Nitrogen adsorption on Nb2C6H4+ cations: the important role of benzyne (ortho-C6H4)†
Received
14th November 2023
, Accepted 26th December 2023
First published on 4th January 2024
Abstract
N2 adsorption is a prerequisite for activation and transformation. Time-of-flight mass spectrometry experiments show that the Nb2C6H4+ cation, resulting from the gas-phase reaction of Nb2+ with C6H6, is more favorable for N2 adsorption than Nb+ and Nb2+ cations. Density functional theory calculations reveal the effect of the ortho-C6H4 ligand on N2 adsorption. In Nb2C6H4+, interactions between the Nb-4d and C-2p orbitals enable the Nb2+ cation to form coordination bonds with the ortho-C6H4 ligand. Although the ortho-C6H4 ligand in Nb2C6H4+ is not directly involved in the reaction, its presence increases the polarity of the cluster and brings the highest occupied molecular orbital (HOMO) closer to the lowest occupied molecular orbital (LUMO) of N2, thereby increasing the N2 adsorption energy, which effectively facilitates N2 adsorption and activation. This study provides fundamental insights into the mechanisms of N2 adsorption in “transition metal–organic ligand” systems.
1. Introduction
Dinitrogen (N2) is the most significant and abundant source of various N-containing compounds.1–5 However, the transformation of N2 at room temperature remains quite challenging due to its chemical inertness, the high HOMO–LUMO energy gap, large triple bond energy (945 kJ mol−1) and lack of a permanent dipole.6,7 In nature, nodule bacteria efficiently perform biological nitrogen fixation at ambient temperatures.8 In the condensed phase, enhancing N2 chemisorption capability on catalyst surfaces is critical for N2 reduction.9,10 Recently, metal–organic complexes, such as metal nitrogen heterocyclic compounds and multi-replacement metallocenes, have been reported to adsorb N2 and break N
N bonds under mild conditions.1,11–17 These unsaturated organic ligands act as electron donors to facilitate N2 adsorption. For example, a diketiminate-supported iron system that progressively activates benzene and N2 to produce aniline derivatives has been reported.18 Despite significant progress in N2 adsorption, the detailed reaction mechanisms and key influencing factors remain unclear. Furthermore, transition metal benzyne (TM-benzyne) complexes usually serve as reactive intermediates in organic synthetic chemistry.19–22 It is interesting to explore the reactivity of TM-benzyne toward N2.
Gas-phase studies on “isolated” reactants provide an ideal venue for investigating reaction mechanisms and kinetics at a strictly molecular level.7,23–26 Recently, some encouraging developments in N2 activation have been reported in the gas phase studies,27–31 and the ligand effect significantly affects cluster reactivity. Gas-phase clusters usually contain metal atoms coupled with inorganic ligands, such as oxygen, boron, hydrogen, carbon, sulfur, and so on.29,32–35 For example, we recently reported that in NbB3O2−, the B3O2 ligand and the single Nb center can form a dual active site to promote N2 activation and transformation.29 In the NbH2−/CO2/N2 system, the NbN2− anion is formed by the reaction of NbH2− with N2, and the N
N triple bond is cleaved with the formation of the C–N bond in the reaction of NbN2− with CO2.33 Notably, only limited studies are about N2 absorption and activation mediated by gas-phase ions containing organic ligands.36–39 He and coworkers recently reported that Fe2VC(C6H6)− can feasibly break the N
N triple bond.39 Considering the high reactivity of TM-benzyne in the condensed-phase reaction, it is crucial to synthesize isolable TM-benzyne ions and investigate their reactivity toward N2. Meanwhile, there is an urgent need to identify the interaction mechanism between metal atoms and organic ligands, which may influence the N2 adsorption rate.
In the present work, the gas-phase TM-benzyne cation Nb2C6H4+ is designed and synthesized through the reaction of Nb2+ and C6H6. We further investigated its reaction with N2. The presence of the ortho-C6H4 ligand increases the adsorption rate of N2 on Nb2C6H4+ compared to bare Nb2+. It is ascribed to the greater polarity and higher HOMO energy level in Nb2C6H4+, which favors N2 adsorption. Exploring typical clusters possessing metal atoms and organic ligands in the gas phase may provide a novel understanding for N2 adsorption.
2. Methods
2.1 Experimental methods
Interactions of ions and reactant gases are investigated using a homemade time-of-flight mass spectrometer (TOF-MS) equipped with a laser ablation ion source, a quadrupole mass filter (QMF), and a linear ion trap (LIT) reactor. Only a concise description of the apparatus is presented here, since details of the experiment have been previously covered in the earlier work.40–42 The Nb1,2+ cations are generated by laser ablation of a niobium disk target (99.999%) in helium carrier gas at a backing pressure of 3.0 standard atmospheres. A 532 nm laser with a repetition rate of 10 Hz as well as 5–8 mJ/pulse energy is employed. The Nb1,2+ cations are mass-selected by QMF, and then reacted with N2 seeded in the cooling gas (He) first. The reaction between Nb2+ and diluted benzene vapor seeded in the cooling gas (Ar) generates the Nb2C6H4+ cluster. N2 is then introduced into the LIT to investigate the N2 adsorption reaction. Finally, the LIT releases reactant and product ions, which are subsequently detected by reflection TOF-MS. The previous study showed that cations were thermally stabilized around (or just approaching) 298.15 K prior to various reactions.42
In our experiments, the concentrations of reactant molecules (C6H6 or N2) are much larger than that of reactant cations (Nb+, Nb2+, or Nb2C6H4+). Thus, the pseudo-first-order reaction model (k1) is assumed in each reaction. The following equation is applied to calculate the rate constant:42
|  | (1) |
in which
IR is the reactant intensity after the reaction and
IT equals the sum of both the reactant and product intensities. In addition,
k,
tR, and
T correspond to the Boltzmann constant, reaction time, and temperature, respectively.
Pe is the effective gas pressure of the reactant in the LIT. Finally, the least squares are used to derive the rate constant of each reaction.
43 The detailed method is described in ref.
43. The errors of
tR (±5%),
T (±2%), and
Pe (±20%) are also taken into account to estimate the error bars.
2.2 Computational methods
All density functional theory (DFT) calculations are carried out in the Gaussian 09 D.01 program package.44 The bond dissociation energies of Nb–Nb, Nb–C, Nb–N, N
N, C–C, C–H, C
N, H–H and N–H bonds are predicted by 20 tested methods, with the M06L functional agreeing well with the experimental data (Table S1, ESI†), and M06L is more adaptable for the group VB transition metal complexes45–47 and selected in the present work. Meanwhile, the density functional theory dispersion correction (DFT-D3) is included.48 The def2-TZVPP basis set is used for the Nb atom,49,50 and the 6-311+G** basis sets51–53 are chosen for the H, C, and N atoms. Each initial structure is optimized for various possible spin multiplicities. Flexible scanning and intrinsic reaction coordinates are used to achieve the optimal geometry of intermediates (Is) and transition states (TSs). Calculations of vibrational frequencies are used to make sure that Is and TSs have zero and only one imaginary frequency, respectively. In each reported energy (ΔH0K in eV), the zero-point vibration correction is taken into account. The rate constant of a reaction without an energy barrier is computed by variational transition state theory.54 NBO 6.0 is used for natural population analysis, which includes natural population analysis (NPA) charge, adaptive natural density partitioning (AdNDP) and percentage of orbital contribution.55 For the simulated density of states (DOS), the vertical detachment energies (VDEs) of the related clusters have been calculated using the same structures as their initial cations, corresponding to the energy difference between the cluster's Hartree–Fock energies and neutral species. Each peak is the result of the cation capturing an electron. Multiwfn 3.8 is applied to obtain localized orbital locator-π (LOL-π), electrostatic potential (ESP), interaction region indicator (IRI), DOS, and frontier molecular orbital.56–60
3. Results
3.1 Experimental results
As shown in Fig. 1, the Nb+, Nb2+, and Nb2C6H4+ cations generated by laser ablation are mass-selected and subsequently react with N2 in the LIT reactor. Note that Nb+ and Nb2+ cations show relatively low reactivity toward N2, Fig. 1b and d, reflected by the low intensities of NbN2+ and Nb2N2+ even at relatively high N2 pressures (up to 1.1 Pa and 0.8 Pa) and rather long reaction times (up to 13.4 ms and 14.9 ms). Interestingly, when the organic ligand ortho-C6H4 is attached on the Nb2+ cation, the adsorption rate is increased dramatically. The Nb2C6H4+ clusters can be produced by the mass-selected Nb2+ cations (Fig. 1c) reacting with C6H6 molecules. As the majority of Nb2+ is converted to Nb2C6H4+ (Fig. 1e, reaction (1)), N2 is pulsed into the LIT reactor. After the reaction with 1.3 Pa N2 for approximately 5.0 ms, a significant peak corresponding to Nb2C6H4N2+ is observed in Fig. 1f, suggesting reaction (2). In the mass spectra, the presence of NbO+, Nb2C6H4O+, and Nb2C6H4C6H6O+ peaks arises from reactions of residual water impurities with Nb+, Nb2C6H4+, and Nb2C6H4C6H6+, respectively.
 |
| Fig. 1 Time-of-flight (TOF) mass spectra for the reactions of mass-selected Nb+ (a) and Nb2+ (c) with He for 11.4 ms; Nb+ (b) and Nb2+ (d) with N2 for 13.4 ms and 14.9 ms, respectively; for the reaction of (e) the generated intermediate product Nb2C6H4+ with N2 (f) for 5.0 ms. The effective reactant gas pressures are shown. In panel (b), the signal magnitude is amplified by a factor of 40 (m/z = 121). | |
The pseudo-first-order rate constant (k1) values of reactions (1) and (2) are estimated to be (1.27 ± 0.26) × 10−9 cm3 molecule−1 s−1 and (2.60 ± 0.53) × 10−13 cm3 molecule−1 s−1, respectively, corresponding to the reaction efficiencies (Φ)61,62 of 126.7% and 0.04%. The results for Φ are consistent for both the surface charge capture and the average dipole orientation models. The rate for the Nb2C6H4+/N2 system is 10 times larger than that of the Nb2N2+/N2 [(2.64 ± 0.59) × 10−14 cm3 molecule−1 s−1], and the presence of the C6H4 ligand indeed increases the adsorption rate of N2. The signal dependence of product ions on N2 pressures and C6H6 are derived and fitted to the experimental data (Fig. 2 and Fig. S1, ESI†).
 |
| Fig. 2 Variations of the relative intensities of the reactant and product cations in the reactions of Nb2C6H4+ (a) and Nb2+ (b) with N2 with respect to the N2 pressures for 5 ms and 14.9 ms, respectively. The solid lines are fitted to the experimental data points by using the equations derived from the approximation of the pseudo-first-order reaction model. | |
3.2 Computational results
DFT calculations are performed to obtain the lowest-energy isomers of Nb2C6H6+ (Fig. S2a, ESI†) and Nb2C6H4+ (Fig. S3a, ESI†), and the dehydrogenation mechanism for reaction (1) is shown in Fig. 3a. Nb2+ has a quartet spin multiplicity.63 A spin crossover from the quartet potential energy surface (PES) to the doublet PES potentially takes place during the formation of I1 (Nb2C6H6+, ΔH0K = −2.61 eV, with respect to the separate reactants). The first hydrogen (H1) transfers from the C6H6 unit in I1 to the Nb atom in Nb2C6H6+viaTS1 to generate I2. Then I3 is formed by rotating H1 to the top of the linear Nb–Nb bond through TS2. The relative position of the Nb–Nb unit and C6H6 has been changed slightly. In the steps of I3 → I4 → I5, the Nb1–Nb2–H1 unit is further rotated relative to the C6H6 unit. It is emphasized that steps I3 → I4 → I5 are crucial to this reaction process because the direct C–H2 (second hydrogen) bond-breaking in I2 is difficult. From I5, the H2 migrates from the C6H6 unit to the Nb1 atom, leading to the formation of intermediate I6. ViaTS6, a planar H2–Nb1–Nb2–H1 unit in I7 is generated by rotating the H2 atom. Two hydrogen atoms bonded with Nb1 are combined to form one H2 unit in I8. Finally, Nb2C6H4+ and H2 are generated. Thus, the dehydrogenation reaction of Nb2C6H6+ cations is thermodynamically and kinetically favourable. As shown in Fig. 3b, the C2–C3 and Nb–Nb bonds in Nb2C6H4+ are elongated by approximately 20 pm and 47 pm, respectively, in comparison to the separate structures of C6H4 and Nb2+. In addition, we have considered benzyne (ortho-C6H4) and its isomers (meta-C6H4 and para-C6H4) in the Nb2C6H4+ structures (Fig. S3, ESI†), and Nb2C6H4+ given in Fig. 3a and b is the ground state. The dehydrogenation reaction of Nb2+ with C6H6 has been previously reported, but the detailed reaction mechanism was not given.64,65
 |
| Fig. 3 M06L-calculated potential energy surfaces (PESs) for the (a) Nb2+ + C6H6, and (c) Nb2C6H4+ + N2 reactions, and the spin multiplicities are doublet. The zero-point vibration-corrected energies (ΔH0K in eV) of the reaction intermediates, transition states, and products relative to the separated reactants are given. (b) Top (upper panel) and side (bottom panel) views of the optimized global minimum structure of the Nb2C6H4+ cluster. The blue arrow represents the vector direction of dipole moment. The superscripts indicate the spin multiplicities. The bond lengths are given in pm. The values of unpaired spin density distributions are given in parentheses. | |
The adaptive natural density partitioning (AdNDP) method is further applied to determine the bonding character of Nb2C6H4+. Two types of bonds exist in the Nb2C6H4+ cluster: the two-center two-electron (2c–2e) bond in Fig. S4 (ESI†) and the three-center two-electron (3c–2e) bond in Fig. 4a. There are two delocalized 3c–2e π-bonds located on the Nb–Nb–C (Nb–Nb–C2 and Nb–Nb–C3) and Nb–C–C (Nb–C1–C2, Nb–C3–C4, and Nb–C5–C6) units. The Nb2 atom carries most of the unpaired spin density (0.91 e) of the Nb2C6H4+ cluster, as shown in Fig. 3b, which provides an ideal site for N2 adsorption (I9 in Fig. 3c).
 |
| Fig. 4 (a) Adaptive natural density partitioning (AdNDP, unit: e) bonding analysis for three-center two-electron (3c–2e) bonds in the Nb2C6H4+ cluster. ON stands for the occupation number. The red and blue colours represent the positive value and negative value, respectively. (b) Two-dimensional localized orbital locator-π (LOL-π, unit: eV) analysis of the ortho-C6H4. A higher LOL-π value indicates a more localized distribution of π-electrons, and the colour depth increases with π-electron density. (c) Electrostatic potentials (ESPs, unit: eV) of the ortho-C6H4. The maximum and minimum values of the electrostatic potential are shown by the blue and orange spheres, respectively. (d) Interaction region indicator (IRI, unit: eV) of Nb2C6H4+. The isosurface map is 0.55. A stronger chemical strength is indicated by a lower value. | |
Nb2C6H4N2+ is observed in the reactions of Nb2C6H4+ and N2, and the calculated PES is given in Fig. 3c. The reaction pathway commences with the formation of encounter complex I9 (−0.92 eV), in which the incoming N2 molecule is adsorbed at the Nb2 atom via an η1-N2 coordination mode. Subsequently, the Nb1–N2 bond is formed in I10 (−1.56 eV) by changing the ∠NbNbN from 119° in I9 to 50° in I10. The most stable structure for Nb2C6H4N2+ is P1 (−1.98 eV), and the N2 unit is in a side-on end-on (η1:η2) coordination mode. The further N–N bond breaking is kinetically impeded by an intrinsic energy barrier TS11. Along the reaction coordinate (Nb2C6H4+ + N2 →→ P1) in Fig. 3b, the N
N bond length is significantly elongated from 109 pm in the free N2 molecule to 121 pm in Nb2C6H4N2+, suggesting that the N
N triple bond is activated.
4. Discussion
Based on the experimental results shown in Fig. 1b, Nb+ is almost inert toward N2, and this may be due to the fast dissociation rate of N2 desorption from the encounter complex (6.0 × 109 s−1 predicted by variational transition state theory). The dual-metal cation Nb2+ greatly enhances the N2 absorption rate, but this rate is still one order of magnitude smaller than that of N2/Nb2C6H4+. This interesting phenomenon gives rise to several questions: (1) What is the distinctive characteristic of the ortho-C6H4 ligand? (2) What is the interaction between Nb2+ and the ortho-C6H4 ligand? (3) Why does the ortho-C6H4 ligand facilitate N2 absorption?
4.1 Characteristic of the ortho-C6H4 ligand and interaction between Nb2+ and ortho-C6H4
Inspired by the conjugated π-bond in C6H6, the localized orbital locator-π (LOL-π) calculation is employed to visualize the distribution of π-electrons in the ortho-C6H4 ligand, as shown in Fig. 4b. The deeper colour in the red region indicates a higher distribution of π-electrons. When hydrogen is abstracted from C6H6, the C–C bond decreases from 139 pm in C6H6 to 125 pm in ortho-C6H4, leading to the increased π-electrons accumulation on this bond (Fig. 4b). Note that other C–C bonds in ortho-C6H4 are just as strong as those in C6H6. Consequently, the intense π-electrons create the negative electrostatic potential region, as depicted in Fig. 4c. The region with the lowest electrostatic potential (−0.65 eV) is situated on the C–C bond, which favors Nb2+ bonding.
Based on the natural population analysis (NPA) charges of the Nb2C6H4+ cation, the electron (0.72 e) transfers from the Nb2+ unit to the ortho-C6H4 ligand. The interaction region indicator (IRI) analysis (Fig. 4d) reveals the formation of a strong chemical bond (green region) between Nb and C atoms. The bonding character and mechanism are further examined using the density of states (DOS) and the frontier orbital analyses, as shown in Fig. 5 and Fig. S6 (ESI†). In Fig. 5a, the 4d-electrons of Nb dominate the DOS around the HOMO. The Nb-4d and C-2p orbitals have strong interaction in the Nb–C bond, as demonstrated by the β-HOMO orbital, as well as the HOMO−1, HOMO−2, and HOMO−3 in both α and β orbitals.
 |
| Fig. 5 Total density of states (TDOS, black line) for (a) Nb2C6H4+ and (b) Nb2C6H4N2+. Projected density of states (PDOS) of Nb (red line) and N2 (purple line) are also given. In each panel, the HOMO position is indicated by the blue dotted line. The HOMO–LUMO energy gaps are also given. The orbital insets are shown to illustrate the bond interactions between the Nb2C6H4+ cluster and N2. The corresponding orbital contributions for Nb, C, and N are also presented. (c) The HOMO and LUMO energy levels of Nb2+, Nb2C6H4+, and N2, respectively. | |
Notably, several significant changes are induced when the Nb2+ bonds with ortho-C6H4. (1) The unpaired spin density (UPSD) distribution is adjusted. In 4Nb2+, the UPSD is evenly distributed over each Nb atom (1.50 e), whereas the UPSD is localized on one Nb atom (0.91 e) in 2Nb2C6H4+. (2) A polar Nb2C6H4+ cluster is formed. The dipole moment is increased from 0 D for Nb2+ to 3.5 D for Nb2C6H4+. The vector direction in Nb2C6H4+ (the blue arrow in Nb2C6H4+, Fig. 3b) is along an axis perpendicular to the Nb–Nb bond, lying in the Nb1–O–Nb2 plane. Large dipole moment favors electron transfer and N2 adsorption. (3) The HOMO energy level is raised from −15.87 eV in Nb2+ to −14.42 eV in Nb2C6H4+ (Fig. 5c), which is closer to the LUMO of N2 (−1.76 eV).
4.2 Interaction between Nb2C6H4+ and N2
In Nb2C6H4N2+ and Nb2N2+, N2 molecules are stably adsorbed on the Nb atoms in terms of side-on end-on (η1:η2) coordination mode. The adsorption energy of N2 on Nb2C6H4+ (−1.98 eV) is larger than that of Nb2+/N2 (−1.01 eV), which is consistent with experimental observations.
As shown in Fig. 5a and Fig. S6a (ESI†), the HOMO of Nb2C6H4+ is closer to the LUMO in N2, compared with that of the Nb2+ cation, resulting in a higher reactivity towards N2. The frontier orbital analyses for Nb2N2+ and Nb2C6H4N2+ cations are performed in Fig. 5b and Fig. S6b (ESI†). As electron donors, Nb atoms directly interact with N2, and the obvious orbital interaction between Nb-4d and N-2p is shown in Fig. 5b (e.g., the α-HOMO−1 and α-HOMO−3) and Fig. S6b (ESI,†e.g., the α-HOMO−3 and α-HOMO−6). The back-donating electrons from the Nb-4d orbital to the antibonding π*-orbital of N2 activate the N
N triple bond. The key factor in enhancing the N2 adsorption rate stems from the addition of the ortho-C6H4 ligand, forming a polar Nb2C6H4+ cluster. The positive-N1 (0.14 e) and negative-N2 (−0.15 e) charges in I9 indicate that N2 has been slightly polarized. A strong interaction occurs between the polarized N2 and the polar Nb2C6H4+. In the condensed-phase system, increasing the polarity of the solid surface can increase the adsorption capacity of polar gas molecules.66,67 Therefore, increasing the polarity of gas-phase clusters favors non-polar N2 adsorption.
In Fig. 6, the natural population analysis charges of Nb, and N atoms, as well as the ortho-C6H4 ligand along the reaction coordinates of the Nb2C6H4+/N2 system are carried out. In the course of R →→ P1, two Nb atoms and the ortho-C6H4 ligand act as an electron reservoir, transferring 0.53 e to N2. In the formation of encounter complex I9 (R → I9), the N1 atom releases negative charges (ΔQ = 0.14 e) to the Nb2 and N2 atoms, forming the Nb2–N2 bond. Then negative charges flow from two Nb atoms to the N2 unit (ΔQ = 0.52 e), leading to the formation of the Nb1–N2 bond (I9 → I10). In the last step for the formation of the Nb1–N1 bond (I10 → P1), the N2 atom donates 0.16 e to the Nb1 and N1 atoms. In a word, although the ortho-C6H4 ligand in Nb2C6H4+ seems like a spectator and does not contribute electrons to the reaction, its presence is significant in increasing polarity, raising the HOMO energy level, and increasing the N2 adsorption energy. These behaviors facilitate N2 adsorption. In addition, Nb2C6H4N2+ has more vibrational degrees of freedom than Nb2N2+ does, and therefore the former has more chance to be stabilized in the experiment.
 |
| Fig. 6 Natural population analysis (NPA) charges on the Nb, N atoms, and ortho-C6H4 ligand along reaction coordinates of N2 activation on the Nb2C6H4+ cation. | |
5. Conclusions
In summary, the gas-phase reaction of Nb2+ with C6H6 is identified, which releases one H2 molecule. The N2 adsorption activity mediated by three different cations, that is, Nb+, Nb2+, and Nb2C6H4+, under thermal collision conditions has been investigated using mass spectrometry experiments and theoretical calculations. The experimental results indicate that the reaction rate constant of Nb2C6H4+/N2 is 10 times faster than that of Nb2+/N2, suggesting that organic ligand ortho-C6H4 facilitates N2 adsorption, while the reaction rate of Nb+/N2 is quite low under similar reaction conditions. Using DFT calculations, the electron transfer, chemical bonding, and orbital interaction mechanisms have been investigated between Nb2+ and ortho-C6H4. In Nb2C6H4+, the enhanced polarity as well as the raised HOMO level facilitate N2 adsorption. The present gas-phase study provides molecular level insight into the mechanism of N2 adsorption mediated by metal active sites with organic ligands.
Author contributions
Feng-Xiang Zhang: conceptualization, data curation, formal analysis, investigation, methodology, software, validation, visualization, writing – original draft and writing – review and editing. Yi-Heng Zhang and Ming Wang: data curation, formal analysis, methodology and validation. Jia-Bi Ma: conceptualization, formal analysis, funding acquisition, project administration, resources, supervision and writing – review and editing.
Conflicts of interest
The authors declare no conflict of interest.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (22222301), the Beijing Natural Science Foundation (2222023), the National Key Research and Development Program of China (2021YFA1601800), and the BIT Research and Innovation Promoting Project (2022YCXY029).
Notes and references
- Q. Zhuo, X. Zhou, T. Shima and Z. Hou, Angew. Chem., Int. Ed., 2023, 62, e202218606 CrossRef CAS PubMed.
- T. T. Liu, D. D. Zhai, B. T. Guan and Z. J. Shi, Chem. Soc. Rev., 2022, 51, 3846–3861 RSC.
- L. J. Wu, Q. Wang, J. Guo, J. Wei, P. Chen and Z. Xi, Angew. Chem., Int. Ed., 2023, 62, e202219298 CrossRef CAS PubMed.
- T. Itabashi, K. Arashiba, A. Egi, H. Tanaka, K. Sugiyama, S. Suginome, S. Kuriyama, K. Yoshizawa and Y. Nishibayashi, Nat. Commun., 2022, 13, 6161 CrossRef CAS PubMed.
- X. Shi, Q. Wang, C. Qin, L. J. Wu, Y. Chen, G. X. Wang, Y. Cai, W. Gao, T. He, J. Wei, J. Guo, P. Chen and Z. Xi, Natl. Sci. Rev., 2022, 9, nwac168 CrossRef CAS PubMed.
- R. J. Burford and M. D. Fryzuk, Nat. Rev. Chem., 2017, 1, 1–13 CrossRef.
- L.-H. Mou, Z.-Y. Li and S.-G. He, J. Phys. Chem. Lett., 2022, 13, 4159–4169 CrossRef CAS PubMed.
- S. Jiang, M.-F. Jardinaud, J. Gao, Y. Pecrix, J. Wen, K. Mysore, P. Xu, S.-C. Carmen, Y. Ruan, M. Zhu, F. Li, E. Wang, P. Poole, P. Gamas and J. D. Murray, Science, 2021, 374, 625–628 CrossRef CAS PubMed.
- M. Wang, J. Ma, Z. Shang, L. Fu, H. Zhang, M. B. Li and K. Lu, J. Mater. Chem. A, 2023, 11, 3871–3887 RSC.
- X. Gao, L. An, D. Qu, W. Jiang, Y. Chai, S. Sun, X. Liu and Z. Sun, Sci. Bull., 2019, 64, 918–925 CrossRef CAS PubMed.
- M. Reiners, D. Baabe, K. Munster, M. K. Zaretzke, M. Freytag, P. G. Jones, Y. Coppel, S. Bontemps, I. D. Rosal, L. Maron and M. D. Walter, Nat. Chem., 2020, 12, 740–746 CrossRef PubMed.
- P. Garrido-Barros, J. Derosa, M. J. Chalkley and J. C. Peters, Nature, 2022, 609, 71–76 CrossRef CAS PubMed.
- H. J. Himmel and M. Reiher, Angew. Chem., Int. Ed., 2006, 45, 6264–6288 CrossRef CAS PubMed.
- X. Xin, I. Douair, Y. Zhao, S. Wang, L. Maron and C. Zhu, J. Am. Chem. Soc., 2020, 142, 15004–15011 CrossRef CAS PubMed.
- Z. J. Lv, Z. Huang, W. X. Zhang and Z. Xi, J. Am. Chem. Soc., 2019, 141, 8773–8777 CrossRef CAS PubMed.
- Y. Ohki, K. Munakata, Y. Matsuoka, R. Hara, M. Kachi, K. Uchida, M. Tada, R. E. Cramer, W. M. C. Sameera, T. Takayama, Y. Sakai, S. Kuriyama, Y. Nishibayashi and K. Tanifuji, Nature, 2022, 607, 86–90 CrossRef CAS PubMed.
- C. J. P. J. Talarmin, Nature, 1985, 317, 652–653 CrossRef.
- S. F. McWilliams, D. L. J. Broere, C. J. V. Halliday, S. M. Bhutto, Q. M. Brandon and P. L. Holland, Nature, 2020, 584, 221–226 CrossRef CAS PubMed.
- S. L. Buchwald, B. T. Watson, R. T. Lum and W. A. Nugent, J. Am. Chem. Soc., 1987, 109, 7137–7141 CrossRef.
- S. L. Buchwald and R. B. Nielsen, Chem. Rev., 1988, 88, 1047–1058 CrossRef CAS.
- S. J. McLain, R. R. Schrock, P. R. Sharp, M. R. Churchill and W. J. Youngs, J. Am. Chem. Soc., 1979, 101, 263–265 CrossRef CAS.
- X. Chen, Z. Xiong, M. Yang and Y. Gong, Chem. Commun., 2022, 58, 7018 RSC.
- B. Yuan, Z. Liu, X.-N. Wu and S. Zhou, Sci. China: Chem., 2022, 65, 1720–1724 CrossRef CAS.
- Y.-Q. Ding, F. Ying, Y. Li, J. Xie and J.-B. Ma, Inorg. Chem., 2023, 62, 6102–6108 CrossRef CAS PubMed.
- X. Sun, S. Zhou, L. Yue, C. Guo, M. Schlangen and H. Schwarz, Angew. Chem., Int. Ed., 2019, 58, 3635–3639 CrossRef CAS PubMed.
- R. A. J. O’Hair and G. N. Khairallah, J. Cluster Sci., 2004, 15, 331–363 CrossRef.
- Y. Zhao, J.-T. Cui, M. Wang, D. Y. Valdivielso, A. Fielicke, L.-R. Hu, X. Cheng, Q.-Y. Liu, Z.-Y. Li, S.-G. He and J.-B. Ma, J. Am. Chem. Soc., 2019, 141, 12592–12600 CrossRef CAS PubMed.
- Y.-Q. Ding, Z.-Y. Chen, Z.-Y. Li, X. Cheng, M. Wang and J.-B. Ma, J. Phys. Chem. A, 2022, 126, 1511–1517 CrossRef CAS PubMed.
- M. Wang, H.-Y. Zhou, A. M. Messinis, L.-Y. Chu, Y. Li and J.-B. Ma, J. Phys. Chem. Lett., 2021, 12, 6313–6319 CrossRef CAS PubMed.
- Y.-Q. Ding, Y. Li, F. Ying, M. Wang and J.-B. Ma, J. Phys. Chem. Lett., 2022, 13, 492–497 CrossRef CAS PubMed.
- L.-Y. Chu, Y.-Q. Ding, M. Wang and J.-B. Ma, Phys. Chem. Chem. Phys., 2022, 24, 14333–14338 RSC.
- M. Gao, Y.-Q. Ding and J.-B. Ma, Int. J. Mol. Sci., 2022, 23, 6976 CrossRef CAS PubMed.
- M. Wang, L.-Y. Chu, Z.-Y. Li, A. M. Messinis, Y.-Q. Ding, L. Hu and J.-B. Ma, J. Phys. Chem. Lett., 2021, 12, 3490–3496 CrossRef CAS PubMed.
- G.-D. Jiang, Z.-Y. Li, L.-H. Mou and S.-G. He, J. Phys. Chem. Lett., 2021, 12, 9269–9274 CrossRef CAS PubMed.
- L.-H. Mou, G.-D. Jiang, Z.-Y. Li and S.-G. He, Chin. J. Chem. Phys., 2020, 33, 507–520 CrossRef CAS.
- V. I. Baranov, G. Javahery and D. K. Bohme, Chem. Phys. Lett., 1995, 239, 339–343 CrossRef CAS.
- R. K. Milburn, V. Baranov, A. C. Hopkinson and D. K. Bohme, J. Phys. Chem. A, 1999, 103, 6373–6382 CrossRef CAS.
- D. Caraiman and D. K. Bohme, Int. J. Mass Spectrom., 2003, 223–224, 411–425 CrossRef CAS.
- X. Cheng, Z.-Y. Li, G.-D. Jiang, X.-X. Liu, Q.-Y. Liu and S.-G. He, J. Phys. Chem. Lett., 2023, 14, 6431–6436 CrossRef CAS PubMed.
- Y. Zhao, J.-C. Hu, J.-T. Cui, L.-L. Xu and J.-B. Ma, Chem. – Eur. J., 2018, 24, 5920–5926 CrossRef CAS PubMed.
- H.-Y. Zhou, M. Wang, Y.-Q. Ding and J.-B. Ma, Dalton Trans., 2020, 49, 14081 RSC.
- Z. Yuan, Z.-Y. Li, Z.-X. Zhou, Q.-Y. Liu, Y.-X. Zhao and S.-G. He, J. Phys. Chem. C, 2014, 118, 14967–14976 CrossRef CAS.
- Z.-Y. Li, Z. Yuan, X.-N. Li, Y.-X. Zhao and S.-G. He, J. Am. Chem. Soc., 2014, 136, 14307–14313 CrossRef CAS PubMed.
-
M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2013 Search PubMed.
- Y. Zhao and D. G. Truhlar, J. Chem. Phys., 2006, 125, 194101 CrossRef PubMed.
- Y. Li, M. Wang, Y.-Q. Ding, C.-Y. Zhao and J.-B. Ma, Phys. Chem. Chem. Phys., 2021, 23, 12592–12599 RSC.
- Q. Chen, Y.-X. Zhao, L.-X. Jiang, H.-F. Li, J.-J. Chen, T. Zhang, Q.-Y. Liu and S.-G. He, Phys. Chem. Chem. Phys., 2018, 20, 4641–4645 RSC.
- S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
- B. Attila, A. M. Steven and Z. Z. Marek, J. Phys. Chem. A, 1998, 102, 6340–6347 CrossRef.
- F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys., 2005, 7, 3297–3305 RSC.
- R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople, J. Chem. Phys., 1980, 72, 650–654 CrossRef CAS.
- T. Clark, J. Chandrasekhar, G. W. Spitznagel and P. V. R. Schleyer, J. Comput. Chem., 1983, 4, 294–301 CrossRef CAS.
- R. Ditchfield, W. J. Hehre and J. A. Pople, J. Chem. Phys., 1971, 54, 724–728 CrossRef CAS.
- J. L. Bao and D. G. Truhlar, Chem. Soc. Rev., 2017, 46, 7548–7596 RSC.
- E. D. Glendening, C. R. Landis and F. Weinhold, J. Comput. Chem., 2013, 34, 1429–1437 CrossRef CAS PubMed.
- T. Lu and F. Chen, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed.
- T. Lu and Q. Chen, Theor. Chem. Acc., 2020, 139, 25 Search PubMed.
- T. Lu and S. Manzetti, Struct. Chem., 2014, 25, 1521–1533 CrossRef CAS.
- T. Lu and Q. Chen, Chem.: Methods, 2021, 1, 231 CAS.
- Z. Liu, T. Lu and Q. Chen, Carbon, 2020, 165, 461–467 CrossRef CAS.
- G. Gioumousis and D. P. Stevenson, J. Chem. Phys., 1958, 29, 294–299 CrossRef CAS.
- G. Kummerlöwe and M. K. Beyer, Int. J. Mass Spectrom., 2005, 244, 84–90 CrossRef.
- K. Balasubramanian, J. Chem. Phys., 1995, 114, 10375–10388 CrossRef.
- C. Berg, T. Schindler, G. Niedner-Schatteburg and V. E. Bondybey, J. Chem. Phys., 1995, 102, 4870–4884 CrossRef CAS.
- S. Roszak, D. Majumdara and K. Balasubramaniana, J. Phys. Chem. A, 1999, 103, 5801–5806 CrossRef CAS.
- F. Mohajer and M. N. Shahrak, Heat Mass Transfer, 2019, 55, 2017–2023 CrossRef CAS.
- M. R. A. Tourrette, R. H. R. Valentin, M. Boissière, J. M. Devoisselle, F. D. Renzo and F. Quignard, Carbohydr. Polym., 2011, 85, 44–53 CrossRef.
|
This journal is © the Owner Societies 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.