Dongxu
Jiao‡
ab,
Yilong
Dong‡
b,
Xiaoqiang
Cui
*b,
Qinghai
Cai
ac,
Carlos R.
Cabrera
d,
Jingxiang
Zhao
*a and
Zhongfang
Chen
*e
aCollege of Chemistry and Chemical Engineering, Key Laboratory of Photonic and Electronic Bandgap Materials, Ministry of Education, Harbin Normal University, Harbin 150025, China. E-mail: xjz_hmily@163.com
bState Key Laboratory of Automotive Simulation and Control, School of Materials Science and Engineering, Key Laboratory of Automobile Materials of MOE, Jilin University, Changchun 130012, China. E-mail: xqcui@jlu.edu.cn
cHeilongjiang Province Collaborative Innovation Center of Cold Region Ecological Safety, Harbin 150025, China
dDepartment of Chemistry and Biochemistry, The University of Texas at El Paso, El Paso, Texas 79968, USA
eDepartment of Chemistry, University of Puerto Rico, Rio Piedras Campus, San Juan, Puerto Rico 00931, USA. E-mail: zhongfangchen@gmail.com
First published on 17th November 2022
By combining theoretical and experimental efforts, we designed the MoP-(101) surface and explored its potential as catalyst for urea production. Our computations revealed that N2 and CO2 reactants can be effectively reduced to urea on the MoP-(101) surface with a low limiting potential (−0.27 V); the competitive side reactions are well suppressed, and the *NHCONH species is a key reaction intermediate for the C–N coupling during urea synthesis. Our experimental measurements confirmed the above theoretical predictions: urea synthesis was achieved with the urea formation rate of 12.4 μg h−1 mg−1 and the faradaic efficiency of 36.5%. This work not only highlights the critical role of the abundant Mo active sites for urea electrosynthesis but also provides a new mechanism for C–N coupling, which may be used to further develop other efficient electrocatalysts.
To date, industrial-scale urea production mainly comes from the reaction of NH3 with CO2.8,9 Note that the Haber–Bosch process still dominates NH3 synthesis from the industrial fixation of N2 and H2,10–12 of which about 80% is consumed for urea generation.13 The Haber–Bosch process requires harsh conditions (400–650 °C and 200–400 bar) due to the inert NN triple bonds; it is responsible for 1–2% of world's energy consumption and ca. 1.4% of CO2 emissions.14,15 In this scenario, the exploration of a sustainable, efficient, and environmentally benign route enabling N2 and CO2 to directly convert into urea at ambient conditions is highly desirable.16 In this regard, electrocatalysis holds great promise for urea synthesis, as it can effectively convert renewable energy and feedstock into high-value-added chemicals under mild conditions.17
Very recently, the simultaneous electrocatalytic activation of N2 and CO2 to generate urea by their C–N coupling has emerged as an attractive alternative strategy to the energy-intensive industrial process.18 Nevertheless, electrocatalytic urea production generally suffers from extremely low activity and selectivity due to the following substantial challenges: (1) weak adsorption of inert N2 and CO2 reactants; (2) difficulty in dissociating NN and C
O bonds; (3) the kinetically sluggish C–N coupling.19 Therefore, these challenges have to be overcome when developing highly efficient catalysts for urea production. Encouragingly, some promising catalysts have been recently developed.19–22 For example, Wang's group successfully realized urea electrosynthesis on a PdCu alloyed nanoparticle by the coupling of N2 and CO2 in H2O under ambient conditions;20 Yuan et al. successfully fabricated BiFeO3/BiVO4,9 InOOH,19 Ni3(BO3)2,21 and Bi–BiVO4 (ref. 22) to boost the formation of urea. However, there is still a lack of highly efficient and selective electrocatalysts with specific and abundant active sites that can simultaneously capture and activate N2 and CO2 to form urea, and developing such catalysts is highly rewarding but remains a huge challenge.
Due to its various oxidation states from −2 to +4, molybdenum (Mo) possesses rich and intriguing chemistry,23–25 and Mo-based materials have emerged as promising electrocatalysts in many energy conversion reactions, especially for nitrogen reduction reaction (NRR) and CO2 reduction reaction (CO2RR).26–31 For example, Jiang and coworkers reported that MoB2 exhibits excellent NRR performance with a high faradaic efficiency of 30.84%,31 while Wu et al. fabricated a peculiar sub-monolayer MoS2−x structure that was revealed to display superior electrocatalytic NRR performance at ultralow overpotential (−0.2 V. vs. RHE).29 In addition, Sun et al. reported that MoP nanoparticles supported on In-doped porous carbon have outstanding performance for CO2 reduction to formate, with a high faradaic efficiency (96.5%) and excellent current density (43.8 mA cm−2).30
Among various Mo-based materials, molybdenum phosphide (MoxPy) materials have been actively investigated as non-precious-metal catalysts in electrocatalysis due to their unique physicochemical properties, such as high chemical stability, excellent conductivity, and earth-abundant reserves.31,32 In particular, as the electronegativity of P (2.1) is slightly larger than that of Mo (1.8), the incorporation of P will induce a small electron transfer from Mo to P, leading to a moderate coupling between Mo-d and P-p orbitals. As a result, the d-band center of MoxPy may move down as compared with the pure Mo metals, leading to an optimal binding strength with adsorbates.33 Furthermore, the positively charged Mo atoms may sufficiently boost the N2 and CO2 activation. In addition, when the electron-deficient Mo layer in MoxPy materials is exposed, abundant Mo active sites become available, which may be beneficial to the simultaneous activation of inert N2 and CO2 reactants in urea synthesis. Therefore, it is very likely that MoxPy materials with exposed Mo surface can achieve the goal of highly efficient urea synthesis at ambient conditions.
Herein, by means of density functional theory (DFT) computations, we explored the catalytic activity of MoP material for urea synthesis. As expected, the Mo-terminated (101) surface, which is energetically the most favorable surface of MoP material, can simultaneously bind N2 and CO2 strongly, with the free energy change of −0.70 eV. Interestingly, the adsorbed N2 and CO2 can be further reduced into urea product with a rather low limiting potential of −0.27 V. Remarkably, our theoretical predictions were verified by our proof-of-concept experiments: the as-synthesized MoP material exhibits good catalytic activity with a urea yield of 12.4 μg h−1 mg−1 and selectivity, with a faradaic efficiency of 36.5% (−0.35 V vs. RHE) for urea synthesis.
The (100), (001) and (101) surfaces of MoP with different terminations were considered to evaluate the stability and catalytic activity for urea synthesis. These surface models were built by cleaving the bulk MoP structure, and a 3 × 3 supercell was adopted for these surfaces. All these models consist of four atomic layers; the bottom two atomic layers were fixed, while the upper two layers were fully relaxed during geometry optimizations. The vacuum space was set to be 20 Å in the z-direction to avoid the interactions between periodic images. A 3 × 3 × 1 Monkhorst–Pack k-point mesh was adopted to sample the Brillouin zone.
To assess the thermodynamic stability of different MoP surfaces, the surface energy (γ) was determined by where A is the surface area, Erelaxslab denotes the energy of the relaxed slab, Natoms is the number of atoms in the slab, and Ebulk is the bulk energy per atom. According to this definition, a smaller γ suggests a better thermodynamic stability.
The computational hydrogen electrode (CHE) model41,42 was used to compute the change in the Gibbs free energy (ΔG) of each elementary step during urea electrosynthesis, in which one-half of the chemical potential of hydrogen molecule is equal to the chemical potential of the proton–electron pair. In this CHE model, the ΔG value of each elementary step can be obtained by ΔG = ΔE + ΔEZPE – TΔS + ΔGU, where ΔE t represents the reaction energy difference of reactant and product, which can be directly computed from DFT computations. ΔEZPE and ΔS are the changes in zero-point energies and entropy at room temperature (T = 298.15 K), which can be computed from the vibrational frequencies. The entropies of the free molecules, such as N2, CO2, and H2, were taken from the NIST database. ΔGU = −eU, where e is the transferred charge and U represents the applied potential at the electrode. The limiting potential (UL) was employed to assess the catalytic activity for urea formation, which can be computed by: UL = −max(ΔG1, ΔG2, ΔG3, ΔG4, …, ΔGi)/e, where ΔGi represents the free energy change of each elementary step in the whole electrochemical process. According to this definition, a less negative UL on a given catalyst denotes a less energy input, thus suggesting its higher catalytic activity toward urea formation.
Furthermore, we examined the electronic properties of the MoP-(101) surface by computing the partial density of states (PDOSs) of Mo and P atoms. There are obvious hybridizations between the Mo-4d and P-3p orbitals (Fig. S2†), suggesting their strong chemical bonds and thus ensuring the high stability of this material. According to the Bader population analysis, there is about 0.49e charge transfer from Mo to P (Fig. 1c), making the 4dz2 orbital of Mo atom empty (locating above the Fermi level, Fig. 1d), while most of the other 4d orbitals are occupied (Fig. S2†). Thus, the co-existence of empty and occupied d orbitals in the surface Mo atoms makes it possible for the surface Mo atoms to accept the lone-pair electrons of N2 and CO2, and at the same time donate electrons into anti-bonding orbitals of N2 and CO2, thus leading to an “acceptance–donation” process, which could be conducive to the sufficient activation of the two inert reactants during urea synthesis.
With these three key stages in mind, we first examined the adsorption of N2 and CO2 molecules on a MoP-(101) slab with Mo-termination to check if the first stage of urea production can proceed well. After considering various possible adsorption configurations (end-on and side-on configurations), we found that the N2 molecule prefers to be adsorbed on the bridge sites of two Mo atoms via the side-on pattern, forming two Mo–N bonds with the lengths of 1.98 Å (Fig. 3a). Furthermore, the computed adsorption energy of N2 (ΔEN2) is −1.05 eV, and the corresponding ΔG value is −0.43 eV, suggesting a spontaneous chemisorption process. Due to this strong interaction, the N–N bond length is elongated from 1.10 Å of the free N2 molecule to 1.20 Å of the adsorbed N2. After N2 is strongly adsorbed, the chemisorption of CO2 on its adjacent bridge site is also favorable, with the computed ΔECO2 of −1.09 eV (ΔG is −0.37 eV). In this configuration, both C and O atoms of *CO2 bind with Mo active sites, forming Mo–O and Mo–C bonds with the lengths of 2.18 and 2.08 Å, respectively (Fig. 3b). As a result, the O–C–O bond angle is significantly bent (at about 47°) as compared with the isolated linear CO2 molecule. In addition, the presence of H protons can hinder the adsorption of absorbent on the catalyst; thus, we also examined the adsorption energy of *H species on the MoP-(101) surface. Our results showed that the *H species will be adsorbed on the Mo sites, leading to the formation of three new Mo–H bonds with the adsorption energy of −0.81 eV, which is much lower than that of the adsorbed N2 (−1.05 eV) and CO2 (−1.09 eV), suggesting a more favorable adsorption of CO2 and N2 on the active sites. The aforementioned results undoubtedly suggested that both N2 and CO2 can be strongly chemisorbed and sufficiently activated on the MoP-(101) surface, which can be further verified by the strong hybridization with each other in the computed PDOSs (Fig. S3a†). Furthermore, the MoP-(101) surface transfers significant electrons to the adsorbed N2 (0.71|e|) and CO2 (0.82|e|) molecules due to the “acceptance–donation” process, in which the MoP-(101) surface has empty d orbitals that can accept the lone-pair electrons of CO2 and N2, and simultaneously donate electrons into the anti-bonding orbitals of CO2 and N2. Interestingly, accumulating negative charges (Fig. S3b†) on the absorbates facilitate their further hydrogenation by interacting with the positively charged H+ species.
![]() | ||
Fig. 4 The computed free energy profiles for urea electrosynthesis on Mo-terminated MoP-(101) surface at zero and applied potentials. The structures of the involved reaction intermediates are presented in Fig. S8.† |
After confirming that N2 and CO2 molecules can be well activated upon co-adsorption on Mo-terminated MoP-(101) surface, we explored the feasibility of the C–N coupling reaction. Our computational results showed that the direct coupling between the adsorbed *N2 and *CO2 molecules on MoP-(101) slab is not possible, as they will be spontaneously separated after full atomic relaxation (Fig. S4†). To this end, we further explored the hydrogenation of the adsorbed N2 and CO2 species. Interestingly, after one proton-coupled electron transfer (PCET) step, the O atom of *CO2 prefers to be hydrogenated to *COOH with a ΔG of −0.51 eV, which is much lower than that of *N2H formation (ΔG = −0.06 eV). Subsequently, the approach of a second (H+ + e−) to *COOH generates the key *CO intermediate by releasing a H2O molecule. Remarkably, this elementary step (*N2 + *COOH + H+ + e− → *N2 + *CO + H2O) is uphill by only 0.17 eV in the free energy profile, which is slightly lower than that of the formation of (*N2H + *COOH, 0.21 eV). These computational results suggested that the formation of intermediate species *N2 and *CO is more energetically favorable than that of other species on the MoP-(101) surface.
Next, we explored the possibility of *NCON formation by the C–N coupling reaction on this MoP surface, inspired by the recent pioneering work of Li's group:20,43 the coupling between *CO and *N2 to generate a tower-like *NCON intermediate is a pivotal step in urea synthesis. As shown in Fig. S5,† we found that the reaction of *N2 + *CO → *NCON is endothermic by 0.48 eV, and the involved energy barrier (1.87 eV) is also much higher than those of the reported Pd–Cu catalyst (0.79 eV),20 MBene materials (0.42–1.00 eV),43,44 and double-atom catalysts (1.18–1.62 eV) anchored on N-doped graphene.45 Thus, we predicted that the formation of *NCON through the coupling between the adsorbed *N2 and *CO species on MoP catalyst is unfavorable both thermodynamically and kinetically.
Since the adsorbed *N2 and *CO intermediates cannot couple with each other to generate *NCON species, we examined their further hydrogenation to give (*N2H + *CO), (*N2 + *COH), or (*N2 + *HCO). The computational results showed that the formation of (*N2H + *CO) is energetically more favorable; its ΔG value (0.09 eV) is much lower than the other two hydrogenation steps (1.34 and 0.91 eV, respectively, see Table S1†). In addition, to further demonstrate that N2 is preferentially hydrogenated, the kinetic process of the first step of hydrogenation was also considered. Our results showed that the barrier energy to form *N2H + *CO, *N2 + *COH, and *N2 + *HCO is 1.07, 2.38 and 1.58 eV, respectively, suggesting that the formed *N2H + *CO is kinetically most favorable (Fig. S6†). Subsequently, the formed (*N2H + *CO) is facilely hydrogenated to (*N2H2 + *CO) due to its lower ΔG value (−0.16 eV) than those of other competitive steps, including formation of HNCON (−0.06 eV), *N2H + *CHO (0.94 eV), and *N2H + *COH (1.41 eV). For the (*N2H2 + *CO) intermediate, the *CO can be inserted into the N–N bond, leading to the formation of *NHCOHN*. Notably, in this structure, the N–N bond is broken with a distance of 2.37 Å, accompanied with the formation of two C–N bonds with the lengths of 1.39 Å. Especially, from the thermodynamic perspective, the formation of *NHCOHN* is exothermic, and the ΔG value is −0.25 eV. In contrast, the hydrogenation of (*N2H2 + *CO) is endothermic by 0.16 eV to (*N2H3 + *CO), 0.98 eV to (*N2H2 + *CHO), and 1.45 eV to (*N2H2 + *COH). Kinetically, the computed energy barrier of this coupling process is 1.30 eV (Fig. S7†), which is comparable to those of previously reported catalysts (0.58–1.62 eV)20,43,45 and is even lower than the thermodynamics overpotential value of 1.67 eV.46 Thus, the C–N coupling of *N2H2 and *CO on the MoP-(101) surface is thermodynamically and kinetically feasible, playing a critical role for efficient urea production.
After the generation of the key *NHCOHN* intermediate through C–N coupling, its further hydrogenation on Mo-(101) slab was explored. It was found that *NHCONH is reduced to the urea product via two PCET steps with ΔG values of 0.27 and 0.25 eV, respectively. Finally, the formed urea is easily released from the MoP-(101) surface, as urea has a quite small adsorption energy (−0.28 eV) on this catalyst, much lower than that of the Pd–Cu catalysts (−1.68 eV).20
Overall, the most favorable pathway for urea production from the cooperative reduction of N2 and CO2 on Mo-terminated MoP-(101) surface can be summarized as: (Fig. 4 and S8†). Since the formation of
exhibits the largest ΔG value (0.27 eV) among all the elementary steps, this step is the potential-limiting step, and the corresponding limiting potential for urea production (UureaL) is −0.27 V, which is much less negative than that of the state-of-the-art Pd–Cu catalyst in experiments and those of other theoretically designed electrocatalysts (−0.49 to −0.79 V),20,45 indicating the superior electrocatalytic activity of MoP toward urea formation. Furthermore, to simulate a real aqueous solution, we computed the free energy to generate urea by using an explicit solvent model (Fig. S9†). Our results showed that the potential limiting step changes from formation of *NHCONH2 (vacuum) to the formation of *N2 + *COOH (water), with a small change in the limiting potential. Notably, we also computed the Gibbs free energy of the MoP-(001) and MoP-(100) surface. Our results showed that the limiting potential of MoP-(001) and MoP-(100) is 1.34 and 0.73 eV, respectively; these values are larger than those of the MoP-(101) surface, and thus the MoP-(101) surface has the highest urea activity (Fig. S10†).
Then, we measured the prepared catalyst's performance for the electrochemical synthesis of urea in an H-type two-chamber electrolytic cell with 0.1 M KHCO3 solution. Ultrahigh-purity CO2 (99.999%) and N2 (99.999%) gas mixture was continuously fed into the cathode chamber during the reaction. The liquid product (urea) formed in the electrolyte was determined by the diacetyl monoxime method. The calibration curve is displayed in Fig. S14.† Constant potential electrolysis tests were used to quantitatively evaluate the ability for urea synthesis at different working potentials; the samples were kept for two hours at each potential (Fig. S15†). The experimental results illustrate that the highest faradaic efficiency of 36.5% (Fig. 5e) and the highest urea formation rate of 12.4 μg h−1 mgcat−1 (Fig. 5f) both occur at −0.35 V. Particularly, the achieved FE and yield rate for urea production on MoP nanoparticles are comparable to previously reported catalysts (Table S2†),20,22,46 suggesting their good catalytic performance for urea generation. In addition, we also tested the MoP samples from different producers. Our results showed that the FE and yield of urea are generally comparable, further suggesting that MoP can be an effective electrocatalyst for urea synthesis (Fig. S16†).
Furthermore, to prove that the urea is generated during the simultaneous reduction of N2 and CO2 by MoP as a catalyst, we conducted potentiostatic tests under different conditions. We barely detected urea either using Ar as feed gas or pouring the mixed gas (N2 and CO2) into the electrolyte tank under the condition of open circuit potential. Besides, if we only pump N2 or CO2 into the electrolytic cell, it was challenging to discover urea in the electrolyte after the reaction (Fig. S17†). These results prove that urea was generated from the reduction of CO2 and N2 using MoP as the catalyst. In order to illustrate that MoP is the active substance for catalyzing urea synthesis, we used different electrodes for electrochemical tests, like carbon paper with nothing loaded and MoO2 or MoO3 loaded on carbon paper, under the same conditions; all three electrodes produced much less urea than MoP loaded on carbon paper (Fig. S18–S20†). This indicates that MoP is the catalytically active substance, rather than MoOx or carbon paper. To further explore the stability of the catalyst, we carried out XRD tests on the electrode after reaction. In the XRD pattern (Fig. S21†), it can be observed that the characteristic peaks of MoP after reaction still exist and no other peak appears, which indicates that the MoP was not reduced during the electrochemical reaction. Besides that, the SEM image of MoP after the reaction (Fig. S22†) also proves that the MoP has no significant change in morphology during the reaction. Further, we tested the cyclic stability of MoP, and it can be seen that after five consecutive cycles (Fig. S23†), there was only a slight decrease in FE, but the urea formation rate dropped by nearly 30%, probably because the catalyst fell off the carbon paper during the reaction. In addition, we tested the XRD after reactions at different potentials. Our results showed that the lattice of MoP-(101) facets at different potentials has no significant change (Fig. S24†). These prove that the stability of the catalyst is relatively good. To further determine whether the adsorption of N2 and CO2 is stable at the operating voltage (for details, see Fig. S25†), we computed N2 and CO2 co-adsorption energy on the MoP-(101) surface at −0.35 V, under which the highest faradaic efficiency and the highest urea formation rate were obtained experimentally. Our results showed that the co-adsorption energy of (*N2 + *CO2) on MoP-(101) is −1.15 eV, indicating that the co-adsorption of N2 and CO2 is stable at operating voltages.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ta07531h |
‡ Dongxu Jiao and Yilong Dong contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2023 |