Next Article in Journal
Toward Enhanced Humidity Stability of Triboelectric Mechanical Sensors via Atomic Layer Deposition
Previous Article in Journal
Assessing the Toxicological Relevance of Nanomaterial Agglomerates and Aggregates Using Realistic Exposure In Vitro
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Linking Bi-Metal Distribution Patterns in Porous Carbon Nitride Fullerene to Its Catalytic Activity toward Gas Adsorption

Research Group Plasmant, NANO Lab Center of Excellence, Department of Chemistry, University of Antwerp, 2610 Antwerp, Belgium
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(7), 1794; https://doi.org/10.3390/nano11071794
Submission received: 27 May 2021 / Revised: 6 July 2021 / Accepted: 7 July 2021 / Published: 9 July 2021

Abstract

:
Immobilization of two single transition metal (TM) atoms on a substrate host opens numerous possibilities for catalyst design. If the substrate contains more than one vacancy site, the combination of TMs along with their distribution patterns becomes a design parameter potentially complementary to the substrate itself and the bi-metal composition. By means of DFT calculations, we modeled three dissimilar bi-metal atoms (Ti, Mn, and Cu) doped into the six porphyrin-like cavities of porous C24N24 fullerene, considering different bi-metal distribution patterns for each binary complex, viz. TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24 (with x, y, z = 0–6). We elucidate whether controlling the distribution of bi-metal atoms into the C24N24 cavities can alter their catalytic activity toward CO2, NO2, H2, and N2 gas capture. Interestingly, Ti2Mn4@C24N24 and Ti2Cu4@C24N24 complexes showed the highest activity and selectively toward gas capture. Our findings provide useful information for further design of novel few-atom carbon-nitride-based catalysts.

Graphical Abstract

1. Introduction

Porous carbon-based catalysts are widely used as sorbents and support materials in heterogeneous catalysis [1,2,3]. The archetypical example is C60 fullerene, which has a closed-cage structure and can be synthesized with a highly defective surface and abundant holes [4,5]. It has high thermal stability, unique mechanical properties, high electronegativity, and high electron affinity [6]. Therefore, both pure and metal-doped C60 show promising applications in energy conversion [7,8], fuel cells [9,10], and for biomedical applications [11].
Recently, the adsorption and conversion of gas molecules on doped C60 fullerene by single transition metal (TM) or heteroatoms, especially nitrogen atoms, has gained significant interest [12,13]. Upon N-doping, carbon π electrons are activated by conjugating with the lone-pair electrons from N. Thus, the C atoms neighboring N become active centers for catalytic reactions. For instance, Chen et al. [14] theoretically investigated the oxygen reduction reaction (ORR) mechanisms and catalytic abilities of pure and N-doped fullerenes of various sizes (C20, C40, C60, and C180). They found that the two pure and N-doped C20 and C180 structures are not active toward the adsorption of common intermediates produced during the ORR process. In contrast, C39N showed the largest decrease in reaction energy of the rate-determining step in the relative energy profile, suggesting its ORR activity is the best among all the different sizes of fullerenes. Experimentally, N-doped carbon materials are prepared using chemical vapor deposition (CVD) or reactive magnetron sputtering [15,16]. For example, Usachov et al. [17] synthesized a N-doped graphene nanosheet from s-triazine molecules. Recently, Zhai et al. [18] for the first time synthesized metal-free N-doped graphene films on glass through plasma-assisted hot filament CVD using N2 gas as dopant. They found that both the hot filament and plasma source are essential for growing N-doped graphene of high quality. By adjusting the N2 flow, the authors could easily modulate the N content, transmittance, and electrical properties of the graphene films.
The chemical inertness of pure fullerenes prevents their possible application for gas capture. Single metal doping (M-C60), like C58Pt and C59Pt [19], significantly modifies the electronic structure of C60, rendering it chemically active. However, one of the main challenges in the synthesis of M-C60 is the often-observed aggregation of metal atoms [8,20]. Recently, extensive efforts have been put into developing single-site catalysts such as pyrolyzed TM-modified porphyrin complexes (TM = Fe, Co, Mn, Cr, Ni) [21,22,23,24,25,26,27,28,29,30,31], which can be used in various applications like ORR and batteries [24,32,33,34,35,36]. The single TM atom doped in these porphyrin units was firmly fixed, preventing metal aggregation. Recently, the porphyrin-like porous C24N24 has been of great interest as truncated N-doped C60 nanocage. The C24N24 fullerene has six N4 cavities with eight s-triazine rings, which are connected via C-C bonds. Each porphyrin-like N4 cavity of C24N24 can host a single TM atom. Recently, TM-doped C24N24 and C24B24 fullerenes were studied for hydrogen storage [37,38,39], ORR [40], and gas conversion [41].
In this work, we investigate the catalytic activity of bi-metal atom doping in C24N24, using Ti, Mn, and Cu as metals. In particular, we here focus on the influence of the metal atom distribution in the catalytic activity of the modified structures. We carefully investigated the adsorption characteristics, electronic properties, and charge transfer properties of our novel catalysts. We found that varying the TM ratio indeed has an important effect on the properties and catalytic activities of the catalysts toward adsorption of gas species. Our calculations can provide the fundamental adsorption mechanism of such a novel material, supporting its possible exploitation to be applied as a green catalyst for gas detection.

2. Computational Details

All quantum chemical DFT computations are performed using the Gaussian16 package [42]. We first tested fourteen DFT functional/basis set combinations (including GGA, meta-GGA, and hybrid functionals) for geometry optimization (see Table S1 of Supplementary Materials). As can be seen in Table S1, the obtained Eb values calculated with the B3LYP-D3/6-311G* are in close agreement with the reported Eb for H2 (−4.52 eV) [43,44], formation energy (Ef) of pristine C24N24 (−7.40 eV), and Eb of Ti6@C24N24 (−8.14 eV), as well as the geometry values for C-C (1.55 Å) and C-N (1.34 Å) bond length [45].
The cohesive energy (per atom) of C24N24 is defined as:
Ecoh = (1/48)(EC24N24 − 24 EC − 24 EN)
where the EC24N24, EC, and EN are the calculated total energy of the pure nanocage, carbon, and nitrogen atom, respectively. The Eb of each doped TM in C24N24 for a homogeneous TM-doped C24N24 (TM6@C24N24) and with different metal ratios (x, y, z = 0–6) was calculated using Equations (2) and (3), respectively:
Eb(TM6@C24N24) = (1/6)(ETM6@C24N24 − 6 ETM − EC24N24)
Eb(TixMnyCuz@C24N24) = (1/6)(ETixMnyCuz-C24N24 − EC24N24 − xETi − yEMn − zECu), (x, y, z = 0–6)
where, in Equation (2), the ETM6@C24N24 and ETM refers to the total energy of one type TM-doped C24N24 (Ti, Mn, or Cu) and the TM atom, respectively. In Equation (3), the ETixMnyCuz@C24N24 is the total energy of the doped C24N24 complex and xETi, yEMn, and zECu are defined as the total energy of Ti, Mn, and Cu metal atoms times the number of doped Ti atoms (x), Mn atoms (y), and Cu atoms (z) into the six porphyrin cavities of C24N24 nanocage, respectively.
The adsorption energy (Eads) of each gas moiety over the modified cage is defined as:
Eads = Eads@TixMnyCuz-C24N24 − ETixMnyCuz-C24N24 − Eadsorbate
where Eads@TixMnyCuz-C24N24, ETixMnyCuz-C24N24, and Eadsorbate are the total energy of the adsorbate on the complex, pure complex, and the adsorbate molecule, respectively. The x, y, z values represent the number of each TM atom doped into the porphyrin-like C24N24 cavities and are in the range of 0 to 6.
For each system, the zero-point energy (EZPE) is calculated by summing vibrational frequencies over all (real) normal modes. Enthalpy and Gibbs energy changes are calculated at 298.15 K following the standard procedure outlined in the reference [46] To follow the nature of the adsorption process, the Wiberg bond indices (WBIs) were computed using NBO analysis. The WBIs are known as a better measure of bond strength relative to the overlap population, which is basis-set-dependent and often does not correlate well to bond strength. The WBIs are often similar in magnitude to the bond order expected from valence bond theory and have been used to propose trigger bonds in various energetic materials.
The WBI is a measure of the density between two atoms A and B. It determines by the sum of the off-diagonal square of the density matrix P (p#q) [47]:
WBI AB = P A q B P 2 pq

3. Results

3.1. Geometry of Pristine C24N24

The porous C24N24 is formed by first removing the 12 C atoms in C60 that connect two pentagons, thus creating six di-vacancies, as shown with a red circle in Figure 1, and subsequently substituting four undercoordinated C-atoms with N-atoms, thus creating six porphyrin-like N4 cavities and eight connected s-triazine rings (see Figure 1). The calculated C-C and C-N bond lengths are 1.55 Å and 1.34 Å, respectively. The calculated cohesive energy per atom in C24N24 is Ecoh = −7.81 eV, which is close to that reported by Ghosh et al. (−7.40 eV) [45] but significantly higher than the value of Tang et al. (−5.78 eV) [38].
The HOMO-LUMO gap of C24N24 is calculated to be to be Eg = 2.82 eV, which is similar to the value reported by Ma et al. [39] and Song et al. [48] and is higher than the value reported in other investigations [40,45].

3.2. Metal Distribution Patterns

3.2.1. Geometry and Electronic Properties of Bi-Metal Complexes

The agglomeration of catalytically active TMs into clusters is a major challenge. This can be prevented by the strong interaction between TMs and support. C24N24 fullerene possesses natural N4 rings that can host TM atoms (see Figure 1). We considered a combination of two metal atoms for the selected Ti, Mn, and Cu TMs with 3d24s2, 3d54s2, and 3d104s1 valence electrons, respectively, i.e., TiCu, TiMn, and MnCu. Then, various distribution patterns of these dissimilar bi-metal atoms into the six porphyrin-like C24N24 cavities were studied. Due to the presence of six cavities, the TM ratios will vary between zero to six. Therefore, we have TixCuz, TixMny, and MnyCuz doped C24N24 fullerene (TixCuz@C24N24, TixMny@C24N24, MnyCuz@C24N24) with x, y, z = 0–6. Figure 2a–j shows a schematic presentation of possible bi-metal distribution patterns in the N4 cavities of C24N24 cage. There are two main doping sites available in the C24N24 cavities: equatorial and axial. As can be seen in Figure 2, equatorial and axial positions are placed in x, y (red and black dashed lines), and z (green dashed line) axis directions, respectively. Two metal atoms in a complex can be distributed into the C24N24 cavities through seven different distribution patterns, as listed in Table 1.

3.2.2. Binding Energy (Eb)

To assess the stability of the modified complexes, the binding energy of each configuration was calculated (see Tables S2–S4). The computed Eb for each complex changes in the range of −3.55 eV to −8.11 eV. The lowest and highest Eb values correspond to the homogeneous doping with Cu (Cu6@C24N24) and Ti (Ti6@C24N24), respectively. The calculated Eb for Ti6@C24N24 is in good agreement with previous studies (−8.61 eV and −8.14 eV) [39,45]. The binding energy of TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24 (x, y, z = 0–6) configurations is plotted and compared in Figure 3.
The high Eb values indicate a strong chemisorption of two dissimilar TMs at six N4 cavities of C24N24, which inhibits the diffusion of TMs into the C24N24 fullerene, increasing the stability of nanocages by metal doping. Due to the higher binding energy of Ti6@C24N24 and Mn6@C24N24 than the cohesive energy of bulk Ti (4.85 eV/atom) and Mn (2.92 eV/atom) [49], TM aggregation is not expected, and these materials are likely to be stable enough to be used in catalytic processes. In contrast, Cu6@C24N24 shows an Eb of −3.55 eV, slightly higher than the cohesive energy of bulk Cu (3.49 eV/atom) [50], such that homogeneously Cu-doped C24N24 is much less stable against TM agglomeration.
One can see from Figure 3 and Tables S2–S4 that by increasing the ratio of Cu-to-Ti, Mn-to-Ti, and Cu-to-Mn atoms in TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24, the Eb decreases (i.e., becomes less negative).

3.2.3. NBO Charge Analysis

Figure 4 shows the NBO charge accumulation on individual TMs in different configurations (a–j) for each complex. The precise values are reported in Tables S2–S4. Ti and Cu atoms in TixCuz@C24N24 and MnyCuz@C24N24 complexes have the highest charge transfer values from the metal atom to the nanocage, in the range of 1.34 to 1.44 |e| and 0.67 to 0.69 |e|, respectively. The high charge on the metal atoms corresponds to the high electron donation induced by the four surrounding N atoms in each porphyrin-like cavity where the d orbitals of Ti, Mn, and Cu atoms overlap with the neighboring nitrogen sp2 orbitals of the cavity to form a sp2d hybridization. The charge transfer leads to the elongation of the average C-C bond (≈1.55 Å) compared to that of pristine C60 (1.45 Å), indicating the activation of C24N24 fullerene upon bi-metal doping.

3.2.4. Thermodynamic Properties and Energy Gap (Eg)

Figure 5 shows the changes in Gibbs free energy for each doped configuration, showing that bi-metallic doping is exothermic and exergonic at room temperature. The values are reported in Tables S2–S4.
Doping C24N24 with various Ti, Mn, or Cu distributions significantly narrows the HOMO-LUMO gap of the C24N24, leading to a noticeable energy gap (Eg) reduction. The highest Eg reduction occurs in Ti2Mn4@C24N24, where the calculated Eg is reduced from 2.82 eV in pure C24N24 to 0.13 eV, making the complex act as a semi-conductor (see Figure 5b, complex g, purple line). The calculated HOMO−LUMO gap for doped configurations does not follow any particular order but in general is less than <1.58 eV, and thus all configurations can be classified as semiconductors. Since the electrical conductivity is exponentially related to Eg, we expect that Ti2Mn4@C24N24 may show good electrical conductivity.
Besides the effects of doping on the geometric, thermodynamic, and electronic properties of the complexes as described above, we find that the location of the second introduced TM atom in axial or equatorial positions has little effect. Therefore, we expect that the complexes c and d, e and f, and g and h exhibit similar catalytic activities.

4. Catalytic Behavior of Bi-Metal Complexes toward Adsorption of Gas Species

4.1. Geometric Properties

To explore the effect of bi-metal-doping on the catalytic behavior of each complex, we investigated the individual adsorption of four gas molecules (CO2, NO2, H2, and N2) on each complex and their subgroup configurations. As shown in previous investigations, the existence of two dissimilar TM atoms in a catalyst induces various active sites [51,52,53,54]. Thus, it is important to first determine the available active sites in each substrate (see Figure S1). Considering each complex as a sphere, TM atoms are placed in x, y, and z directions. In homogeneous doping with one TM atom, one active site will be available (Figure S1a–j). By adding the second TM atom to the complex, it can be doped into one of the porphyrin vacancies along the x, y, or z direction, forming the complexes that are shown in Figure S1b. With two different TM atoms present in the complex, one would expect to have two different active sites available. However, for the adsorbent, there are not only two sites. We have labeled the TM atoms in Figure S1 (denoted as i, j, and k) to show different possible positions for the adsorbents to adsorb on the TM sites of each catalyst. It is clear that i, k/j sites are located on the axial (perpendicular to the plane of the ring)/equatorial (in the plane of the ring) axis of the catalyst, respectively. Depending on the TM ratio, the adsorbent can adsorb on either i, j, or k positions. One can see that for each configuration, two active sites are available except in configurations b and i, in which, depending on the adsorption position of the adsorbent in the equatorial (j) or axial (i and k) axis, three possible active sites are available. In addition, each adsorbent adopts mainly three adsorption modes on each active site of the substrates: parallel (side-on) or vertical (end-on) to the surface, and bridge positions.
Our results reveal that CO2, H2, and N2 adsorb strongly on TixCuz@C24N24 and TixMny@C24N24 and weakly on MnyCuz@C24N24, while NO2 adsorbs quite strongly with |Eads| > 6 eV on all three complexes (see Figure 6 and Table S5). The adsorption behavior of each set of configurations is discussed below.

4.2. Adsorption of CO2 and H2

Ti2Cu4@C24N24 and Ti2Mn4@C24N24 tend to chemisorb CO2 and H2 species with high adsorption energies (see Figure 7). The covalent nature of the adsorbed species is confirmed by the calculated WBIs. As a result, the structure of CO2 is drastically distorted upon its adsorption on these complexes. It is bent over the Ti atom binding via its C and O and forms a triangular ring above the nanocage. The O-C-O angle is bent to 131.12° in its adsorbed form and the C = O bond length is elongated to 1.35 Å, which we attribute to the significant charge transfer of 0.49 |e| from the complex to the 2π* orbitals of the CO2 molecule. The obtained adsorption energies for CO2 are lower than those reported on B80 fullerene (−3.49 eV) [55] and higher than those of N-S dual doped graphene (−0.25 eV) [56] and Ti-doped C2N (Eads = −0.95 eV) [57]. The WBIs of both Ti-C and Ti-O bonds in Ti2Cu4@C24N24 and Ti2Mn4@C24N24 are 0.84, confirming the covalent bond between TM and CO2 atoms and therefore its chemisorption over these substrates. H2 adsorbs dissociatively in a barrier-less reaction, forming two covalent Ti-H bonds above the Ti atom with H-H bond length of >2.0 Å (see Figure 7). The NBO charge analysis (see Table S5) along with the calculated WBIs for both Ti-H bond lengths (0.89) confirms the covalent binding between Ti and H atoms and consequently the H2 chemisorption. Interestingly, we find that the adsorption energy of one hydrogen molecule on these complexes is higher than the adsorption of six H2 molecules on Ti6@C24N24 (Eads = −0.48 eV) [39] [39], Fe-B38 (Eads = −0.42 eV), Co-B38 (Eads = −0.72 eV), Ni-B38 (Eads = −0.89 eV) [58], and 2D carbon allotrope Ψ graphene (Eads = −0.34 eV) [59].
Due to the lower activity of the Cu6@C24N24 complex, CO2 and H2 physisorb on this structure in their gas-phase form. The calculated Eads of H2 on Cu6@C24N24 is lower than that reported on Ti2C- and Ti2CN-Mxenes (Eads in the range of −0.99 to −1.4 eV) [60]. A negligible charge transfer from CO2 (H2) to Cu6@C24N24, the large Cu-C (Cu-H) bond length, and the low WBIs value of 0.12 (0.2) confirm physisorption of these CO2 (H2) species on Cu6@C24N24.

4.3. Adsorption of N2 and NO2

Two different orientations were considered for N2 adsorption on each structure: side-on or end-on. The ideal catalyst would provide strong binding sites for the N2 molecule and thus weaken the N−N bond. Our results indicate that the ideal orientation for N2 adsorption on Ti2Cu4@C24N24 and Ti2Mn4@C24N24 and Mn3Cu3@C24N24 is the end-on (see Figure 8). Upon N2 adsorption, the N-N bond length increases from 1.09 Å in the gas phase to 1.11 and 1.12 Å in its adsorbed form on Ti2Cu4@C24N24 and Ti2Mn4@C24N24 complexes, respectively. These values are in between the double and triple bond lengths, indicating the activation of N2 upon adsorption on Ti sites. The empty d orbitals of the Ti atom can accept the lone-pair electrons of N2. In turn, the Ti’s capacity to donate electrons to the antibonding π* orbital of N2 is also significant for N2 binding to Ti. Therefore, this electron acceptance/donation process between the nanocage and N2 plays an important role in N2 activation (see Table S5). Ti2Mn4@C24N24 has a greater tendency for N2 activation with higher adsorption energy than that reported on Co-doped graphitic carbon nitride (−1.63 eV) [61] and Fe doped phosphorene (−0.81 eV) [62].
Although the Mn3Cu3@C24N24 complex adsorbs N2 molecule with lower Eads, due to its lower catalytic activity toward gas adsorption, this value is still higher than that on Mn-Fe bi-metal atoms anchored pyridinic nitrogen-doped graphene (Eads = −0.53 eV) [63]. The calculated WBIs (≈0.0001) and charge transfer confirm the N2 physisorption on Mn3Cu3@C24N24.
Regardless of all possible bi-metal distributions into the porphyrin C24N24 cavities, the three TixCuz@C24N24, TizMny@C24N24, and MnyCuz@C24N24 (x, z, y = 0–6) complexes with all the metal ratios exhibit an outstanding activity toward NO2 adsorption and activation. However, the more stable configurations are shown in Figure 8. One can see that the NO2 molecule binds via two O atoms with the Ti atom in Ti6@C24N24 and Ti2Mn4@C24N24 and with the Mn atom in the Mn5Cu@C24N24 complex, respectively. Owing to the great NBO charge transfer from the nanocage to the NO2 molecule (0.46 |e|) reported in Table S5, the N-O bond length increases compared to that of the gas phase (1.19 Å).

4.4. Electronic and Thermodynamic Properties

To see if the adsorption of CO2, NO2, H2, and N2 gas species on the selected twelve more energetically stable complexes affects their electronic properties, we also investigated the LUMO-HOMO energy gap of each structure and compared our results with those of pristine bi-metal doped nanocages reported in Figure 5. The obtained Eg values for adsorption structures are listed in Table S5, showing that Eg indeed increases upon gas adsorption. This confirms the tunable electronic properties of C24N24 nanocage induced by hosting six dissimilar bi-metals doping into its N4 cavities with various distribution patterns. All adsorption reactions are exothermic and exergonic, except for H2 adsorbed on the Cu6@C24N24 complex, which is slightly endergonic (see Table S5).

4.5. Lifetime of the Adsorbed Gas Species on Bi-Metal Complexes

The retention time of a molecule on a surface can be calculated with the Frenkel equation [64]:
τ = τ 0 e Q R T
where τ 0 is 10−12 to 10−13 s and Q is the adsorption energy. We calculated the lifetime of each adsorbent over each configuration from which the energetically more favorable structures were chosen and plotted versus temperature. Figure 9 shows the computed lifetime vs. temperature for adsorbed CO2, NO2, H2, and N2 on the twelve nanocages discussed above. As can be seen in Figure 9, the lifetime of gas species on TixCuz@C24N24 and TixMny@C24N24 at 400 K is higher than that on MnyCuz@C24N24 (x, y, z = 0–6), indicating that TixCuz@C24N24 and TixMny@C24N24 are likely to be more efficient for gas capture. Obviously, increasing the temperature reduces the lifetime leading. However, TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24 catalysts capture NO2 so actively that it will not desorb from the catalyst even at high temperatures. Therefore, we can conclude that, except for NO2, which effectively adsorbs and is retained on the catalysts, the gas species adsorb and remain on the Ti active site of TixCuz@C24N24 and TixMny@C24N24 fullerene at room temperature, whereas they are not retained on MnyCuz@C24N24 sufficiently for gas capture.

5. Conclusions

In this work, the effect of dissimilar bi-metal doping into the six porphyrin-like cavities of a C24N24 nanocage on the catalytic activity and adsorption characteristic of a number of greenhouse gases are investigated by means of DFT calculations. The binding energy and bulk cohesive energy calculations reveal that the selected TM atoms are stably trapped in the C24N24 cavities, especially Ti/Cu atoms in TixCuz@C24N24 and Ti/Mn atoms in TixMny@C24N24 (x, y, z = 0–6) complexes, suggesting the durability of the catalysts. Studying the adsorption behavior of these catalysts toward H2, CO2, NO2, and N2 sensing show that the Ti2Mn4@C24N24 is more active for adsorption of all species. Furthermore, the lifetime of each gas species on TixCuz@C24N24 and TixMny@C24N24 at 400 K is higher than that on MnyCuz@C24N24 (x, y, z = 0–6), indicating that TixMny@C24N24 and TixCuz@C24N24 are likely to be more efficient for gas capture. Overall, this work systematically provides the unique fundamental understanding of catalytic properties of two dissimilar bi-metal atom catalysts that could open a way for the future design and development of novel few-atom catalysts.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano11071794/s1, Figure S1: A schematic representative of available active sites (shown with red stars) on various BM@C24N24 configurations.Color code: white, Ti; orange, Cu., Table S1: The calculated cohesive energy (Ecoh) per atom of C24N24, binding energy (Eb) of hydrogen molecule and Ti6-doped C24N24 along the energy gap of pristine C24N24 using different functional and basis sets, Table S2: The calculated binding energy (Eb), change of enthalpy (ΔH298), change of Gibbs free energy (ΔG298), energy gap (Eg), and the average NBO charge on TixCuz@C24N24. All the values are in eV, Table S3: The calculated binding energy (Eb), change of enthalpy (ΔH298), change of Gibbs free energy (ΔG298), energy gap (Eg), and the average NBO charge on TixMny@C24N24. All the values are in eV, Table S4: The calculated binding energy (Eb), change of enthalpy (ΔH298), change of Gibbs free energy (ΔG298), energy gap (Eg), and the average NBO charge on MnyCuz@C24N24. All the values are in eV, Table S5: The calculated total adsorption energy (Eads), energy gap (Eg), changes of enthalpy(ΔH298), changes of free energy(ΔG298), and NBO charge analysis for the energetically more stable adsorption configurations.

Author Contributions

Conceptualization, P.N.; methodology, P.N.; software, P.N.; validation, P.N.; formal analysis, P.N.; investigation, P.N.; resources, P.N.; data curation, P.N.; writing—original draft preparation, P.N.; writing—review and editing, P.N. and E.C.N.; visualization, P.N.; supervision, E.C.N.; project administration, P.N.; funding acquisition, E.C.N. and P.N. Both authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fund of Scientific Research Flanders (FWO), Belgium, Grant number 1261721N. The work was carried out in part using the Turing HPC infrastructure of the CalcUA core facility of the Universiteit Antwerpen, a division of the Flemish Supercomputer Centre VSC, funded by the Hercules Foundation, the Flemish Government (department EWI), and the University of Antwerp.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Trogadas, P.; Fuller, T.F.; Strasser, P. Carbon as catalyst and support for electrochemical energy conversion. Carbon 2014, 75, 5–42. [Google Scholar] [CrossRef]
  2. Srinivasu, K.; Modak, B.; Ghosh, S.K. Porous graphitic carbon nitride: A possible metal-free photocatalyst for water splitting. J. Phys. Chem. C 2014, 118, 26479–26484. [Google Scholar] [CrossRef]
  3. Albero, J.; Vidal, A.; Migani, A.; Concepción, P.; Blancafort, L.; García, H. Phosphorus-doped graphene as a metal-free material for thermochemical water reforming at unusually mild conditions. ACS Sustain. Chem. Eng. 2018, 7, 838–846. [Google Scholar] [CrossRef]
  4. Pan, Y.; Liu, X.; Zhang, W.; Liu, Z.; Zeng, G.; Shao, B.; Liang, Q.; He, Q.; Yuan, X.; Huang, D.; et al. Advances in photocatalysis based on fullerene C60 and its derivatives: Properties, mechanism, synthesis, and applications. Appl. Catal. B Environ. 2020, 265, 118579. [Google Scholar] [CrossRef]
  5. Han, F.; Wang, R.; Feng, Y.; Wang, S.; Liu, L.; Li, X.; Han, Y.; Chen, H. On demand synthesis of hollow fullerene nanostructures. Nat. Commun. 2019, 10, 1–9. [Google Scholar] [CrossRef] [Green Version]
  6. Zettergren, H.; Alcami, M.; Martín, F. First- and second-electron affinities ofC60andC70isomers. Phys. Rev. A 2007, 76, 043205. [Google Scholar] [CrossRef]
  7. Wu, Q.; Yang, L.; Wang, X.; Hu, Z. Carbon-based nanocages: A new platform for advanced energy storage and conversion. Adv. Mater. 2020, 32, e1904177. [Google Scholar] [CrossRef]
  8. El Mahdy, A.M. Density functional investigation of CO and NO adsorption on TM-decorated C60 fullerene. Appl. Surf. Sci. 2016, 383, 353–366. [Google Scholar] [CrossRef]
  9. Coro, J.; Suárez, M.; Silva, L.S.; Eguiluz, K.; Banda, G. Fullerene applications in fuel cells: A review. Int. J. Hydrogen Energy 2016, 41, 17944–17959. [Google Scholar] [CrossRef]
  10. Wang, Y.; Jiao, M.; Song, W.; Wu, Z. Doped fullerene as a metal-free electrocatalyst for oxygen reduction reaction: A first-principles study. Carbon 2017, 114, 393–401. [Google Scholar] [CrossRef]
  11. Cai, Q.; Hu, Z.; Zhang, Q.; Li, B.; Shen, Z. Fullerene (C60)/CdS nanocomposite with enhanced photocatalytic activity and stability. Appl. Surf. Sci. 2017, 403, 151–158. [Google Scholar] [CrossRef]
  12. Sood, P.; Kim, K.C.; Jang, S.S. Electrochemical properties of boron-doped fullerene derivatives for lithium-ion battery ap-plications. ChemPhysChem 2018, 19, 753–758. [Google Scholar] [CrossRef] [PubMed]
  13. Paraknowitsch, J.P.; Thomas, A. Doping carbons beyond nitrogen: An overview of advanced heteroatom doped carbons with boron, sulphur and phosphorus for energy applications. Energy Environ. Sci. 2013, 6, 2839–2855. [Google Scholar] [CrossRef] [Green Version]
  14. Chen, X.; Chang, J.; Ke, Q. Probing the activity of pure and N-doped fullerenes towards oxygen reduction reaction by density functional theory. Carbon 2018, 126, 53–57. [Google Scholar] [CrossRef]
  15. Hultman, L.; Stafström, S.; Czigány, Z.; Neidhardt, J.; Hellgren, N.; Brunell, I.F.; Suenaga, K.; Colliex, C. Cross-linked nano-onions of carbon nitride in the solid phase: Existence of a novel C 48 N 12 Aza-fullerene. Phys. Rev. Lett. 2001, 87, 225503. [Google Scholar] [CrossRef] [PubMed]
  16. Otero, G.; Biddau, G.; Sanchez, C.S.; Caillard, R.; López, M.F.; Rogero, C.; Palomares, F.J.; Cabello, N.; Basanta, M.A.; Ortega, J.; et al. Fullerenes from aromatic precursors by surface-catalysed cyclodehydrogenation. Nat. Cell Biol. 2008, 454, 865–868. [Google Scholar] [CrossRef]
  17. Usachov, D.; Vilkov, O.; Grüneis, A.; Haberer, D.; Fedorov, A.; Adamchuk, V.; Preobrajenski, A.B.; Dudin, P.; Barinov, A.; Oehzelt, M.; et al. Nitrogen-doped graphene: Efficient growth, structure, and electronic properties. Nano Lett. 2011, 11, 5401–5407. [Google Scholar] [CrossRef]
  18. Zhai, Z.; Shen, H.; Chen, J.; Li, X.; Jiang, Y. Direct growth of nitrogen-doped graphene films on glass by plasma-assisted hot filament CVD for enhanced electricity generation. J. Mater. Chem. A 2019, 7, 12038–12049. [Google Scholar] [CrossRef]
  19. Hayashi, A.; Xie, Y.; Poblet, J.M.; Campanera, J.M.; Lebrilla, C.B.; Balch, A.L. Mass spectrometric and computational studies of heterofullerenes ([C58Pt]-, [C59Pt]+) Obtained by laser ablation of electrochemically deposited films. J. Phys. Chem. A 2004, 108, 2192–2198. [Google Scholar] [CrossRef]
  20. Gao, B.; Zhao, J.-X.; Cai, Q.-H.; Wang, X.-G.; Wang, X.-Z. Doping of calcium in C60Fullerene for enhancing CO2Capture and N2O transformation: A theoretical study. J. Phys. Chem. A 2011, 115, 9969–9976. [Google Scholar] [CrossRef]
  21. He, T.; Santiago, A.R.P.; Du, A. Atomically embedded asymmetrical dual-metal dimers on N-doped graphene for ultra-efficient nitrogen reduction reaction. J. Catal. 2020, 388, 77–83. [Google Scholar] [CrossRef]
  22. Guo, C.; Zhang, T.; Liang, X.; Deng, X.; Guo, W.; Wang, Z.; Lu, X.; Wu, C.-M.L. Single transition metal atoms on nitrogen-doped carbon for CO2 electrocatalytic reduction: CO production or further CO reduction? Appl. Surf. Sci. 2020, 533, 147466. [Google Scholar] [CrossRef]
  23. Gao, X.; Zhou, Y.; Liu, S.; Cheng, Z.; Tan, Y.; Shen, Z. Single cobalt atom anchored on N-doped graphyne for boosting the overall water splitting. Appl. Surf. Sci. 2020, 502, 144155. [Google Scholar] [CrossRef]
  24. Asset, T.; Atanassov, P. Iron-nitrogen-carbon catalysts for proton exchange membrane fuel cells. Joule 2020, 4, 33–44. [Google Scholar] [CrossRef]
  25. Nematollahi, P.; Neyts, E. Direct methane conversion to methanol on M and MN4 embedded graphene (M=Ni and Si): A comparative DFT study. Appl. Surf. Sci. 2019, 496, 143618. [Google Scholar] [CrossRef]
  26. Liu, S.; Cheng, L.; Li, K.; Yin, C.; Tang, H.; Wang, Y.; Wu, Z. RuN4 doped graphene oxide, a highly efficient bifunctional catalyst for oxygen reduction and CO2 reduction from computational study. ACS Sustain. Chem. Eng. 2019, 7, 8136–8144. [Google Scholar] [CrossRef]
  27. Holst-Olesen, K.; Silvioli, L.; Rossmeisl, J.; Arenz, M. Enhanced oxygen reduction reaction on Fe/N/C catalyst in acetate buffer electrolyte. ACS Catal. 2019, 9, 3082–3089. [Google Scholar] [CrossRef]
  28. Jia, Y.; Xiong, X.; Wang, D.; Duan, X.; Sun, K.; Li, Y.; Zheng, L.; Lin, W.; Dong, M.; Zhang, G.; et al. Atomically dispersed Fe-N4 modified with precisely located S for highly efficient oxygen reduction. Nano-Micro Lett. 2020, 12, 1–13. [Google Scholar] [CrossRef]
  29. Chung, H.T.; Cullen, D.A.; Higgins, D.; Sneed, B.T.; Holby, E.F.; More, K.L.; Zelenay, P. Direct atomic-level insight into the active sites of a high-performance PGM-free ORR catalyst. Science 2017, 357, 479–484. [Google Scholar] [CrossRef] [Green Version]
  30. Wan, X.; Liu, X.; Li, Y.; Yu, R.; Zheng, L.; Yan, W.; Wang, H.; Xu, M.; Shui, J. Fe–N–C electrocatalyst with dense active sites and efficient mass transport for high-performance proton exchange membrane fuel cells. Nat. Catal. 2019, 2, 259–268. [Google Scholar] [CrossRef]
  31. Zitolo, A.; Ranjbar-Sahraie, N.; Mineva, T.; Emiliano, F.; Jia, Q.; Stamatin, S.; Harrington, G.F.; Lyth, S.M.; Krtil, P.; Mukerjee, S.; et al. Identification of catalytic sites in cobalt-nitrogen-carbon materials for the oxygen reduction reaction. Nat. Commun. 2017, 8, 1–11. [Google Scholar] [CrossRef] [PubMed]
  32. Kattel, S.; Wang, G. A density functional theory study of oxygen reduction reaction on Me–N4 (Me = Fe, Co, or Ni) clusters between graphitic pores. J. Mater. Chem. A 2013, 1. [Google Scholar] [CrossRef]
  33. Kattel, S.; Atanassov, P.; Kiefer, B. Density functional theory study of Ni–Nx/C electrocatalyst for oxygen reduction in al-kaline and acidic media. J. Phys. Chem. C. 2012, 116, 17378–17383. [Google Scholar] [CrossRef]
  34. Du, Z.; Chen, X.; Hu, W.; Chuang, C.; Xie, S.; Hu, A.; Yan, W.; Kong, X.; Wu, X.; Ji, H. Cobalt in nitrogen-doped graphene as single-atom catalyst for high-sulfur content lithium-sulfur batteries. J. Am. Chem. Soc. 2019, 141, 3977–3985. [Google Scholar] [CrossRef]
  35. Bi, W.; Li, X.; You, R.; Chen, M.; Yuan, R.; Huang, W.; Wu, X.; Chu, W.; Wu, C.; Xie, Y. Surface immobilization of transition metal ions on nitrogen-doped graphene realizing high-efficient and selective CO2 reduction. Adv. Mater. 2018, 30, e1706617. [Google Scholar] [CrossRef]
  36. Zhang, C.; Sha, J.; Fei, H.; Liu, M.; Yazdi, S.; Zhang, J.; Zhong, Q.; Zou, X.; Zhao, N.; Yu, H.; et al. Single-atomic ruthenium catalytic sites on nitrogen-doped graphene for oxygen reduction reaction in acidic medium. ACS Nano 2017, 11, 6930–6941. [Google Scholar] [CrossRef] [Green Version]
  37. Guo, J.; Liu, Z.; Liu, S.; Zhao, X.; Huang, K. High-capacity hydrogen storage medium: Ti doped fullerene. Appl. Phys. Lett. 2011, 98, 023107. [Google Scholar] [CrossRef]
  38. Tang, C.; Chen, S.; Zhu, W.; Kang, J.; He, X.; Zhang, Z. Transition metal Ti coated porous fullerene C24B24: Potential material for hydrogen storage. Int. J. Hydrogen Energy 2015, 40, 16271–16277. [Google Scholar] [CrossRef]
  39. Ma, L.-J.; Hao, W.; Han, T.; Rong, Y.; Jia, J.; Wu, H.-S. Sc/Ti decorated novel C24N24 cage: Promising hydrogen storage materials. Int. J. Hydrogen Energy 2021, 46, 7390–7401. [Google Scholar] [CrossRef]
  40. Modak, B.; Srinivasu, K.; Ghosh, S.K. Exploring metal decorated porphyrin-like porous fullerene as catalyst for oxygen re-duction reaction: A DFT study. Int. J. Hydrogen Energy 2017, 42, 2278–2287. [Google Scholar] [CrossRef]
  41. Shakerzadeh, E.; Hamadi, H.; Esrafili, M.D. Computational mechanistic insights into CO oxidation reaction over Fe decorated C24N24 fullerene. Inorg. Chem. Commun. 2019, 106, 190–196. [Google Scholar] [CrossRef]
  42. Ogliaro, F.; Bearpark, M.; Heyd, J.; Brothers, E.; Kudin, K.; Staroverov, V.; Keith, T.; Kobayashi, R.; Normand, J.; Raghavachari, K. Gaussian 16; Revision A. 03; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  43. Lide, D.R. CRC Handbook of Chemistry and Physics; CRC Press: Boca Raton, FL, USA, 2004. [Google Scholar]
  44. De Lara-Castells, M.P.; Stoll, H.; Civalleri, B.; Causà, M.; Voloshina, E.; Mitrushchenkov, A.O.; Pi, M. Communication: A combined periodic density functional and incremental wave-function-based approach for the dispersion-accounting time-resolved dynamics of 4He nanodroplets on surfaces: 4He/graphene. J. Chem. Phys. 2014, 141, 151102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Srinivasu, K.; Ghosh, S.K. Transition metal decorated porphyrin-like porous fullerene: Promising materials for molecular hydrogen adsorption. J. Phys. Chem. C 2012, 116, 25184–25189. [Google Scholar] [CrossRef]
  46. Ochterski, J.W. Thermochemistry in Gaussian; Gaussian Inc.: Wallingford, CT, USA, 2000. [Google Scholar]
  47. Wiberg, K. Application of the pople-santry-segal CNDO method to the cyclopropylcarbinyl and cyclobutyl cation and to bicyclobutane. Tetrahedron 1968, 24, 1083–1096. [Google Scholar] [CrossRef]
  48. Song, Y.-D.; Wang, L.; Wu, L.-M. How the alkali metal atoms affect electronic structure and the nonlinear optical properties of C24N24 nanocage. Optik 2017, 135, 139–152. [Google Scholar] [CrossRef]
  49. Holcomb, D.; Kittel, C. Introduction to solid state physics. Am. J. Phys. 1967, 35. [Google Scholar] [CrossRef] [Green Version]
  50. Philipsen, P.; Baerends, E. Cohesive energy of 3D transition metals: Density functional theory atomic and bulk calculations. Phys. Rev. B 1996, 54, 5326. [Google Scholar] [CrossRef] [Green Version]
  51. Wei, Z.; He, J.; Yang, Y.; Xia, Z.; Feng, Y.; Ma, J. Fe, V-co-doped C2N for electrocatalytic N2-to-NH3 conversion. J. Energy Chem. 2021, 53, 303–308. [Google Scholar] [CrossRef]
  52. Wang, Z.; Lee, H.; Chen, J.; Wu, M.; Leung, D.Y.; Grimes, C.A.; Lu, Z.; Xu, Z.; Feng, S.-P. Synergistic effects of Pd–Ag bimetals and g-C3N4 photocatalysts for selective and efficient conversion of gaseous CO2. J. Power Sources 2020, 466, 228306. [Google Scholar] [CrossRef]
  53. Bian, Z.; Das, S.; Wai, M.H.; Hongmanorom, P.; Kawi, S. A review on bimetallic nickel-based catalysts for CO2 reforming of methane. ChemPhysChem 2017, 18, 3117–3134. [Google Scholar] [CrossRef] [Green Version]
  54. Wei, X.; Yang, X.-F.; Wang, A.-Q.; Li, L.; Liu, X.-Y.; Zhang, T.; Mou, C.-Y.; Li, J. Bimetallic Au–Pd alloy catalysts for N2O decomposition: Effects of surface structures on catalytic activity. J. Phys. Chem. C. 2012, 116, 6222–6232. [Google Scholar] [CrossRef]
  55. Sun, Q.; Wang, M.; Li, Z.; Du, A.; Searles, D.J. Carbon dioxide capture and gas separation on B80Fullerene. J. Phys. Chem. C 2014, 118, 2170–2177. [Google Scholar] [CrossRef]
  56. Li, J.; Hou, M.; Chen, Y.; Cen, W.; Chu, Y.; Yin, S. Enhanced CO2 capture on graphene via N, S dual-doping. Appl. Surf. Sci. 2017, 399, 420–425. [Google Scholar] [CrossRef]
  57. Deshpande, S.S.; Deshpande, M.D.; Hussain, T.; Ahuja, R. Investigating CO2 storage properties of C2N monolayer functionalized with small metal clusters. J. CO2 Util. 2020, 35, 1–13. [Google Scholar] [CrossRef]
  58. Liu, P.; Zhang, H.; Cheng, X.; Tang, Y. Transition metal atom Fe, Co, Ni decorated B38 fullerene: Potential material for hydrogen storage. Int. J. Hydrogen Energy 2017, 42, 15256–15261. [Google Scholar] [CrossRef]
  59. Chakraborty, B.; Ray, P.; Garg, N.; Banerjee, S. High capacity reversible hydrogen storage in titanium doped 2D carbon allotrope Ψ-graphene: Density Functional Theory investigations. Int. J. Hydrogen Energy 2021, 46, 4154–4167. [Google Scholar] [CrossRef]
  60. Ding, B.; Ong, W.-J.; Jiang, J.; Chen, X.; Li, N. Uncovering the electrochemical mechanisms for hydrogen evolution reaction of heteroatom doped M2C MXene (M = Ti, Mo). Appl. Surf. Sci. 2020, 500, 143987. [Google Scholar] [CrossRef]
  61. Wang, K.; Gu, G.; Hu, S.; Zhang, J.; Sun, X.; Wang, F.; Li, P.; Zhao, Y.; Fan, Z.; Zou, X. Molten salt assistant synthesis of three-dimensional cobalt doped graphitic carbon nitride for photocatalytic N2 fixation: Experiment and DFT simulation analysis. Chem. Eng. J. 2019, 368, 896–904. [Google Scholar] [CrossRef]
  62. Wei, Z.; Zhang, Y.; Wang, S.; Wang, C.; Ma, J. Fe-doped phosphorene for the nitrogen reduction reaction. J. Mater. Chem. A 2018, 6, 13790–13796. [Google Scholar] [CrossRef]
  63. Qu, M.; Qin, G.; Fan, J.; Du, A.; Sun, Q. Theoretical insights into the performance of single and double transition metal atoms doped on N-graphenes for N2 electroreduction. Appl. Surf. Sci. 2021, 537, 148012. [Google Scholar] [CrossRef]
  64. DeBoer, J.H. The Dynamical Character of Adsorption. Soil Sci. 1953, 76, 166. [Google Scholar] [CrossRef]
Figure 1. The optimized structure of porous C24N24 originated from fullerene. Color code: C, gray; N, blue.
Figure 1. The optimized structure of porous C24N24 originated from fullerene. Color code: C, gray; N, blue.
Nanomaterials 11 01794 g001
Figure 2. A schematic representative of possible distribution patterns of two dissimilar TMs (TM1:TM2) in N4-pyridinic cavities of the C24N24 nanocage: (a) 6:0, (b) 5:1, (c,d) 4:2, (e,f) 3:3, (g,h) 2:4, (i) 1:5, (j) 0:6.Color code: white, Ti; orange, Cu.
Figure 2. A schematic representative of possible distribution patterns of two dissimilar TMs (TM1:TM2) in N4-pyridinic cavities of the C24N24 nanocage: (a) 6:0, (b) 5:1, (c,d) 4:2, (e,f) 3:3, (g,h) 2:4, (i) 1:5, (j) 0:6.Color code: white, Ti; orange, Cu.
Nanomaterials 11 01794 g002
Figure 3. The calculated binding energy (Eb) of TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24 (x, y, z = 0–6). a–j refer to the bi-metal configurations with different metal ratios listed in Table 1. All values are in eV.
Figure 3. The calculated binding energy (Eb) of TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24 (x, y, z = 0–6). a–j refer to the bi-metal configurations with different metal ratios listed in Table 1. All values are in eV.
Nanomaterials 11 01794 g003
Figure 4. NBO charge analysis of individual Ti, Mn, and Cu atoms in TixCuz@C24N24, TixMny@ C24N24, and MnyCuz@ C24N24. The a–j complexes refer to the metal distribution patterns presented in Figure 2. All values are in eV.
Figure 4. NBO charge analysis of individual Ti, Mn, and Cu atoms in TixCuz@C24N24, TixMny@ C24N24, and MnyCuz@ C24N24. The a–j complexes refer to the metal distribution patterns presented in Figure 2. All values are in eV.
Nanomaterials 11 01794 g004
Figure 5. (a) The calculated changes in free energy ( Δ G, at 298 K and 1 atm) (b) and energy gap for TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24 (x, y, z = 0–6). a–j refer to the bi-metal configurations with different metal ratios listed in Table 1.
Figure 5. (a) The calculated changes in free energy ( Δ G, at 298 K and 1 atm) (b) and energy gap for TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24 (x, y, z = 0–6). a–j refer to the bi-metal configurations with different metal ratios listed in Table 1.
Nanomaterials 11 01794 g005
Figure 6. The more active bi-metal configurations of TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24 (x, y, z = 0–6) toward CO2, NO2, H2, and N2 gas capture.
Figure 6. The more active bi-metal configurations of TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24 (x, y, z = 0–6) toward CO2, NO2, H2, and N2 gas capture.
Nanomaterials 11 01794 g006
Figure 7. The optimized configurations of CO2 and H2 on bi-metal complexes. All bond distances are in Å. Color code: white, Ti; orange, Cu; purple, Mn; red, O; grey, C; blue, N.
Figure 7. The optimized configurations of CO2 and H2 on bi-metal complexes. All bond distances are in Å. Color code: white, Ti; orange, Cu; purple, Mn; red, O; grey, C; blue, N.
Nanomaterials 11 01794 g007
Figure 8. The optimized configurations of N2 and N2O on bi-metal complexes. All bond distances are in Å. Color code: white, Ti; orange, Cu; purple, Mn; red, O; grey, C; blue, N.
Figure 8. The optimized configurations of N2 and N2O on bi-metal complexes. All bond distances are in Å. Color code: white, Ti; orange, Cu; purple, Mn; red, O; grey, C; blue, N.
Nanomaterials 11 01794 g008
Figure 9. The calculated lifetime of the adsorbed gas species on TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24.
Figure 9. The calculated lifetime of the adsorbed gas species on TixCuz@C24N24, TixMny@C24N24, and MnyCuz@C24N24.
Nanomaterials 11 01794 g009
Table 1. Distribution patterns of bi-metal atoms in C24N24 cavities.
Table 1. Distribution patterns of bi-metal atoms in C24N24 cavities.
Complex/Metal-Ratio6:05:14:23:32:41:50:6
abc, de, fg, hij
TixCuz@C24N24Ti6Ti5CuTi4Cu2Ti3Cu3Ti2Cu4TiCu5Cu6
TixMny@C24N24Ti6Ti5MnTi4Mn2Ti3Mn3Ti2Mn4TiMn5Mn6
MnyCuz@C24N24Mn6Mn5CuMn4Cu2Mn3Cu3Mn2Cu4MnCu5Cu6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nematollahi, P.; Neyts, E.C. Linking Bi-Metal Distribution Patterns in Porous Carbon Nitride Fullerene to Its Catalytic Activity toward Gas Adsorption. Nanomaterials 2021, 11, 1794. https://doi.org/10.3390/nano11071794

AMA Style

Nematollahi P, Neyts EC. Linking Bi-Metal Distribution Patterns in Porous Carbon Nitride Fullerene to Its Catalytic Activity toward Gas Adsorption. Nanomaterials. 2021; 11(7):1794. https://doi.org/10.3390/nano11071794

Chicago/Turabian Style

Nematollahi, Parisa, and Erik C. Neyts. 2021. "Linking Bi-Metal Distribution Patterns in Porous Carbon Nitride Fullerene to Its Catalytic Activity toward Gas Adsorption" Nanomaterials 11, no. 7: 1794. https://doi.org/10.3390/nano11071794

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