Next Article in Journal
Microwave-Assisted Metal-Organic Frameworks Derived Synthesis of Zn2GeO4 Nanowire Bundles for Lithium-Ion Batteries
Previous Article in Journal
PMMA/SWCNT Composites with Very Low Electrical Percolation Threshold by Direct Incorporation and Masterbatch Dilution and Characterization of Electrical and Thermoelectrical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Single-Atom Anchored g-C3N4 Monolayer as Efficient Catalysts for Nitrogen Reduction Reaction

1
School of Physics, Henan Normal University, Xinxiang 453007, China
2
College of Physics and Electronic Engineering, Zhengzhou Normal University, Zhengzhou 450044, China
3
School of Materials Science and Engineering, Henan Institute of Technology, Xinxiang 453000, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(8), 1433; https://doi.org/10.3390/nano13081433
Submission received: 22 February 2023 / Revised: 9 April 2023 / Accepted: 18 April 2023 / Published: 21 April 2023
(This article belongs to the Topic Condensed Matter Physics and Catalysis)
(This article belongs to the Section Theory and Simulation of Nanostructures)

Abstract

:
Electrochemical N2 reduction reaction (NRR) is a promising approach for NH3 production under mild conditions. Herein, the catalytic performance of 3d transition metal (TM) atoms anchored on s-triazine-based g-C3N4 (TM@g-C3N4) in NRR is systematically investigated by density functional theory (DFT) calculations. Among these TM@g-C3N4 systems, the V@g-C3N4, Cr@g-C3N4, Mn@g-C3N4, Fe@g-C3N4, and Co@g-C3N4 monolayers have lower ΔG(*NNH) values, especially the V@g-C3N4 monolayer has the lowest limiting potential of −0.60 V and the corresponding limiting-potential steps are * N 2 + H + + e = * N N H for both alternating and distal mechanisms. For V@g-C3N4, the transferred charge and spin moment contributed by the anchored V atom activate N2 molecule. The metal conductivity of V@g-C3N4 provides an effective guarantee for charge transfer between adsorbates and V atom during N2 reduction reaction. After N2 adsorption, the p-d orbital hybridization of *N2 and V atoms can provide or receive electrons for the intermediate products, which makes the reduction process follow acceptance-donation mechanism. The results provide an important reference to design high efficiency single atom catalysts (SACs) for N2 reduction.

1. Introduction

Ammonia (NH3) is not only an essential substance to produce fertilizers, explosives, dyes, and pharmaceuticals, but also an important clean energy carrier [1,2,3,4]. The growing demand of NH3 has spurred researchers to seek more efficient artificial nitrogen (N2) fixation. Presently, large-scale production of NH3 mainly depends on the Haber–Bosch process. However, the process not only needs to operate under harsh conditions (300–500 °C and 200–300 atm) [5], but also requires huge energy input, and simultaneously generates a large number of greenhouse gases carbon dioxide (CO2) [6,7,8]. Considering energy consumption and environmental protection, researchers hope to find a method for converting N2 to NH3 using renewable energy, no polluting emissions, and mild operating conditions.
Electrochemical N2 reduction reaction (NRR), a promising approach for sustainable NH3 production under ambient conditions, has received extensive and increasing attention both in experiment and theoretic studies [9,10,11]. However, the strong bonding energy of N≡N triple bond (945 KJ/mol) and the weak adsorption of nonpolarized N2 are two major challenges in the NRR process [12,13,14]. Therefore, an efficient catalyst for N2 activation and reduction is urgent to improve NRR activity. On the other hand, the transition metal can accept the lone pair electrons of the N2 molecule and weak the N≡N triple bond [15,16,17]. Dispersing metal atoms on suitable supporting materials can not only provide more active sites, but also regulate the electronic properties of substrate and enhance the catalytic efficiency. Two-dimensional materials (2D) are widely used as catalyst substrates for N2 reduction due to their excellent chemical stability, thermal and electrical properties, and the ability to construct defect active sites through surface functionalization [18,19,20,21].
In recent years, transition metal single atom catalysts (SACs) constructed by anchoring single atom to two-dimensional materials have gained increasing attentions for the electrocatalysis applications and have shown excellent catalytic performances in NRR [22,23,24,25]. To avoid diffusion and agglomeration of metal atoms on the substrate, vacancies are usually constructed in the substrate to increase the stability of TM atom, or a two-dimensional material with a pore structure is used as the supporting substrate [26,27]. For example, the anchored Mo and Au on the N-doped porous carbon exhibited excellent NRR catalytic performance [28]. The Mo-embedded C2N monolayer has been demonstrated a very high NRR catalytic activity with the onset potential (Uonset) of 0.17 V by Zhao et al. [29]. Thus, selecting the right metal atom and anchoring it to a suitable substrate with a pore structure can not only prevent the diffusion and agglomeration of the metal atom, but also higher catalytic activity can be obtained by the d-orbital electrons of metal atom inducing the π-back donation.
Graphitic carbon nitride (g-C3N4) has recently attracted great attentions because of its good physicochemical stability and superior properties [30,31]. tri-s-triazine-based g-C3N4 with large pore and s-triazine-based g-C3N4 with small pore are two common structures [32]. The uniform distribution of pore sites in both structures provide uniform nitrogen coordinators to capture metal atoms. Many studies [33,34] have shown that metal atoms anchored at tri-s-triazine-based g-C3N4 have excellent NRR performance. For example, Zhao et al. have found W-anchored tri-s-triazine-based g- C3N4 exhibits a high catalytic activity toward NRR with a limiting potential of −0.35 V [33]. Wang et al. have found that B/g-C3N4 can reduce N2 to NH3 with a lower onset potential (0.20 V) [34]. For s-triazine-based g-C3N4, it also has a pore structure surrounded by three nitrogen atoms and can be used as a metal-free substrate to anchor metal atoms. Hu et al. [35] have found that V@g-C3N4 with lying-on adsorbed N2 pattern has the lowest limiting potential of −0.79 V. However, implicit solvation model was not used to simulate the electrolyte solution, and the origin of the catalytic activity has not been explored in detail. It is necessary to systematically study the N2 reduction process in the electrolyte solution and explore the origin of the catalytic activity by anchoring single TM atom on s-triazine-based g-C3N4 (TM@g-C3N4).
Inspired by the above studies, the NRR catalytic behaviors of TM (TM = Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu) atoms on intrinsic s-triazine-based g-C3N4 have been systematically investigated by density functional theory (DFT) theory. Theoretical calculations show that TM (V, Cr, Mn, Fe, Co) atoms anchored at g-C3N4 exhibits better NRR activities than Ru (0001) surface. V@g-C3N4 monolayer possesses the lowest limiting potential of −0.60 V, which is lower than that calculated by Hu et al. [35], and potential-limiting step (PLS) is *N2-*NNH for both alternating and distal mechanisms. The charge density difference, spin electrons distribution, density of states (DOS), the variation of charge transfer, DV-N, LN-N, and Mtot are used to explore the origin of the excellent catalytic performance. These results provide an important reference for the research of s-triazine-based g-C3N4 SACs.

2. Computational Methods

DFT methods are performed using the Vienna Ab Initio Simulation Package (5.4.4, 2017, VASP Software GmbH, Vienna, Austria) [36,37]. The exchange-correlation potentials are described through the Perdew–Burke– Ernerhof (PBE) parametrization within the generalized gradient approximation (GGA) [38,39]. The DFT-D3 method is utilized to describe the weak van der Waals (vdW)-like interaction [40]. The kinetic energy cutoff of plane-wave expansion is set to 520 eV and a vacuum space of 20 Å is inserted in the z-direction to eliminate the interaction between adjacent periodic units [41]. A 3 × 3 × 1 supercell is constructed. For geometrical optimizations, a 1 × 1 × 1 Monkhorst–Pack mesh of k-points is employed to sample the first Brillouin zero. A finer 5 × 5 × 1 k-points grid is chosen for the density of states (DOS). The convergence tolerances for the force and energy are set to 0.02 eV/Å and 10−5 eV, respectively. Bader charge is used analyze electron transfer [42].
The adsorption energies (Eads) of single TM atom and N2 molecule on substrate are calculated by the following formula:
E a d s = E a d s o r b a t e @ s u b s t r a t e E s u b s t r a t e E a d s o r b a t e
where E a d s o r b a t e @ s u b s t r a t e , E s u b s t r a t e , and E a d s o r b a t e are the total energies of adsorbate @substrate, substrate, and an isolate adsorbate, respectively. Negative Eads indicates that the adsorbate can be stably bond to the substrate.
The cohesive energy (Ecoh) is calculated by the following formula:
E c o h = E T M E T M , b u l k
where E T M and E T M , b u l k are the energies of an isolated TM atom and bulk metal, respectively.
Computational hydrogen electrode (CHE) mode [43] is used to calculate the reaction free energy (ΔG). A solvation model is used to simulate the solution environment [44], and a relative permittivity of 80 is set for water [45]. The ΔG is evaluated as follows:
Δ G = Δ E + Δ Z P E T Δ S + Δ G U + Δ G P H
where Δ E is the reaction energy directly obtained by DFT calculations, Δ Z P E and Δ S are the changes of zero-point energy (ZPE) and entropy computed from the vibrational frequencies at 298.15 K. Δ G U is the free energy contribution with applied potential U, ΔGU = eU, where e and U are the number of electrons transferred and the applied electrode potential. Δ G p H = k B × p H × l n 10 , where pH = 0 in this work for simplicity. The limiting potential (UL) is defined as U L = Δ G m a x / e .
Using the CHE model, the lowest G p H , U was evaluated. The G ( p H , U ) values can be projected onto the ( p H , U ) plane of lowest Gibbs free energy, resulting in a 2D plot with pH and U descriptors to build the Pourbaix diagram [46]. The Pourbaix diagrams can provide the reliability electrocatalytic reaction results under certain conditions of pH and potential U.

3. Results and Discussion

3.1. Structural and Electronic Properties of TM@g-C3N4

Optimized structure of s-triazine-based g-C3N4 monolayer is displayed in Figure 1a. It contains two C-N bonds, the bond lengths dC-N are 1.33 Å and 1.46 Å, respectively. The optimized lattice constant is g-C3N4 is 4.78 Å, which is consistent with previous theoretical results [47,48]. The symmetrical distribution of spin-up and spin-down electrons in Figure 1b indicates that the g-C3N4 is a non-magnetic semiconductor. TM (Sc-Cu) atom is anchored at five high-symmetry sites (Figure 1a) to investigate the stable configurations, namely, the top sites of C (T1), pyridine N (T2), and graphite N (T3), hollow sites above the three pyridine N (H1) and the triazine ring (H2). Taking Mn atom as an example, the calculated adsorption energies (Eads) are −5.16 eV, −6.01 eV, −4.49 eV, −6.20 eV, −3.70 eV, respectively, and the H1 site is the most stable adsorption site with the lowest Eads −6.20 eV. All the TM atoms have a same stable adsorption site H1. The stable configuration of Mn atom adsorbed at H1 site is shown in Figure 1c. The adsorption of Mn atom causes the distortion of g-C3N4 monolayer, but no bond breaks. The other configurations of TM@g-C3N4 are displayed in Figure S1.
The calculated adsorption energies (Eads), energy difference between adsorption energy and cohesive energy (ΔE), total magnetic moments (Mtot), the charge (ΔQ) transferred from TM atoms to g-C3N4 monolayer, and the electronic structures (ES) of TM@g-C3N4 are shown in Table 1. The premise of good catalytic performance of SAC is that single metal atom has suitable Eads on the substrate. The calculated Eads values of TM@g-C3N4 are in the range of −9.55~−4.76 eV, which indicates that TM atoms can bind stably to the substrate. At the same time, the energy difference (ΔE) between the adsorption energy (Eads) and the cohesive energy (Ecoh) is calculated to investigate the aggregation of TM atoms on g-C3N4. The value of ΔE is in the range of −5.36~−0.92 eV, and a negative value of ΔE means that the adsorption of metal atoms on g-C3N4 is stronger than the cohesive of atoms. Hence, the aggregation of TM atoms on g-C3N4 can be suppressed efficiently. TM (Sc-Cu) atoms anchored to g-C3N4 holes have good stability.
The spin densities distribution of TM@g-C3N4 are shown in Figure S2. It can be seen that spin electrons are located at TM (TM= Sc, Ti, V, Cr, Mn, Fe, Co, Cu) and surrounding atoms. Compared with Figure 1b, it is obvious that the anchored TM atoms (TM= Sc, Ti, V, Cr, Mn, Fe, Co, Cu) induce the magnetic moment of TM@g-C3N4. V@g-C3N4 has the most spin densities, which is consistent with the largest moment (4.88 μB) in Table 1.
As is well known, the anchored TM atoms can regulate the electronic structure of substrate, and the catalytic performance of the catalyst is closely related to the electronic structure. Hence, the total density of states (TDOS) of TM@g-C3N4 and partial DOS (PDOS) are calculated and shown in Figure S3. It is seen that the anchored TM (TM=Sc, Ti, V, Cr, Mn, Fe, Co, Cu) atoms induce the asymmetry of spin-up and spin-down electrons of TM@g-C3N4, which is consistent with the spin electrons distribution in Figure S2. Second, the adsorption of Cr, Mn, Co, and Ni atoms retain the semiconductor property, but the band gap decreases from 1.57 eV to 0.38 eV, 0.46 eV, 0.58 eV, and 1.23 eV, respectively. The TDOSs of Ti@g-C3N4, V@g-C3N4, and Fe@g-C3N4 exhibit metallic properties and many electronic states are located near the Fermi level, which ensures a rapid transfer of charge in the reaction.

3.2. Adsorption of N2 Molecule

It is well known that the N2 adsorption on the catalytic is the first step to investigate the NRR performance, which reflects the sensitivity of the catalyst and determines the most favorable adsorption manner. Stable adsorption of N2 is a prerequisite for the subsequent reaction to generate NH3.
The adsorption energies [Eads(*N2)] of TM@g-C3N4 (TM = Sc-Cu) are illustrated in Figure 2a. For TM@g-C3N4 (TM = Ti, V, Cr, Mn, Fe, Co, Ni, Cu), both end-on and side-on configurations can exist, while Eads(*N2) (−1.76~−0.58 eV) of end-on adsorption N2 molecule are lower than those (−1.45~−0.26 eV) of side on patterns, which means that end-on patterns are more energetically favorable than the side-on patterns. Therefore, in the following studies of this paper, only the favorable end-on adsorption manner is discussed. These results indicate that the TM atoms (TM = Ti, V, Cr, Mn, Fe, Co, Ni, Cu) anchored at g-C3N4 are active sites and have strong ability to capture N2 molecules. This result is consistent with the result calculated by Hu et al. [35] except for Ti atom.
In order to further study the adsorption of N2 on TM@g-C3N4, the distance between TM atom and the N2 molecule (DTM-N), and N-N bond length (LN-N) are displayed in Figure 2b. The DTM-N (TM = Ti, V, Cr, Mn, Fe, Co, Ni, Cu) are in the range of 1.76~2.09 Å and the LN-N are stretched to 1.13~1.14 Å, which indicates that the N≡N triple bonds are weakened and good for reduction reaction. For Sc@g-C3N4, the adsorption energies for end-on and side-on patterns are −0.01 eV and 0.52 eV, the DSc-N is 4.14 Å and the LN-N is 1.11 Å, which means that Sc@g-C3N4 has poor NRR activity. TM@g-C3N4 (TM = Ti, V, Cr, Mn, Fe, Co, Ni, Cu) will be considered as SAC candidates in the following studies.

3.3. N2 Electrocatalytic Reduction Reaction

The second criterion for good NRR catalyst is that it has a lower limiting potential in the N2 reduction reaction. To evaluate the catalytic activities of TM@g-C3N4 (TM = Ti, V, Cr, Mn, Fe, Co, Ni, Cu), the NRR mechanisms are investigated to obtain the potential-limiting step (PLS) and limiting potential.
Theoretical and experimental studies [49] on NRR show that the end-on adsorbed N2 molecule on TM@g-C3N4 (TM = Ti, V, Cr, Mn, Fe, Co, Ni, Cu) can be reduced to NH3 via a distal or alternating mechanism. In the distal mechanism, the proton (H+) continuously attacks the terminal N atom until the first NH3 molecule is released and then the other. While for the alternating mechanism, the proton (H+) alternating bonds the two N atoms to release NH3 molecule.
Previous studies suggest that the first H+ reaction * N 2 + H + + e = * N N H is usually the potential-limiting step (PLS) for NRR [16,50]. The free energy barrier ΔG(*NNH) is calculated and displayed in Figure 3a. It can be seen that the ΔG(*NNH) values of V@g-C3N4 (0.60 eV), Cr@g-C3N4 (0.76 eV), Mn@g-C3N4 (0.78 eV), Fe@g-C3N4 (0.63 eV), and Co@g-C3N4 (0.71 eV) are lower than that on Ru (0001) surface. On the other hand, since the hydrogen evolution reaction (HER) is a major competitor for NRR [17,51], the calculated free energies change of H [ΔG(*H)] are shown in Figure 3b. It is seen that ΔG(*H) are positive for V@g-C3N4, Cr@g-C3N4, and Mn@g-C3N4, which means that H atom cannot be adsorbed on V, Cr and Mn atoms. For Fe@g-C3N4 and Co@g-C3N4, the ΔG(*H) is negative, while the values of ΔG(*N2) are lower than ΔG(*H) and they prefer to adsorb N2 molecule. In conclusion, V, Cr, Mn, Fe, and Co atoms anchored by g-C3N4 are selected as potential SACs for NRR by the second criterion.
In order to explore the whole N2 reduction process, it is necessary to explore the PLS of the process. The PLS dominates the catalytic efficiency of the catalyst for NRR. The results show that V@g-C3N4 is the best catalyst with the lowest ΔG(*NNH) value 0.60 eV compared with Cr@g-C3N4, Mn@g-C3N4, Fe@g-C3N4, and Co@g-C3N4. The whole free energy diagrams of V@g-C3N4 during NRR process is further investigated with the favorable end-on adsorption manner.
The NRR free energy diagrams of V@g-C3N4 are displayed in Figure 4. Optimized adsorption structures of intermediate via the alternating and distal pathways are depicted in Figure S4. The free energy diagrams via alternating and distal mechanisms are very similar. During N2 reduction process, the N2 adsorption and protonation into *NNH along the two mechanisms are same. The adsorption process of * + N 2 = * N 2 is downhill with ΔG value −0.89 eV, which shows that V@g-C3N4 can spontaneously adsorb the N2 molecule. The step of * N 2 + H + + e = * N N H is uphill with ΔG value 0.60 eV, which means that the protonation process is endothermic. The second proton-electron may bond with another N atom of *N2 molecule to generate *NHNH (alternating mechanism) or the same N atom (distal mechanism) to generate *NNH2, and the hydrogenation in both ways are both uphill with ΔG value 0.48 eV and 0.13 eV, respectively. In the following steps of two mechanisms, ΔG values are all negative. For alternating mechanism in Figure 4a, the calculated ΔG values are −0.89, 0.60, 0.48, −0.44, −0.12, −1.81 eV, and −0.27 eV, respectively. The first protonation step has the highest ΔG values and is confirmed to be the PLS with an uphill value 0.60 eV. When the UL is applied to −0.60 V, all the reaction steps are exothermic, therefore, the limiting potential is −0.60 V. For distal mechanism in Figure 4b, the ΔG values are −0.89, 0.60, 0.13, −0.77, −0.55, −0.70, and −0.27 eV, respectively, and it is obvious that it has the same PLS and UL as the alternating mechanism. The calculated limiting potentials of V@g-C3N4 are lower than those calculated by Hu et al., where the limiting potential of V@g-C3N4 with lying-on adsorbed N2 is −0.79 V, while that of standing-on pattern is −0.93 V [35]. The final desorption of the second *NH3 molecule requires the absorption energy of 1.16 eV, and it can also be further protonated to form NH4+ [52].
Pourbaix diagrams can provide an effective guidance for the electrocatalytic reactions, which will not be discussed in this article.

3.4. Origin of Catalytic Activity

To explore the excellent catalytic efficiency of V@g-C3N4, the electronic properties, charge density difference, and spin electrons distribution are investigated. The charge density difference is shown in Figure 5a. It can be seen that the electron densities of the three N atoms anchoring the V atom increases, and Bader charge analysis finds that the V atom loses 1.92 e, which makes the V atom be a good active site for trapping N2 molecule. Figure 5b shows that the spin density is mainly located at the V atom, which makes V@g-C3N4 possess a total spin moment of 4.88 μB. These results prove the conclusion that the magnetism of the catalyst can increase its activity. The density of states (DOS) is presented in Figure 5c. Different from the intrinsic s-triazine-based g-C3N4 monolayer of direct bandgap semiconductor in Figure 1b [47,48], V@g-C3N4 exhibits a metallic property owing to the V atom. The spin moment is vital to activate N2 molecule, and the excellent electrical conductivity is essential to ensure good charge transfer for efficient electroreduction reaction [48,53]. Both of them play an important role in the outstanding NRR catalytic performance of V@g-C3N4.
In order to further explore the understanding NRR catalytic performance of V@g-C3N4, the charge density difference and spin electrons distribution after N2 adsorption are calculated. As shown in Figure 6a, the electrons accumulate near the *N2 molecule, and it is confirmed by Bader charge analysis that 0.42 e transferred from V atom to the N2 molecule. The spin electrons distribution in Figure 6b reveals that after N2 adsorption, some spin moment on V atom can be transferred to the N2 molecule, so that the N2/V@g-C3N4 only possesses a magnetic moment 1.00 μB. The change in charge and spin magnetic moment drives its further activation and reactions. The DOS of N2/V@g-C3N4 are displayed in Figure 6c. Obvious overlap between the *N2-2p and V-3d orbitals around the Fermi energy happens, in which the empty d orbitals of V can accept the lone-pair electrons in N2, at the same time, the occupied d orbitals can donate electrons to the antibonding orbitals of N2, and this process follows acceptance-donation mechanism.
To further explore the charge transferred among intermediate adsorbent, the variations of the charge transferred between them are studied. Based on the previous studies [54,55], NxHy/V@g-C3N4 is divided into three parts (Figure S5): g-C3N4 substrate without V and the three N atoms bonded with it (moiety 1), V-3N (moiety 2), and the adsorbed intermediate NxHy (moiety 3). The charge variations along distal and alternating mechanisms are shown in Figure 7 (the charge variation is the charge difference of each moiety between the present step and the previous step). For N2 adsorption on V@g-C3N4 monolayer in Figure 7a, V-3N offers 0.42 e and 0.45 e to *N2 and g-C3N4, respectively. During the protonation process * N 2 + H + + e = * N N H , V-3N and g-C3N4 substrate provide 0.15 e and 0.19 e, respectively. However, during the second protonation process * N N H + H + + e = * N N H 2 , V-3N gains 0.52 e from the substrate and *NNH2. In the following hydrogenation and reduction steps, obvious charge fluctuation occurs for the three moieties, which means that V-3N is the charge transferred medium between NxHy and g-C3N4 substrate, providing or receiving electrons to the adsorbed intermediate. The above results once again prove that the excellent catalytic performance of V@g-C3N4 in NRR process follows the acceptance-donation mechanism. The charge variation in the alternating mechanism shown in Figure 7b is similar to those the distal mechanism.
In order to explore the change of adsorbates configuration and magnetic properties during the reduction of N2, the V-N distance (DV-N), N-N bond length (LN-N), and the total magnetic moments (Mtot) are examined and displayed in Figure 8. In Figure 8a, DV-N along distal pathway decreases during protonation and reaches a minimum when the first NH3 molecule is released, and then continuously increases. While in alternating mechanism, when the proton (H+) attacks the distal N atom, DV-N decreases, and when it attacks the N atom bonded to V, DV-N increases. DV-N increases in a zigzagging fluctuation way during the whole reduction process. For LN-N, as shown in Figure 8b, it is stretched continuously until the first NH3 is released for both distal and alternating mechanisms. Figure 8c shows the magnetic moments of the NxHy, and it can be seen that the protonation process makes the Mtot show zigzag fluctuations for both mechanisms. These results provide strong evidence for the excellent catalytic performance of V@g-C3N4.

4. Conclusions

In this paper, the catalytic performance of 3d TM atoms anchored on s-triazine-based g-C3N4 (TM@g-C3N4) for N2 reduction is systematically investigated. The results show that N2 molecules can be stably adsorbed on TM@g-C3N4 (TM=Ti, V, Cr, Mn, Fe, Co, Ni, Cu), and the end-on modes are more energetically favorable than the side-on modes. The ΔG(*NNH) values indicate that TM@g-C3N4 (TM = V, Cr, Mn, Fe, Co) are potential NRR candidates with lower ΔG(*NNH) value than that on Ru (0001) surface. The free energy diagrams show that V@g-C3N4 exhibits the highly catalytic performance for NRR with a limiting potential −0.60 V, and the PLS is the first protonation of *N2 for both alternating and distal mechanisms. In order to explore the origin of excellent NRR performance of V@g-C3N4, the charge density difference, spin electrons distribution and DOS plots before and after N2 adsorption, and their corresponding charge variation of adsorbates during N2 reduction reaction are discussed. First, for V@g-C3N4, V atom as an active site provide 1.92 e and has the ability to trap N2 molecule. A spin moment of 4.88 μB of V@g-C3N4 contributed by V atom can activate N2 molecule. Second, after N2 adsorption, the strong interaction between *N2-2p and V-3d orbitals ensures the charge transferred during N2 reduction reaction. Third, the charge variation of moieties indicate that V-3N acts as a medium between adsorbates NxHy and g-C3N4 substrate, which ensures the efficient reduction of N2. The DV-N, LN-N and Mtot of NxHy/V@g-C3N4 are calculated to illustrate the change of adsorbate configurations and magnetic properties during the N2 reduction reactions, which provides an important evidence for the excellent catalytic performance of SACs TM@g-C3N4.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13081433/s1, Figure S1: Stable configurations of TM@ g-C3N4. Figure S2: Spin electrons distribution of TM@ g-C3N4. Figure S3: DOS of TM@ g-C3N4. Figure S4 Optimized adsorption structures of intermediate via the alternating and distal mechanisms. Figure S5. Diagram of the moiety 1(g-C3N4), moiety 2 (V-3N) and moiety 3 (the adsorbed intermediate NxHy).

Author Contributions

H.C. and W.C.: investigation, writing—original draft, project administration; Z.F. and Y.L.: data curation, formal analysis; M.Z. and J.S.: software, validation, resources, methodology; Y.T. and X.D.: conceptualization, supervision, writing—review and editing and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

The National Natural Science Foundation of China (Nos. 62074053, 12204430, 12204431), the Natural Science Foundation of Henan (Grant No. 222300420587).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on a reasonable request from the corresponding author.

Acknowledgments

This work is financially supported by the National Natural Science Foundation of China (Nos. 62074053, 12204430, 12204431), the Natural Science Foundation of Henan (Grant No. 222300420587), the Key Scientific Research Project of Henan College (20A140030), the Aid program for Science and Technology Innovative Research Team, and Open Research Fund of Zhengzhou Normal University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Erisman, J.W.; Sutton, M.A.; Galloway, J.; Klimont, Z.; Winiwarter, W. How a century of ammonia synthesis changed the world. Nat. Geosci. 2008, 1, 636–639. [Google Scholar]
  2. Schlögl, R. Catalytic synthesis of ammonia—A “never-ending story”? Angew. Chem. Int. Ed. 2003, 42, 2004–2008. [Google Scholar] [CrossRef]
  3. Galloway, J.N.; Townsend, A.R.; Erisman, J.W.; Bekunda, M.; Cai, Z.; Freney, J.R.; Martinelli, L.A.; Seitzinger, S.P.; Sutton, M.A. Transformation of the nitrogen cycle: Recent trends, questions, and potential solutions. Science 2008, 320, 889–892. [Google Scholar] [CrossRef] [PubMed]
  4. Canfield, D.E.; Glazer, A.N.; Falkowski, P.G. The evolution and future of Earth’s nitrogen cycle. Science. 2010, 330, 192–196. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, H. Ammonia synthesis catalyst 100 years: Practice, enlightenment and challenge. Chin. J. Catal. 2014, 35, 1619–1640. [Google Scholar] [CrossRef]
  6. Kitano, M.; Kanbara, S.; Inoue, Y.; Kuganathan, N.; Sushko, P.V.; Yokoyama, T.; Hara, M.; Hosono, H. Electride support boosts nitrogen dissociation over ruthenium catalyst and shifts the bottleneck in ammonia synthesis. Nat. Commun. 2015, 6, 6731. [Google Scholar] [CrossRef]
  7. Chen, P. Across the board: Ping chen. ChemSusChem 2018, 11, 2469–2471. [Google Scholar] [CrossRef]
  8. McEnaney, J.M.; Singh, A.R.; Schwalbe, J.A.; Kibsgaard, J.; Lin, J.C.; Cargnello, M.; Jaramillo, T.F.; Nørskov, J.K. Ammonia synthesis from N2 and H2O using a lithium cycling electrification strategy at atmospheric pressure. Energy Environ. Sci. 2017, 10, 1621–1630. [Google Scholar] [CrossRef]
  9. Hao, Y.C.; Guo, Y.; Chen, L.; Shu, M.; Wang, X.Y.; Bu, T.; Gao, W.; Zhang, N.; Su, X.; Feng, X.; et al. Promoting nitrogen electroreduction to ammonia with bismuth nanocrystals and potassium cations in water. Nat. Catal. 2019, 2, 448–456. [Google Scholar] [CrossRef]
  10. Guo, X.; Gu, J.; Hu, X.; Zhang, S.; Chen, Z.; Huang, S. Coordination tailoring towards efficient single-atom catalysts for N2 fixation: A case study of iron-nitrogen-carbon (Fe@N-C) systems. Catal. Today 2020, 350, 91–99. [Google Scholar] [CrossRef]
  11. Xiao, B.B.; Yang, L.; Yu, L.B.; Song, E.H.; Jiang, Q. The VN3 embedded graphane with the improved selectivity for nitrogen fixation. Appl. Surf. Sci. 2020, 513, 145855. [Google Scholar] [CrossRef]
  12. Yang, C.; Zhu, Y.; Liu, J.; Qin, Y.; Wang, H.; Liu, H.; Chen, Y.; Zhang, Z.; Hu, W. Defect engineering for electrochemical nitrogen reduction reaction to ammonia. Nano Energy. 2020, 77, 105126. [Google Scholar] [CrossRef]
  13. Chen, C.; Liang, C.; Xu, J.; Wei, J.; Li, X.; Zheng, Y.; Li, J.; Tang, H.; Li, J. Size-dependent electrochemical nitrogen reduction catalyzed by monodisperse Au nanoparticles. Electrochim. Acta. 2020, 335, 135708. [Google Scholar] [CrossRef]
  14. Wu, T.; Zhu, X.; Xing, Z.; Mou, S.; Li, C.; Qiao, Y.; Liu, Q.; Luo, Y.; Shi, X.; Zhang, Y.; et al. Greatly improving electrochemical N2 reduction over TiO2 nanoparticles by iron doping. Angew. Chem. Int. Ed. 2019, 58, 18449–18453. [Google Scholar] [CrossRef] [PubMed]
  15. Honkala, K.; Hellman, A.; Remediakis, I.N.; Logadottir, A.; Carlsson, A.; Dahl, S.; Christensen, C.H.; Nørskov, J.K. Ammonia synthesis from first-principles calculations. Science 2005, 307, 555–558. [Google Scholar] [CrossRef]
  16. Skulason, E.; Bligaard, T.; Gudmundsdóttir, S.; Studt, F.; Rossmeisl, J.; Abild-Pedersen, F.; Vegge, T.; Jo, H.; Nørskov, J.K. A theoretical evaluation of possible transition metal electro-catalysts for N2 reduction. Phys. Chem. Chem. Phys. 2012, 14, 1235–1245. [Google Scholar] [CrossRef]
  17. Wan, Y.; Xu, J.; Lv, R. Heterogeneous electrocatalysts design for nitrogen reduction reaction under ambient conditions. Mater. Today 2019, 27, 69–90. [Google Scholar] [CrossRef]
  18. Jiao, D.; Liu, Y.; Cai, Q.; Zhao, J. Coordination tunes the activity and selectivity of the nitrogen reduction reaction on single-atom iron catalysts: A computational study. J. Mater. Chem. A 2021, 9, 1240–1251. [Google Scholar] [CrossRef]
  19. Li, H.; Zhao, Z.; Cai, Q.; Yin, L.; Zhao, J. Nitrogen electroreduction performance of transition metal dimers embedded into N-doped graphene: A theoretical prediction. J. Mater. Chem. A 2020, 8, 4533–4543. [Google Scholar] [CrossRef]
  20. Yang, W.; Huang, H.; Ding, X.; Ding, Z.; Wu, C.; Gates, I.; Gao, Z. Theoretical study on double-atom catalysts supported with graphene for electroreduction of nitrogen into ammonia. Electrochim. Acta. 2020, 335, 135667. [Google Scholar] [CrossRef]
  21. Wu, D.; Lv, P.; Wu, J.; He, B.; Li, X.; Chu, K.; Jia, Y.; Ma, D. Catalytic active centers beyond transition metals: Atomically dispersed alkaline-earth metals for the electroreduction of nitrate to ammonia. J. Mater. Chem. A 2023, 11, 1817–1828. [Google Scholar] [CrossRef]
  22. Liu, K.; Fu, J.; Zhu, L.; Zhang, X.; Li, H.; Liu, H.; Hu, J.; Liu, M. Single-atom transition metals supported on black phosphorene for electrochemical nitrogen reduction. Nanoscale 2020, 12, 4903–4908. [Google Scholar] [CrossRef] [PubMed]
  23. Han, M.; Wang, G.; Zhang, H.; Zhao, H. Theoretical study of single transition metal atom modified MoP as a nitrogen reduction electrocatalyst. Phys. Chem. Chem. Phys. 2019, 21, 5950–5955. [Google Scholar] [CrossRef] [PubMed]
  24. Ma, D.; Zeng, Z.; Liu, L.; Huang, X.; Jia, Y. Computational evaluation of electrocatalytic nitrogen reduction on TM single-, double-, and triple-atom catalysts (TM = Mn, Fe, Co, Ni) based on graphdiyne monolayers. J. Phys. Chem. C 2019, 123, 19066–19076. [Google Scholar] [CrossRef]
  25. Ou, P.; Zhou, X.; Meng, F.; Chen, C.; Chen, Y.; Song, J. Single molybdenum center supported on N-doped black phosphorus as an efficient electrocatalyst for nitrogen fixation. Nanoscale. 2019, 11, 13600–13611. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, B.; Fan, T.; Xie, N.; Nie, G.; Zhang, H. Versatile applications of metal single-atom @ 2D material nanoplatforms. Adv. Sci. 2019, 1901787. [Google Scholar] [CrossRef] [PubMed]
  27. Zhu, Y.; Peng, W.; Li, Y.; Zhang, G.; Zhang, F.; Fan, X. Modulating the electronic structure of single-atom catalysts on 2D nanomaterials for enhanced electrocatalytic performance. Small Methods 2019, 1800438. [Google Scholar] [CrossRef]
  28. Han, L.; Liu, X.; Chen, J.; Lin, R..; Liu, H.; Xin, H. Atomically dispersed molybdenum catalysts for efficient ambient nitrogen fixation. Angew. Chem. 2019, 131, 2343–2347. [Google Scholar] [CrossRef]
  29. Wang, Z.; Yu, Z.; Zhao, J. Computational screening of a single transition metal atom supported on the C2N monolayer for electrochemical ammonia synthesis. Phys. Chem. Chem. Phys. 2018, 20, 12835–12844. [Google Scholar] [CrossRef]
  30. Zhao, Z.; Sun, Y.; Dong, F. Graphitic carbon nitride based nanocomposites: A review. Nanoscale 2015, 7, 15–37. [Google Scholar] [CrossRef]
  31. Liu, J.; Cheng, B.; Yu, J. A new understanding of the photocatalytic mechanism of the direct Z-scheme g-C3N4/TiO2 heterostructure. Phys. Chem. Chem. Phys. 2016, 18, 31175–31183. [Google Scholar] [CrossRef] [PubMed]
  32. Zhu, B.; Cheng, B.; Zhang, L.; Yu, J. Review on DFT calculation of s-triazine-based carbon nitride. Carbon Energy 2019, 1, 32–56. [Google Scholar] [CrossRef]
  33. Chen, Z.; Zhao, J.; Cabrera, C.; Chen, Z. Computational screening of efficient single-atom catalysts based on graphitic carbon nitride (g-C3N4) for nitrogen electroreduction. Small Methods 2019, 3, 1800368. [Google Scholar] [CrossRef]
  34. Ling, C.; Niu, X.; Li, Q.; Du, A.; Wang, J. Metal-free single atom catalyst for N2 fixation driven by visible light. J. Am. Chem. Soc. 2018, 140, 14161–14168. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, L.; Zhao, W.; Zhang, W.; Chen, J.; Hu, Z. gt−C3N4 coordinated single atom as an efficient electrocatalyst for nitrogen reduction reaction. Nano. Res. 2019, 12, 1181–1186. [Google Scholar] [CrossRef]
  36. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  37. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  38. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  39. Perdew, J.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  40. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  41. Bermudez, A.; Jelezko, F.; Plenio, M.; Retzker, A. Electron-mediated nuclear-spin interactions between distant nitrogen-vacancy centers. Phys. Rev. Lett. 2011, 107, 150503–150508. [Google Scholar] [CrossRef] [PubMed]
  42. Chai, H.; Chen, W.; Li, Y.; Tang, Y.; Dai, X. The adsorption properties and stable configurations of hydroxyl groups at Mo edge of MoS2 (100) surface. Mater. Chem. Phys. 2022, 283, 126051–126057. [Google Scholar] [CrossRef]
  43. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  44. Mathew, K.; Sundararaman, R.; Letchworth-Weaver, K.; Arias, T.A.; Hennig, R.G. Implicit solvation model for density-functional study of nanocrystal surfaces and reaction pathways. J. Chem. Phys. 2014, 140, 084106. [Google Scholar] [CrossRef]
  45. Reda, M.; Hansen, H.A.; Vegge, T. DFT study of stabilization effects on N-doped graphene for ORR catalysis. Catal. Today 2018, 312, 118–125. [Google Scholar] [CrossRef]
  46. López, M.; Exner, K.S.; Viñes, F.; Illas, F. Computational Pourbaix diagrams for MXenes: A key ingredient toward proper theoretical electrocatalytic studies. Adv. Theory. Simul. 2022, 2200217. [Google Scholar] [CrossRef]
  47. Zhu, B.; Wageh, S.; Al-Ghamdi, A.; Yang, S.; Tian, Z.; Yu, J. Adsorption of CO2, O2, NO and CO on s-triazine-based g-C3N4 surface. Catal. Today 2019, 335, 117–127. [Google Scholar] [CrossRef]
  48. Praus, P. A brief review of s-triazine graphitic carbon nitride. Carbon Lett. 2022, 32, 703–712. [Google Scholar] [CrossRef]
  49. Li, X.; Li, Q.; Cheng, J.; Liu, L.; Yan, Q.; Wu, Y.; Zhang, X.H.; Wang, Z.Y.; Qiu, Q.; Luo, Y. Conversion of dinitrogen to ammonia by FeN3-embedded graphene. J. Am. Chem. Soc. 2016, 138, 8706–8709. [Google Scholar] [CrossRef]
  50. Liu, C.; Li, Q.; Zhang, J.; Jin, Y.; MacFarlane, D.R.; Sun, C. Conversion of dinitrogen to ammonia on Ru atoms supported on boron sheets: A DFT study. J. Mater. Chem. A 2019, 7, 4771–4776. [Google Scholar] [CrossRef]
  51. Suryanto, B.H.R.; Du, H.; Wang, D.; 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]
  52. Gao, L.; Wang, F.; Yu, M.-a.; Wei, F.; Qi, J.; Lin, S.; Xie, D. A novel phosphotungstic acid-supported single metal atom catalyst with high activity and selectivity for the synthesis of NH3 from electrochemical N2 reduction: A DFT prediction. J. Mater. Chem. A 2019, 7, 19838–19845. [Google Scholar] [CrossRef]
  53. Zhao, J.; Chen, Z. Single Mo atom supported on defective boron nitride monolayer as an efficient electrocatalyst for nitrogen fixation: A computational study. J. Am. Chem. Soc. 2017, 139, 12480–12487. [Google Scholar] [CrossRef] [PubMed]
  54. Xu, Z.; Song, R.; Wang, M.; Zang, X.; Liu, G.; Qiao, G. Single atom-doped arsenene as electrocatalyst for reducing nitrogen to ammonia: A DFT study. Phys. Chem. Chem. Phys. 2020, 22, 26223–26230. [Google Scholar] [CrossRef]
  55. Xue, Z.; Zhang, X.; Qin, J.; Liu, R. Anchoring Mo on C9N4 monolayers as an efficient single atom catalyst for nitrogen fixation. J. Energy Chem. 2021, 57, 443–450. [Google Scholar] [CrossRef]
Figure 1. (a) Optimized structure and (b) DOS of s-tri-based g-C3N4 monolayer, (c) stable configuration of Mn@g-C3N4.
Figure 1. (a) Optimized structure and (b) DOS of s-tri-based g-C3N4 monolayer, (c) stable configuration of Mn@g-C3N4.
Nanomaterials 13 01433 g001
Figure 2. (a) Eads(*N2) and ΔG(*N2) of an N2 molecule on TM@g-C3N4, (b) DTM-N and LN-N of TM@g-C3N4.
Figure 2. (a) Eads(*N2) and ΔG(*N2) of an N2 molecule on TM@g-C3N4, (b) DTM-N and LN-N of TM@g-C3N4.
Nanomaterials 13 01433 g002
Figure 3. (a) The reaction free energy ΔG(*NNH), (b) absorption free energies of N2 molecules ΔG(*N2) and H atoms ΔG(*H) on the TM@g-C3N4 (TM = V, Cr, Mn, Fe, Co).
Figure 3. (a) The reaction free energy ΔG(*NNH), (b) absorption free energies of N2 molecules ΔG(*N2) and H atoms ΔG(*H) on the TM@g-C3N4 (TM = V, Cr, Mn, Fe, Co).
Nanomaterials 13 01433 g003
Figure 4. NRR free energy diagrams via (a) alternating (b) distal mechanisms for V@g-C3N4.
Figure 4. NRR free energy diagrams via (a) alternating (b) distal mechanisms for V@g-C3N4.
Nanomaterials 13 01433 g004
Figure 5. (a) Charge density difference with isosurface 0.004 e/Bohr3, the yellow and blue areas represent the accumulation and depletion of electrons, (b) spin density, the yellow and blue regions represent the spin-up and spin-down states, (c) DOS of V@g-C3N4, the Fermi level is set to 0 eV.
Figure 5. (a) Charge density difference with isosurface 0.004 e/Bohr3, the yellow and blue areas represent the accumulation and depletion of electrons, (b) spin density, the yellow and blue regions represent the spin-up and spin-down states, (c) DOS of V@g-C3N4, the Fermi level is set to 0 eV.
Nanomaterials 13 01433 g005
Figure 6. (a) Charge density difference with isosurface 0.004 e/Bohr3, the yellow and blue areas represent the accumulation and depletion of electrons, (b) spin density, the yellow and blue regions represent the spin-up and spin-down states, (c) DOS of N2/V@g-C3N4, the Fermi level is set to 0 eV.
Figure 6. (a) Charge density difference with isosurface 0.004 e/Bohr3, the yellow and blue areas represent the accumulation and depletion of electrons, (b) spin density, the yellow and blue regions represent the spin-up and spin-down states, (c) DOS of N2/V@g-C3N4, the Fermi level is set to 0 eV.
Nanomaterials 13 01433 g006
Figure 7. Charge variation diagrams of adsorbates via (a) distal and (b) alternating mechanism.
Figure 7. Charge variation diagrams of adsorbates via (a) distal and (b) alternating mechanism.
Nanomaterials 13 01433 g007
Figure 8. DV-N, LN-N and total magnetic moment (Mtot) of adsorbates on V@g-C3N4 via an (ac) distal mechanism and (df) alternating mechanism.
Figure 8. DV-N, LN-N and total magnetic moment (Mtot) of adsorbates on V@g-C3N4 via an (ac) distal mechanism and (df) alternating mechanism.
Nanomaterials 13 01433 g008
Table 1. Adsorption energies (Eads, eV), energy difference (ΔE, eV) between adsorption energy and cohesive energy, total magnetic moments (Mtot, μB), the charge (ΔQ, e) transferred from the TM atoms to g-C3N4 monolayer, and the electronic structures (ES) (SM: semi-metallic, M: metallic and SC: semiconductor).
Table 1. Adsorption energies (Eads, eV), energy difference (ΔE, eV) between adsorption energy and cohesive energy, total magnetic moments (Mtot, μB), the charge (ΔQ, e) transferred from the TM atoms to g-C3N4 monolayer, and the electronic structures (ES) (SM: semi-metallic, M: metallic and SC: semiconductor).
EadsΔEMtotΔQES
Sc@g-C3N4−9.55−5.361.002.35SM
Ti@g-C3N4−8.26−2.802.002.21M
V@g-C3N4−7.50−2.134.881.92M
Cr@g-C3N4−5.97−1.964.001.59SC
Mn@g-C3N4−6.20−2.341.001.48SC
Fe@g-C3N4−5.70−0.92−0.261.35M
Co@g-C3N4−7.25−1.891.001.02SC
Ni@g-C3N4−7.32−2.4200.94SC
Cu@g-C3N4−4.96−1.451.000.86SM
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chai, H.; Chen, W.; Feng, Z.; Li, Y.; Zhao, M.; Shi, J.; Tang, Y.; Dai, X. Single-Atom Anchored g-C3N4 Monolayer as Efficient Catalysts for Nitrogen Reduction Reaction. Nanomaterials 2023, 13, 1433. https://doi.org/10.3390/nano13081433

AMA Style

Chai H, Chen W, Feng Z, Li Y, Zhao M, Shi J, Tang Y, Dai X. Single-Atom Anchored g-C3N4 Monolayer as Efficient Catalysts for Nitrogen Reduction Reaction. Nanomaterials. 2023; 13(8):1433. https://doi.org/10.3390/nano13081433

Chicago/Turabian Style

Chai, Huadou, Weiguang Chen, Zhen Feng, Yi Li, Mingyu Zhao, Jinlei Shi, Yanan Tang, and Xianqi Dai. 2023. "Single-Atom Anchored g-C3N4 Monolayer as Efficient Catalysts for Nitrogen Reduction Reaction" Nanomaterials 13, no. 8: 1433. https://doi.org/10.3390/nano13081433

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