electrocatalysis.
Footnote |
† Electronic supplementary information (ESI) available: Details of physical characterization, electrochemical measurements, and density functional theory calculations. See DOI: https://doi.org/10.1039/d3gc01914d |
This journal is © The Royal Society of Chemistry 2023 |