So Eun
Jang
a,
Jae Young
Kim
*b and
Duck Hyun
Youn
*a
aDepartment of Chemical Engineering, Department of Integrative Engineering for Hydrogen Safety, Kangwon National University, Cheucheon 24341, South Korea. E-mail: youndh@kangwon.ac.kr
bKorea Research Institute of Chemical Technology, Daejeon 34114, South Korea. E-mail: jaeykim@krict.re.kr
First published on 24th October 2024
The electrochemical nitrate reduction reaction (NO3−RR) is emerging as a promising method for ammonia production under ambient conditions while simultaneously addressing nitrate pollution. Due to the complexity of NO3−RR, which involves multi-electron/proton transfer and competes with the hydrogen evolution reaction (HER), the development of efficient electrocatalysts with high activity and stability is crucial. In this study, we report the use of Mo2C nanoparticles homogeneously dispersed on a carbon nanotube-reduced graphene oxide hybrid support (Mo2C/CNT-RGO) as an effective electrocatalyst for NO3−RR. The three-dimensional CNT-RGO hybrid provides a large surface area for electrolyte contact, enhanced electrical conductivity, and prevents the aggregation of Mo2C nanoparticles. Consequently, the Mo2C/CNT-RGO electrocatalyst demonstrated high NO3−RR performance, achieving a maximum NH3 production rate of 5.23 mg h−1 cm−2 with a faradaic efficiency of 95.9%. Mo2C/CNT-RGO also exhibited excellent long-term stability during consecutive cycling tests. This work presents a promising strategy for developing high-performance and durable NO3−RR electrocatalysts.
Recently, the electrochemical nitrogen reduction reaction (NRR) has attracted significant interest as it utilizes water as a sustainable proton source under mild conditions.15–17 However, NRR faces significant challenges, such as extremely low faradaic efficiency and NH3 yield, attributable to the poor solubility of N2 in water, highly stable NN bond (941 kJ mol−1), and competing hydrogen evolution reaction (HER).18–22 As an alternative to NRR, the research on electrochemical nitrate reduction reaction (NO3−RR) is growing extensively.23–25 This growth is driven by substantially higher NH3 production rates compared to NRR, owing to the high solubility of NO3− in aqueous solutions and the lower bond energy of NO (204 kJ mol−1).26–28 Moreover, NO3− is abundantly present as a contaminant in surface and groundwater due to the excessive use of nitrogen-containing fertilizers and chemicals, posing significant health risks, such as blue baby syndrome, non-Hodgkin's lymphoma, and other cancers.26,29 Therefore, NO3−RR offers a “two birds-one stone” solution by converting nitrate pollutants into value-added ammonia while simultaneously purifying water. However, the NO3−RR is a complex process involving an eight-electron and nine-proton transfer with sluggish kinetics, and it includes several intermediates (e.g., NO2−, N2, and N2H4).30–35 Thus, the development of efficient electrocatalysts with high activity and stability for NO3−RR is imperative.
Various noble metal-based catalysts, including Pt, Rh, and Pd, have been employed as NO3−RR catalysts, demonstrating high faradaic efficiency and stability.36–38 However, their high costs and limited availability hinder the large-scale application of NO3−RR systems. Transition metal carbides (TMCs) have emerged as promising alternatives to Pt-group metals in various electrocatalytic processes due to their electronic structures, which resemble those of noble metals.39–41 Among TMCs, molybdenum carbides (Mo2C) have been employed as catalysts for both NRR and NO3−RR owing to their high adsorption capacity and catalytic hydrogenation ability for electron-rich nitrogenous molecules.42,43 Nevertheless, the aggregation and large particle size of Mo2C, typically resulting from high synthesis temperatures, reduce the exposure of active sites, and consequently impair catalytic activity.44 Therefore, there remains significant scope for improving the activity and stability of Mo2C, particularly for NO3−RR applications.
To enhance the electrocatalytic activity of Mo2C, one effective strategy is to compound it with carbonaceous materials as catalyst supports. Carbonaceous materials, such as carbon nanotubes (CNTs) and graphene, offer a large surface area and high electrical conductivity, which can enhance the performance of loaded Mo2C.45–49 However, when used individually, graphene tends to stack due to van der Waals forces and strong π–π interactions between nanosheets.50,51 Similarly, CNTs have a tendency to bundle together, which reduces their surface area and electrochemical performance.52 The combination of CNTs and graphene (forming a CNT-RGO hybrid) addresses the stacking and bundling challenges associated with their individual use by creating a three-dimensional structure. This hybrid CNT-RGO framework not only provides a larger contact area with the electrolyte but also facilitates improved electron transfer, thereby enhancing the overall electrocatalytic activity.
In this study, nanostructured Mo2C supported on a CNT-RGO hybrid (Mo2C/CNT-RGO) was prepared using a modified urea-glass route for NO3−RR.53,54 The Mo2C/CNT-RGO composite comprised uniformly dispersed 7 nm Mo2C nanoparticles on a three-dimensional CNT-RGO hybrid support. The resulting Mo2C/CNT-RGO demonstrated high NO3−RR activity, achieving a maximum NH3 production rate of 5.23 mg h−1 cm−2 and a faradaic efficiency (FE) of 95.9%, outperforming the Mo2C/CNT and Mo2C/RGO catalysts. Additionally, it exhibited excellent long-term stability over ten consecutive cycling tests. The enhanced NO3−RR performance of Mo2C/CNT-RGO can be attributed to the synergistic interaction between the active Mo2C nanoparticles and the CNT-RGO hybrid support, which provided a high surface area and enhanced electrical conductivity.
Fig. 2a shows the TEM image of the Mo2C/CNT-RGO catalyst, where CNTs are randomly distributed across the RGO layers, and spherical Mo2C particles are uniformly dispersed on the CNT-RGO hybrid support without aggregation. The average particle size of Mo2C is approximately 7 nm. The high-resolution TEM (HRTEM) image in Fig. 2b reveals lattice fringes of 0.224 and 0.240 nm, corresponding to the (101) and (002) facets of the Mo2C phase, respectively. Fig. 2c and e present the TEM images of Mo2C/CNT and Mo2C/RGO, showing that ∼8 nm Mo2C nanoparticles are exclusively anchored onto the CNTs and RGO, respectively. Similar to Fig. 2b, lattice spacings of 0.224 and 0.240 nm were detected, confirming the presence of the Mo2C phase (Fig. 2d and f).
The XRD patterns of the prepared catalysts are presented in Fig. 3a. All catalysts exhibit identical XRD patterns, aligning with the reference XRD patterns of hexagonal Mo2C (JCPDS No. 00-035-0787). The observed peaks at 34.4°, 37.9°, 39.4°, 52.2°, 61.6°, 69.6°, 74.6°, and 75.5° correspond to the (100), (002), (101), (102), (110), (103), (112), and (201) planes of Mo2C, respectively. No impurity peaks, such as Mo metal or MoOx, were detected, confirming the successful synthesis of pure Mo2C on the CNT-RGO, CNT, and RGO supports. The absence of peaks around 10° further indicates the conversion of GO to RGO during thermal annealing.56
Fig. 3 (a) XRD patterns of the synthesized catalysts. XPS spectra of Mo2C/CNT-RGO for (b) C 1s, (c) Mo 3d, and (d) N 1s. |
The chemical states of Mo2C/CNT-RGO were investigated via XPS, as shown in Fig. 3b–d. The survey spectrum of Mo2C/CNT-RGO confirms the presence of C, Mo, N, and O elements (Fig. S1a†). In the C 1s spectrum (Fig. 3b), the deconvoluted peaks at 283.2, 284.6, 285.6, and 286.5 eV correspond to the Mo2C, C–C, C–N, and C–O bands, respectively.57,58 Fig. S1b† shows the C 1s spectrum for GO, where broad peaks observed at 280–290 eV indicate the presence of oxygen-containing functional groups such as hydroxyl, carboxyl, and epoxy groups. The significantly reduced peak intensities in the C 1s spectrum of Mo2C/CNT-RGO suggest the conversion of GO to RGO during synthesis. The Mo 3d spectrum in Fig. 3c is deconvoluted into Mo2+ (228.3/231.5 eV), Mo4+ (229.2/233.6 eV), and Mo6+ (232.5/235.6 eV).59 The Mo4+ and Mo6+ species are attributed to MoO2 and MoO3, respectively, due to unavoidable surface oxidation,42,43,60 while the Mo2+ species stem from the Mo2C phase.61–63 The N 1s spectrum (Fig. 3d) exhibits three main peaks at 396.1, 398.3, and 400.6 eV, corresponding to N–Mo, pyridinic N, and graphitic N, respectively. The peak observed at 394.4 eV is assigned to Mo 3p due to binding energy overlap.57 Fig. S2 and S3† display the XPS results of Mo2C/CNT and Mo2C/RGO, with their C 1s, Mo 3d, and N 1s spectra showing similarities to Mo2C/CNT-RGO, indicating comparable chemical states among these catalysts.
The electrical conductivities of the catalysts were measured using the four-point probe method, with results summarized in Table 1. Among the catalysts, Mo2C/CNT-RGO exhibited the highest electrical conductivity of 3.04 × 103 S m−1, followed by Mo2C/CNT (1.98 × 103 S m−1), and Mo2C/RGO (0.64 × 103 S m−1).
Sample | Resistivity [ohm cm] | Electrical conductivity × 103 [S m−1] |
---|---|---|
Mo2C/CNT-RGO | 3.28 × 10−2 | 3.04 (±0.14) |
Mo2C/CNT | 5.05 × 10−2 | 1.98 (±0.57) |
Mo2C/RGO | 1.58 × 10−1 | 0.64 (±0.07) |
The high electrical conductivity, which is crucial for enhanced NO3−RR activity, was effectively achieved by employing the CNT-RGO hybrid support for Mo2C.
Electrochemical NO3−RR measurements were conducted in a 0.1 M KOH solution with 0.1 M NaNO3 under ambient conditions. LSV curves were used to investigate the electrocatalytic responses of the catalysts both in the presence and absence of NO3− (Fig. S4a†). In 0.1 M KOH (dashed line), the LSV curves displayed a delayed onset potential. However, the addition of 0.1 M NaNO3 (solid line) resulted in significant changes, including an increase in current density and a reduction in onset potential, indicating a strong NO3− reduction response. As shown in Fig. 4a, the LSV curves for the prepared catalysts in the presence of 0.1 M NaNO3 demonstrated that Mo2C/CNT-RGO exhibited a higher current density across the entire potential range, indicating its superior electrocatalytic activity for nitrate reduction.
Fig. 4b presents the Tafel plots of the prepared catalysts fitted to the Tafel equation (η = b log|J| + a, where b is the Tafel slope and J is the current density). A lower Tafel slope typically signifies better reaction kinetics. The Tafel slope of Mo2C/CNT-RGO was 140 mV dec−1, which is smaller than that of Mo2C/CNT (172 mV dec−1) and Mo2C/RGO (188 mV dec−1), suggesting that Mo2C/CNT-RGO possesses higher electrocatalytic activity.
The cyclic voltammogram (CV) curves of the catalysts at various scan rates in the non-faradaic region are shown in Fig. S5,† and the corresponding electrochemical double-layer capacitance (Cdl) values are provided in Fig. 4c. Mo2C/CNT-RGO exhibited a larger Cdl of 25.74 mF cm−2 compared to Mo2C/CNT (18.93 mF cm−2) and Mo2C/RGO (17.01 mF cm−2). The Cdl is generally proportional to the contact area between the catalyst and electrolyte. Employing the CNT-RGO hybrid support leads to a larger contact area due to its 3D network structure, contributing to the enhanced NO3−RR activity of Mo2C/CNT-RGO.64,65 This is further supported by the adsorption capacity (qe, the adsorbed NO3− amount per unit catalyst amount) measurements, which confirm the effectiveness of the CNT-RGO hybrid.66 In Fig. S6,† the Mo2C/CNT-RGO recorded a much higher qe for NO3− (67.5 g gcat−1) compared to Mo2C/CNT (42. 9 g gcat−1) and Mo2C/RGO (32.8 g gcat−1), indicating that CNT-RGO provides an enhanced surface area for reactant adsorption.
Fig. 4d shows the Nyquist plots of the catalysts obtained from electrochemical impedance spectroscopy (EIS). The charge transfer resistance (Rct) at the catalyst-electrolyte interface is represented by the semicircle in the Nyquist plot and is inversely proportional to the electrocatalytic activity. The measured Rct values were 16.2, 33.5, and 47.5 Ω for Mo2C/CNT-RGO, Mo2C/CNT, and Mo2C/RGO, respectively, suggesting faster electron transfer and thus enhanced NO3−RR activity of Mo2C/CNT-RGO.
Chronoamperometry (CA) tests were conducted over a potential range from −0.3 V to −0.6 V for 1 h (Fig. S4b†). The concentrations of NO2−, NO3−, and NH3 products were quantified using the colorimetric method, with their standard calibration curves shown in Fig. S7.† As illustrated in Fig. 4e and f, the Mo2C/CNT-RGO demonstrated a higher NH3 yield and FENH3 across the entire potential range compared to Mo2C/CNT and Mo2C/RGO. The NH3 yield for Mo2C/CNT-RGO increased with more negative potentials, reaching a maximum yield of 5.23 mg h−1 cm−2 at −0.6 V (Fig. 4e). In comparison, the maximum NH3 yields for Mo2C/CNT and Mo2C/RGO were 4.55 and 4.53 mg h−1 cm−2, respectively. Additionally, Mo2C/CNT-RGO achieved a maximum FENH3 of 95.92% at −0.3 V and maintained FENH3 values above 90% up to −0.5 V (Fig. 4f), demonstrating superior performance at relatively less negative potentials compared to other reported Mo-based catalysts (Table S1†). A decreasing trend in FENH3 was observed at −0.6 V, primarily due to the increasing dominance of HER at more negative potentials. Although Mo2C/CNT and Mo2C/RGO also demonstrated notable performance with FENH3 values of 85.25% and 84.39% at −0.3 V, respectively, their overall activity was significantly lower than that of Mo2C/CNT-RGO. The three-dimensional CNT-RGO hybrid support offers a larger contact area with the electrolyte (as evidenced by Fig. 4c and Fig. S6†) and exhibits higher electrical conductivity (Table 1). Consequently, when combined with the active Mo2C, this hybrid support structure contributes to the superior NO3−RR performance of Mo2C/CNT-RGO compared to Mo2C/CNT and Mo2C/RGO.
Long-term stability is crucial in determining the performance of catalysts. To evaluate the stability of Mo2C/CNT-RGO, ten consecutive electrolytic cycles were conducted at −0.3 V. As shown in Fig. 5a, both the NH3 yield and FENH3 remained stable throughout the cycling tests, indicating the catalyst's good stability. The morphology of Mo2C/CNT-RGO after the cycling test was examined using TEM. As shown in Fig. S8a,† the Mo2C particles remain well-dispersed on the CNT-RGO hybrid support without significant aggregation. Additionally, the lattice spacings of 0.224 nm and 0.240 nm observed in Fig. S8b† correspond to the (101) and (002) planes of Mo2C, respectively. The similar morphology of Mo2C/CNT-RGO after the cycling test suggests its structural stability.
Possible by-products such as NO2− and H2 were also measured (Fig. 5b). NO2− was detected only in the potential range of −0.3 to −0.5 V and was nearly absent at −0.6 V. As the potential becomes more negative, the FE of H2 gradually increased due to the competing HER. Nevertheless, Mo2C/CNT-RGO maintained high selectivity for NH3 generation across all potentials.
To identify the nitrogen source in the produced NH3, several control experiments were performed. Conducting CA tests at −0.3 V for 1 h in the absence of NO3− and under open-circuit potential (OCP) conditions resulted in negligible ammonia formation (Fig. 5c), confirming that the produced NH3 originated from the NO3−RR. Furthermore, a 15N isotopic labeling experiment was conducted using 1H-NMR. In Fig. 5d, standard solutions of 14NH4Cl and 15NH4Cl exhibited characteristic triplet peaks of 14NH4+ and doublet peaks of 15NH4+, respectively. When 14NO3− (from Na14NO3) was used as the reactant, triplet peaks corresponding to 14NH4+ were observed, while doublet peaks corresponding to 15NH4+ appeared when 15NO3− (from Na15NO3) was used as the reactant. These results strongly confirm that the NH4+ detected in the electrolyte originated from NO3−RR and not from other contaminants.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4dt02817a |
This journal is © The Royal Society of Chemistry 2024 |