Observation of double-ring tubular B20(CO)n+ (n = 1–8): emergence of 2D-to-3D transition in boron carbonyl complexes
Received
15th May 2025
, Accepted 24th July 2025
First published on 25th July 2025
Abstract
The consecutive discoveries of double-ring (DR) tubular D10d B20 and D2d B20+ as embryos of single-walled boron nanotubes have attracted considerable attention in the past two decades. Joint chemisorption experiments and first-principles theory investigations performed herein indicate that, as the only isomer of the monocation observed in gas-phase experiments, DR tubular D2d B20+ can react with CO successively under ambient conditions to form a series of DR tubular boron carbonyl monocations B20(CO)n+ up to n = 8, presenting the largest boron carbonyl complexes observed to date, which mark the 2D-to-3D transition in boron carbonyl complexes. DR tubular D2d B20+ with twenty peripheral boron atoms is found to be about ten times more reactive to chemisorb the first CO than the experimentally known quasi-planar C2v B13+ (B3@B10+) with typical π-aromaticity analogous to benzene's but about ten times less reactive than both quasi-planar Cs B11+ (B2@B9+) and C2v B15+ (B4@B11+) with σ and π conflicting aromaticity. Extensive theoretical calculations and analyses unveil the chemisorption pathways, potential energy profiles, and chemical bonding patterns of DR tubular B20(CO)n+ and its neutral counterpart B20(CO)n, both of which appear to be tubularly aromatic in nature.
Introduction
Carbon monoxide (CO) is one of the most important ligands in chemistry due to its unique capacity to form σ-donation and π-back-donation coordination interactions in metal carbonyl complexes.1–3 In typical transition-metal (TM) carbonyl complexes, the :C
O ligand donates its σ-HOMO lone-pair on carbon to the partially occupied (n − 1)d orbitals of the TM center while concurrently accepting back donation from the (n − 1)d orbitals of the TM center in its two degenerate π*-LUMO antibonding orbitals. σ + π dual coordination interactions dominate the electronic structures and reactivities of TM carbonyl complexes. Main group metal carbonyl complexes have also been discovered in the past decade, including the cubic alkaline-earth metal carbonyl complexes M(CO)8 (M = Ca, Sr, and Ba) and Be(CO)n (n = 1–4), in which the alkaline-earth metals serve as “honorable transition metals” exhibiting transition-metal-like behaviors.4,5
As a prototypical electron-deficient element in the periodic table, boron (He[2s22p1]) can also form various types of carbonyl complexes. The closed-shell carbonyl borane H3BCO and its derivatives are experimentally known as stable boron carbonyl complexes.6 In recent years, various small boron carbonyl clusters, including linear BCO,7 OCBBCO,8 and BBCO,9 V-shaped B(CO)2,10 rhombic B4(CO)2,11 and planar (2D) B(CO)3+, B(CO)4+, B2(CO)4+, and B3(CO)n+ (n = 3–6), have been observed in experiments using infrared photodissociation (IRPD) spectroscopy.12–15 Neutral rhombic B4(CO)3 was also observed in the gas phase.16 In these boron carbonyl complexes, CO ligands serve as donors to coordinate Bn centres as acceptors via effective σ-donations. Based on joint gas-phase mass spectroscopy, collision-induced dissociation (CID), and first-principles theory investigations, our group observed in 2024 the first quasi-planar boron carbonyl aromatics (BCAs) B13(CO)n+ (n = 1–7), analogous to benzene C6H6 under ambient conditions.17 CID experiments confirmed that CO ligands are molecularly coordinated to the aromatic B13+ core in these BCAs without being activated or disassociated. Similarly, in 2025, quasi-planar B11(CO)n+ (n = 1–6) and B15(CO)n+ (n = 1–5) with π and σ conflicting aromaticity were observed in the gas phase, presenting the largest boron carbonyl complexes observed to date.18 Neutral double-ring tubular B20(CO)n (n = 1–8) complexes based on the framework of aromatic DR tubular B20 have also been predicted in theory recently.19 However, to the best of our knowledge, there have been no three-dimensional (3D) boron carbonyl complexes observed to date in experiments, leaving an important question to be addressed in this area.
Based on the fact that neutral B20 has a perfect double-ring (DR) tubular D10d structure as its well-defined global minimum (GM),20–22 B20+ monocation possesses a slightly distorted DR tubular D2d configuration as its deep-lying GM, which was the only stable isomer of the system observed in ion-mobility experiments.23 While both of them can be considered as the embryos of single-walled boron nanotubes, we report herein the observation of a series of DR tubular boron carbonyl monocations B20(CO)n+ (n = 1–8) in gas phase under ambient conditions, as revealed by detailed chemisorption experiments and extensive first-principles theory calculations, presenting the largest boron carbonyl complexes observed to date, which mark the 2D-to-3D structural transition in boron carbonyl complexes. DR tubular D2d B20+ with twenty periphery boron atoms was found to be much more reactive to chemisorb the first CO than the previously reported planar π-aromatic C2v B13+, but much less reactive than both 2D Cs B11+ (B2@B9+) and C2v B15+ (B4@B11+) with σ and π conflicting aromaticity. Detailed theoretical calculations and analyses reveal the chemisorption mechanisms and bonding patterns of the experimentally observed DR tubular B20(CO)n+ monocations and their neutral DR tubular counterparts B20(CO)n (n = 1–8), with both appearing to be aromatic in nature.
Methods
Experimental methods
A homemade reflection time-of-flight mass spectrometer (TOF-MS)17,18,24,25 equipped with a laser ablation cluster source, a quadrupole mass filter (QMF),26 and a linear ion trap (LIT)27 reactor was employed in this work. Bare boron cluster monocations (Bn+) were generated by laser ablation of a rotating and translating 11B target (99% enriched) under a 6 atm He carrier gas. A 532 nm pulsed laser (Nd3+:YAG second harmonic) with an energy of 3–5 mJ per pulse and a repetition rate of 10 Hz was used to ablate the boron target.
B20+ monocations were mass-selected by the QMF and entered into the LIT, where they were confined and thermalized by collisions with a pulse of He gas and then interacted with a pulse of CO reactant gas. The cluster ions ejected from the LIT reactor were detected using TOF-MS. The pseudo-first-order rate constants (k1) of the reaction between B20+ and CO were determined using the following equation,
|
 | (1) |
in which
IR is the signal intensity of the reactant cluster ions after the reaction,
IT is the total ion intensity including contributions from the products,
Peffective is the effective pressure of the reactant gas in the ion trap reactor,
kB is the Boltzmann constant,
T is the temperature (∼300 K), and
tR is the reaction time. More details about the method to derive
k1 can be found in ref.
28.
Theoretical methods
Extensive density functional theory (DFT) calculations at the PBE0/6-311+G(d) level29,30 were performed using the Gaussian 16 program.31 Previous CID experiments on BCAs B13(CO)n+ indicate that CO ligands in boron carbonyl complexes are coordinated molecularly to the peripheral atoms of the Bm+ cores, without being activated or dissociated.17 Using CO as molecular ligands to coordinate the experimentally observed DR tubular D2d B20+ core, which is the only stable isomer of the monocation existing in the gas phase,23 the structures of the B20(CO)n+ (n = 1–8) boron carbonyl complexes, intermediates (IMs), and transition states (TSs) were extensively manually constructed and fully optimized in this work. The hybrid PBE0 functional, which has proven to be reliable for boron cluster calculations,17,18,32–34 was employed in this work. Vibrational frequency analyses confirmed that all IMs and TSs were true minima and transition states of the systems, respectively. Intrinsic reaction coordinate (IRC) calculations35,36 were further conducted to verify that each TS with one imaginary vibrational frequency (as indicated in Fig. S10–S21) connects two appropriate IMs. Based on the fact that DR tubular D2d B20+ is the only isomer of the monocation observed in ion-mobility experiments23 and CO ligands are molecularly coordinated to the Bm+/0 cores in CID measurements,17 the optimized structures obtained herein by extensive manual structural constructions based on D2d B20+ are believed to be reliable, though no global structural optimization and direct IRPD spectroscopic comparison were performed at the current stage. PBE0-D3/6-311+G(d) structural optimizations with dispersion correction (Becke–Johnson damping D3 correction)37 included were also performed to compare with the results obtained using PBE0/6-311+G(d). More accurate single-point DLPNO-CCSD(T)38 calculations were performed on the most favorable chemisorption pathways of the experimentally observed B20(CO)n+ (n = 1–8) monocations using the ORCA program.39 Adaptive natural density partitioning (AdNDP 2.0) bonding analyses40,41 were performed on the concerned species. Iso-chemical shielding surfaces (ICSSs) were computed using the Multiwfn 3.8 program42 and visualized using the VMD 1.9.3 software.43 Detailed anisotropies of the current-induced density (ACID) analyses were realized using the ACID 2.0 code44 to further check the tubular aromaticity of the systems, with the ring current maps finally generated using POV-Ray 3.7 render.45 In addition, energy decomposition analyses with natural orbitals for chemical valence (EDA-NOCV)46–48 were performed using the ADF(2023) program package49 at the PBE0/TZ2P level combined with the zeroth-order regular approximation (ZORA)50 to elucidate the coordination bonding characteristics of B20(CO)n+/0.
Results and discussion
Cluster reactivity measurements
The TOF mass spectra of the reactions of B20+ with CO in Fig. 1A indicate that B20+ can consecutively chemisorb up to eight CO ligands under ambient conditions to form the B20(CO)n+ (n = 1–8) complex series, presenting the highest coordination number of n = 8 in boron carbonyl complexes observed to date. Upon reacting B20+ with 356 mPa CO for 2 ms, distinct signals of B20CO+ and B20(CO)2+ corresponding to the first and second CO adsorptions were clearly observed, respectively (Fig. 1(a2)). When the CO pressure was increased to 893 mPa (Fig. 1(a3)), the product B20(CO)3+ emerged. Increasing the CO pressure to 1276 mPa (Fig. 1(a4)), the signal intensities of both B20+ and B20CO+ were significantly reduced, with a weak signal appearing at B20(CO)4+. To further investigate the chemisorption capacity of B20+ toward CO, the reaction time was extended to 10 ms (Fig. 1(a5)). Under these conditions, the reactant signal of B20+ and product signal of B20CO+ disappeared completely, the B20(CO)2+ signal was obviously weakened, while the relative intensity of B20(CO)4+ increased effectively and new product signals of both B20(CO)5+ and B20(CO)6+ started to emerge. These results indicate that no inert isomers of B20+ exist in our experiments. Further prolonging the reaction time to 20 ms (Fig. 1(a6)) and 30 ms (Fig. 1(a7)) led to a notable enhancement of the B20(CO)6+ signal and the ultimate emergence of both B20(CO)7+ and B20(CO)8+. No signal beyond, corresponding to the ninth CO adsorption product (B20(CO)9+), was detected in our experiments (Fig. 1(a7)). The experimental results observed above show that B20+ can successively chemisorb up to eight CO molecules in the following chemisorption reaction: |
B20+ + nCO → B20(CO)n+ (n = 1–8)
| (2) |
 |
| Fig. 1 (A) Measured TOF mass spectra for the reactions of mass-selected B20+ with nCO (n = 1–8) and variations in the measured relative signal intensities of the reactant and product ions with respect to CO gas pressures in the reactions of (B) B20+ + nCO → B20(CO)n+ (n = 1–2) and (C) B20(CO)2+ + (n − 2)CO → B20(CO)n+ (n = 3–6), with the solid lines fitted to experimental data points with the approximation of pseudo-first-order reaction mechanisms. | |
Based on detailed chemisorption measurements and the least-squares fitting procedure of eqn (1), the pseudo-first-order rate constants (k1) for the reactions of B20+ + nCO → B20(CO)n+ (n = 1–2) were estimated to be k1 = (2.65 ± 0.53) × 10−11 [ϕ = (1.91 ± 0.38)%] and (1.84 ± 0. 37) × 10−11 [ϕ = (1.33 ± 0.27)%] cm3 molecule−1 s−1 for n = 1 and 2, respectively (Fig. 1B). For consecutive chemisorptions of B20(CO)2+ with more CO molecules to form B20(CO)n+ (n = 3–6), the k1 values were estimated to be k1 = (5.76 ± 1.15) × 10−11, (1.10 ± 0.22) × 10−11, (7.64 ± 1.53) × 10−13, and (1.95 ± 0.39) × 10−12 cm3 molecule−1 s−1 for n = 3, 4, 5, and 6, with the corresponding reaction efficiencies of ϕ = (4.15 ± 0.83)%, (0.79 ± 0.16)%, (0.06 ± 0.01)%, and (0.14 ± 0.03)%, respectively (Fig. 1C). It is noticed that B20(CO)4+ possesses the highest mass intensity in Fig. 1(a4)–(a7) in the B20(CO)n+ (n = 2–8) series and exhibits the most abundant accumulative intensity in Fig. 1C in the B20(CO)n+ (n = 2–6) series. These kinetic results suggest that B20(CO)4+ possesses a unique structure, which is chemically more stable than its neighbours and relatively less reactive to chemisorb more CO molecules. Previous ion mobility measurements23 indicate DR tubular D2d B20+ is the well-defined GM of the monocation and the only stable isomer observed in the gas phase. The k1 value estimated here shows that the DR tubular D2d B20+ is about ten times more reactive to chemisorb the first CO than the typical planar π-aromatic C2v B13+ analogous to benzene,17 but about ten times less reactive than both quasi-planar Cs B11+ and C2v B15+ with σ and π conflicting aromaticity,18 suggesting that relative reactivities of the concerned Bn+ towards the first CO are closely related to the geometries and aromaticities of the systems.
Chemisorption pathway analyses
The optimized lowest-lying isomers and corresponding most favourable chemisorption pathways of B20(CO)n+ (n = 1–8) monocations and B20(CO)n (n = 1–8) neutrals are depicted in Fig. 2A and C, respectively. Alternative low-lying isomers and minor chemisorption pathways starting from local minima, which are slightly less stable than the corresponding lowest-lying isomers on the most favourable chemisorption pathways, are collectively shown in Fig. S1–S6 and Fig. S8–S21, respectively. The CO molecular ligands appear to prefer to be chemisorbed along both the top and bottom B10 rings of DR tubular B20+/0 at two opposite sides. As shown in Fig. 2A, the first CO is coordinated to the top B10 ring of DR tubular B20+ without an energy barrier, forming the most stable adduct Cs B20(CO)+ (1A) with a chemisorption energy of 1.08 eV. Therefore, the observed B20(CO)+ mass signal should correspond to Cs B20(CO)+ (1A), which represents the most favourable configuration in both thermodynamics and dynamics. Similarly, the experimentally observed B20(CO)2+, B20(CO)3+, and B20(CO)4+ can be assigned to C2 2A, C1 3A, and Cs 4A (Fig. S8), which possess the most favourable chemisorption energies, respectively, via barrierless chemisorption processes. We notice that the experimentally observed Cs B20(CO)4+ (4A) can also be formed from the third lowest-lying isomer C1 B20(CO)3+ (3C) in a barrierless minor chemisorption process (Fig. S9). Notably, with the largest chemisorption energy of 1.31 eV (Fig. 2A) and lowest reaction rate of k1 = 7.64 × 10−13 cm3 molecule−1 s−1, the observed Cs B20(CO)4+ (4A) with four CO ligands symmetrically distributed at two opposite sites exhibits the highest mass intensity in the B20(CO)n+ (n = 1–8) series (Fig. 1(a4)–(a7)). The formation of B20(CO)5+ (5A) proceeds through the transition state TS1, which lies 0.07 eV lower than the entrance channel (Fig. 3A and Fig. S10). Further analysis reveals that the adsorption of the sixth CO molecule to form the most stable adduct C2 B20(CO)6+ (6A) from C1 B20(CO)5+ (5A) is also a barrierless process (Fig. 2A and Fig. S8). The subsequent chemisorption pathways C2 B20(CO)6 (6A) + CO → C1 B20(CO)7+ (7A) and C1 B20(CO)7+ (7A) + CO → D2 B20(CO)8+ (8A) involve the transition states TS2 and TS3, which lie 0.04 eV and 0.03 eV lower than the entrance channels (Fig. 2A and Fig. S11, S12), respectively, indicating that these processes are also barrierless, with the axially chiral D2 B20(CO)8+ (8A) possessing the highest symmetry in the series with a symmetrical CO ligand distribution. However, as shown in Fig. S13, further calculations indicate that the chemisorption process of D2 B20(CO)8+ (8A) + CO → C1 B20(CO)9+ (9A) possesses a positive energy barrier of +0.15 eV, indicating that the formation of B20(CO)9+ is kinetically unfavourable under ambient conditions, consistent with the observation that no mass signal of B20(CO)9+ was detected in experiments (Fig. 1(a7)). As the most favourable chemisorption pathway, the D2d B20+ → Cs 1A → C2 2A → C1 3A → Cs 4A → C1 5A → C2 6A → C1 7A → D2 8A process in Fig. 2A possesses an overall exothermicity of 9.22 eV, enabling the consecutive chemisorptions of eight CO ligands in the B20(CO)n+ series (n = 1–8). As indicated in Fig. S7, the PBE0 and PBE0-D3 relative energies and energy barriers on the most favourable chemisorption pathway of the experimentally observed B20(CO)n+ monocations (n = 0–8) are well supported by single-point DLPNO-CCSD(T) calculations, the most accurate calculations performed in this work. Inspiringly, as clearly indicated in Fig. 2(B), the chemisorption energies (Ec) with respect to B20+ + nCO → B20(CO)n+ exhibit an almost perfect linear relationship of Ec = 1.17n + 0.03 with the number (n) of CO ligands involved in the complexes, with the average chemisorption energy of Ec = 1.17 eV per CO, indicating that DR tubular B20+ monocation consecutively chemisorbs the eight CO molecular ligands almost independently.
 |
| Fig. 2 Optimized structures and chemisorption pathways of (A) B20(CO)n+ monocations (n = 1–8) and (C) B20(CO)n neutrals (n = 1–8), with their chemisorption energies and corresponding energy barriers indicated in eV at the PBE0/6-311+G(d) and PBE0-D3/6-311+G(d) (in square brackets) levels, respectively. Variations in the calculated chemisorption energies (Ec) with the numbers (n) of CO ligands in the concerned complexes with respect to (B) B20+ + nCO → B20(CO)n+ and (D) B20 + nCO → B20(CO)n (n = 1–8) at PBE0/6-311+G(d). | |
 |
| Fig. 3 AdNDP bonding patterns of the DR tubular D10d B20 (A), D2d B20+ (B), Cs B20(CO)4+ (C), and D2 B20(CO)8+ (D) on the B20 DR tubular frameworks, with the occupation numbers (ON) indicated. | |
Further PBE0 calculations indicate that, starting from neutral DR tubular D10d B20, which possesses twenty equivalent peripheral boron atoms, the neutral Cs B20(CO) (1a) can be formed by overcoming a marginal energy barrier of +0.02 eV, while C2 B20(CO)2 (2a) and C1 B20(CO)4 (4a) can be spontaneously generated via barrierless processes (Fig. 2C and Fig. S14, S15, S17). The sequential B20(CO)n can be achieved by overcoming the small energy barriers of +0.01, +0.04, +0.23, +0.04, and +0.05 eV for n = 3, 5, 6, 7, and 8, respectively (Fig. 2C and Fig. S16, S18–S21), with an overall exothermicity of 7.22 eV. The carbonylation of the typical aromatic neutral DR tubular D10d B20 turns out to be both thermodynamically and dynamically less favourable than that of the more electron-deficient DR tubular monocation D2d B20+. Interestingly, as shown in Fig. 2(D), the chemisorption energies with respect to B20 + nCO → B20(CO)n also exhibit an almost perfect linear relationship Ec = 0.90n + 0.11 with the number of CO ligands involved in the systems, with an obviously lower average chemisorption energy of Ec = 0.90 eV per CO. Interestingly, as indicated in Fig. 2, with D3 dispersion corrections included, the PBE0-D3 approach produces similar optimized structures and relative energies with PBE0 for both the B20(CO)n+ and B20(CO)n (n = 1–8) series, with PBE0-D3 generating small negative energy barriers for all the concerned chemisorption processes, except B20(CO)5 (5a) + CO → B20(CO)6 (6a), which has a small positive energy barrier of +0.13 eV.
Bonding pattern and aromaticity analyses
To better understand the high stability of the B20(CO)n+/0 (n = 0–8) series, detailed AdNDP bonding pattern analyses were performed, as shown in Fig. 3 and Fig. S22–S29. Fig. 3 clearly indicates that the perfect DR tubular neutral D10d B20 and experimentally observed DR tubular D2d B20+ possess very similar bonding patterns, which can be categorized into three subgroups. For simplicity, we first discuss the closed-shell neutral D10d B20 in Fig. 3A. The first subgroup in B20 contains 20 2c-2e localized B–B σ-bonds on the two B10 rings with the occupation numbers (ON) of ON = 1.79 |e|. These bonds can also be analysed as 20 3c-2e σ-bonds with slightly higher ON values (1.97 |e|), which are responsible for connecting the two adjacent B10 rings, similar to the situation reported in DR tubular Co@B16−.51 The second group consists of 5 20c-2e σ + σ bonds with ON = 2.00 |e|, which represent in-phase orbital overlaps between the two adjacent B10 rings. These completely delocalized σ-bonds match the 4n + 2 Hückel rule (n = 2) and render σ-aromaticity to DR tubular D10d B20. The third group possesses 5 completely delocalized 20c-2e π + π bonds formed positively between the two B10 rings with ON = 2.00 |e|, which match the 4n + 2 Hückel rule (n = 2) and make the system π-aromatic in nature. As clearly shown in Fig. 3B, the slightly distorted open-shell D2d B20+ possesses a bonding pattern very similar to that of neutral D10d B20, with the only difference occurring at the last delocalized 20c-2e π + π bond, which turns out to be singly occupied with ON = 1.00 |e|. For the corresponding boron carbonyl complex monocations, as demonstrated, Fig. 3C and D show the bonding patterns of Cs B20(CO)4+ and D2 B20(CO)8+ on their DR tubular B20 frameworks, respectively. Obviously, the 20 2c-2e localized σ-bonds formed on the two B10 rings and 5 20c-2e σ + σ delocalized σ-bonds and 5 20c-2e π + π delocalized π-bonds formed positively over the DR tubular B20 framework in D2d B20+ (Fig. 3B) have all been practically well inherited in both Cs B20(CO)4+ and D2 B20(CO)8+, though with slightly lower ON values. Similar bonding patterns exist in the whole B20(CO)n+ and B20(CO)n species (n = 1–8), as depicted in Fig. S21–S28.
To comprehend the overall aromaticity of these DR tubular systems, Fig. 4 depicts the calculated ICSS surfaces of B20(CO)n+ and B20(CO)n (n = 2, 4, 6, 8) series, compared with that of the perfect DR D10d B20 on the top, which is known to possess typical tubular aromaticity.20,52,53 Interestingly, both the B20(CO)n+ and B20(CO)n series exhibit similar ICSS surfaces with that of D10d B20, where the yellow areas with negative NICS-ZZ values inside the B20 tube and within about 1.0 Å above the tube in the vertical direction belong to the chemical shielding regions, while the green regions with positive NICS-ZZ values like a belt around the B20 tube in horizontal directions belong to the chemical de-shielding regions. More specifically, B20(CO)n+ and B20(CO)n (n = 2, 4, 6, 8) possess the negative nuclear independent chemical shift values of NICS = −19.75, −14.53, −7.46, and −5.36 and NICS = −26.06, −16.82, −9.57, and −4.94 ppm at the centres of the DR B20 tubes, respectively. Similar ICSSs exist in other B20(CO)n+ and B20(CO)n (n = 1, 3, 5, 7) series (Fig. S30). These results indicate that, similar to DR tubular D10d B20, which has the largest negative NICS value of −39.71 ppm at the centre, both the distorted DR tubular B20(CO)n neutrals and B20(CO)n+ monocations (n = 1–8) exhibit tubular aromaticity, rendering extra-stability to help stabilize these DR tubular complexes. As indicated in Fig. S31, the calculated ring current maps of the slightly or severely distorted DR tubular B20(CO)n+ and B20(CO)n (n = 2, 4, 6, 8) using the ACID approach44 further evidence the tubular aromaticity of the systems.
 |
| Fig. 4 Calculated ICSS surfaces of D10d B20, C2 B20(CO)2+, Cs B20(CO)4+, C2 B20(CO)6+, D2 B20(CO)8+, C2 B20(CO)2, C2 B20(CO)4, C1 B20(CO)6, and C1 B20(CO)8 at the PBE0 level, with the calculated nuclear independent chemical shift values at the centres of the DR tubular B20 tubes (NICS(0)) indicated in ppm. The yellow and green areas with negative and positive NICSzz values represent the shielding and de-shielding regions, respectively. | |
Effective σ-donations and weak π-back-donations
Detailed EDA-NOCV analyses in Fig. 5 and Tables S1, S2 reveal the corresponding deformation densities Δρ and shapes of the most important interacting orbitals of the pairwise orbital interactions depicted of Cs B20(CO)+ and Cs B20(CO), more specifically, with D2d B20+ and CO and D10d B20 and CO as reacting fragments, respectively. Obviously, one effective σ-donation from the HOMO of CO to the LUMO of B20+ in radial direction, which contributes 65.37% to the overall orbital interaction (ΔEorb = −105.2 kcal mol−1) between B20+ and CO and two weak π-back-donations perpendicular to each other from the HOMO and HOMO−9 of B20+ to the two degenerate LUMOs of CO, which contribute 15.86% and 7.97% (Fig. 5A), respectively, coexist in the coordination interactions of Cs B20(CO)+, with the dominant σ-donation bond being well reflected in the AdNDP 2c-2e B–C σ-bond in Fig. S21. As shown in the red → blue charge flow colour code, the doubly occupied σ-HOMO of CO serves as a lone-pair donor in the effective σ-donation interaction, while the two unoccupied degenerate antibonding π*-LUMOs of CO perpendicular to each other function as electron acceptors in the two weak π-back-donations. Similarly, one effective σ-donation, which contributes 58.76% to the overall orbital interactions and two weak π-back-donations, which contribute 21.07% and 8.71%, respectively, coexist in Cs B20(CO), as clearly shown in Fig. 5B.
 |
| Fig. 5 Plots of the deformation densities (Δρ) and shapes of the most important interacting orbitals of the pairwise orbital interactions between B20+ and CO in Cs B20CO+ (A) and B20 and CO in Cs B20CO (B), with the orbital interaction energies (ΔEorb) in kcal mol−1 and their percentage contributions to the overall orbital interactions indicated. The colour code of the charge flow is from red to blue. | |
Similar coordination bonding patterns exist in the whole B20(CO)n+/0 series between the CO ligands and the DR tubular B20 core. The effective B20 ← CO σ-donations and weak B20 → CO π-back-donations in these boron carbonyl complexes appear to be similar to that of the TM–CO coordination interactions in classic TM carbonyl complexes, indicating that the peripheral boron atoms in these DR tubular boron carbonyl complexes can also be viewed as “honorable transition metals”,4,5 similar to the situations observed in the previously reported 2D B13(CO)n+, B11(CO)n, and B15(CO)n+.17,18
Conclusions
Extensive chemisorption experiments and first-principles theory calculations performed in this work indicate that previously experimentally observed gas-phase DR tubular D2d B20+ can chemisorb up to eight CO molecules consecutively under ambient conditions to form a series of DR tubular boron carbonyl monocations B20(CO)n+ (n = 1–8), unveiling the emergence of a 2D-to-3D transition in boron carbonyl complexes. Both DR tubular B20(CO)n+ and their neutral counterparts B20(CO)n appear to be tubularly aromatic in nature. Further joint experimental and theoretical investigations on larger boron carbonyl complexes Bm(CO)n+/−, which may have a 3D cage-like, bilayer and core–shell structures, are currently in progress. Carbonyl coordination is expected to be an effective approach to stabilize boron nanoclusters and low-dimensional nanomaterials to further enrich the chemistry of boron.
Conflicts of interest
There are no conflicts to declare.
Data availability
All the data are available online on the website of PCCP.
All the necessary low-lying isomers and their chemisorption pathways, bonding patterns, ICSS surfaces, ring currentmaps, and EDA-NOCV analyses provided. See DOI: https://doi.org/10.1039/d5cp01827g
Acknowledgements
This work was supported by the National Natural Science Foundation of China (Grant No. 22003034, 22373061, and 92461303).
References
- M. F. Zhou, L. Andrews and C. W. Bauschlicher, Jr., Chem. Rev., 2001, 101, 1931 CrossRef CAS
. - Q. Q. Tian, X. Yin, R. J. Sun, X. F. Wu and Y. H. Li, Coord. Chem. Rev., 2023, 475, 214900 CrossRef CAS
. - A. M. Ricks, Z. E. Reed and M. A. Duncan, J. Mol. Spectrosc., 2011, 266, 63 CrossRef CAS
. - X. Wu, L. Zhao, J. Jin, S. Pan, W. Li, X. Jin, G. Wang, M. Zhou and G. Frenking, Science, 2018, 361, 912 CrossRef CAS
. - S. K. Purkayastha, S. S. Rohman, P. Parameswaran and A. K. Guha, Phys. Chem. Chem. Phys., 2024, 26, 12573 RSC
. - A. B. Burg and H. I. Schlesinger, J. Am. Chem. Soc., 1937, 59, 780 CrossRef CAS
. - M. Zhou, N. Tsumori, L. Andrews and Q. A. Xu, J. Phys. Chem. A, 2003, 107, 2458 CrossRef CAS
. - M. Zhou, N. Tsumori, Z. Li, K. Fan, L. Andrews and Q. Xu, J. Am. Chem. Soc., 2002, 124, 12936 CrossRef CAS
. - M. Zhou, Z. X. Wang, P. V. R. Schleyer and Q. Xu, ChemPhysChem, 2003, 4, 763 CrossRef CAS PubMed
. - T. R. Burkholder and L. Andrews, J. Phys. Chem., 1992, 96, 10195 CrossRef CAS
. - M. Zhou, Q. Xu, Z. X. Wang and P. V. R. Schleyer, J. Am. Chem. Soc., 2002, 124, 14854 CrossRef CAS PubMed
. - J. Y. Jin, G. J. Wang and M. F. Zhou, Chin. J. Chem. Phys., 2016, 29, 47 CrossRef CAS
. - J. Jin, G. Wang and M. Zhou, J. Phys. Chem. A, 2018, 122, 2688 CrossRef CAS
. - J. Jin, G. Wang, M. Zhou, D. M. Andrada, M. Hermann and G. Frenking, Angew. Chem., Int. Ed., 2016, 55, 2078 CrossRef CAS PubMed
. - J. Jin and M. Zhou, Dalton Trans., 2018, 47, 17192 RSC
. - Y. Zhao, T. Wang, C. Wang, Z. Zhang, H. Zheng, S. Jiang, W. Yan, H. Xie, G. Li, J. Yang, G. Wu, W. Zhang, D. Dai, X. Zheng, H. Fan, L. Jiang, X. Yang and M. Zhou, ChemPhysChem, 2022, 23, e202200060 CrossRef CAS
. - R. N. Yuan, J. J. Chen, Q. Chen, Q. W. Zhang, H. Niu, R. Wei, Z. H. Wei, X. N. Li and S. D. Li, J. Am. Chem. Soc., 2024, 146, 31464 CrossRef CAS PubMed
. - R. N. Yuan, Q. Chen, H. Niu, C. Y. Gao, X. N. Zhao, Y. B. Wu, S. G. He and S. D. Li, Phys. Chem. Chem. Phys., 2025, 27, 7279 RSC
. - J. Wang, W. Fan, S. B. Cheng and J. Chen, J. Phys. Chem. A, 2024, 128, 7869 CrossRef CAS
. - B. Kiran, S. Bulusu, H. J. Zhai, S. Yoo, X. C. Zeng and L. S. Wang, Proc. Natl. Acad. Sci. U. S. A., 2005, 102, 961 CrossRef CAS
. - W. An, S. Bulusu, Y. Gao and X. C. Zeng, J. Chem. Phys., 2006, 124, 154310 CrossRef
. - T. B. Tai, N. M. Tam and M. T. Nguyen, Chem. Phys. Lett., 2012, 530, 71 CrossRef CAS
. - E. Oger, N. R. M. Crawford, R. Kelting and P. Weis, Angew. Chem., Int. Ed., 2007, 46, 8503 CrossRef CAS
. - X. N. Wu, B. Xu, J. H. Meng and S. G. He, Int. J. Mass Spectrom., 2012, 310, 57 CrossRef CAS
. - Z. Yuan, Y. X. Zhao, X. N. Li and S. G. He, Int. J. Mass Spectrom., 2013, 354, 105 CrossRef
. - R. A. J. O’Hair, Chem. Commun., 2006, 1469 RSC
. - L. D. Socaciu, J. Hagen, U. Heiz, T. M. Bernhardt, T. Leisner and L. Wöste, Chem. Phys. Lett., 2001, 340, 282 CrossRef CAS
. - 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 CrossRef CAS
. - C. Adamo and V. Barone, J. Chem. Phys., 1999, 110, 6158 CrossRef CAS
. - R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople, J. Chem. Phys., 1980, 72, 650 CrossRef CAS
. - 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, et al., Gaussian 16, Revision C.01, Gaussian, Inc., Wallingford, CT, USA, 2016 Search PubMed
. - Q. Chen, W. L. Li, Y. F. Zhao, S. Y. Zhang, H. S. Hu, H. Bai, H. R. Li, W. J. Tian, H. G. Lu, H. J. Zhai, S. D. Li, J. Li and L. S. Wang, ACS Nano, 2015, 9, 754 CrossRef CAS
. - H. J. Zhai, Y. F. Zhao, W. L. Li, Q. Chen, H. Bai, H. S. Hu, Z. A. Piazza, W. J. Tian, H. G. Lu, Y. B. Wu, Y. W. Mu, G. F. Wei, Z. P. Liu, J. Li, S. D. Li and L. S. Wang, Nat. Chem., 2014, 6, 727 CrossRef CAS
. - J. Tian, X. N. Chen, S. D. Li, A. I. Boldyrev, J. Li and L. S. Wang, Chem. Soc. Rev., 2019, 48, 3550 RSC
. - C. Gonzalez and H. B. Schlegel, J. Chem. Phys., 1989, 90, 2154 CrossRef CAS
. - C. Gonzalez and H. B. Schlegel, J. Phys. Chem., 1990, 94, 5523 CrossRef CAS
. - S. Grimme, S. Ehrlich and L. Goerigk, J. Comput. Chem., 2011, 32, 1456 CrossRef CAS PubMed
. - Y. Guo, C. Riplinger, U. Becker, D. G. Liakos, Y. Minenkov, L. Cavallo and F. Neese, J. Chem. Phys., 2018, 148, 011101 CrossRef
. - F. Neese, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2022, 12, e1606 Search PubMed
. - N. V. Tkachenko and A. I. Boldyrev, Phys. Chem. Chem. Phys., 2019, 21, 9590 RSC
. - D. Y. Zubarev and A. I. Boldyrev, Phys. Chem. Chem. Phys., 2008, 10, 5207 RSC
. - T. Lu and F. Chen, J. Comput. Chem., 2012, 33, 580 CrossRef CAS PubMed
. - W. Humphrey, A. Dalke and K. Schulten, J. Mol. Graph., 1996, 14, 33 CrossRef CAS PubMed
. - D. Geuenich, K. Hess, F. Köhler and R. Herges, Chem. Rev., 2005, 105, 3758–3772 CrossRef CAS
. - Povray, Persistence of vision raytracer, POV-Ray 3.7, https://www.povray.org/.
- T. Ziegler and A. Rauk, Theor. Chim. Acta, 1977, 46, 1 CrossRef CAS
. - M. P. Mitoraj, A. Michalak and T. Ziegler, J. Chem. Theory Comput., 2009, 5, 962 CrossRef CAS
. - M. Mitoraj and A. Michalak, Organometallics, 2007, 26, 6576 CrossRef CAS
. - G. T. E. Velde, F. M. Bickelhaupt, E. J. Baerends, C. F. Guerra, S. J. A. Van Gisbergen, J. G. Snijders and T. Ziegler, J. Comput. Chem., 2001, 22, 931 CrossRef
. - E. VanLenthe, E. J. Baerends and J. G. Snijders, J. Chem. Phys., 1993, 99, 4597 CrossRef CAS
. - I. A. Popov, T. Jian, G. V. Lopez, A. I. Boldyrev and L. S. Wang, Nat. Commun., 2015, 6, 8654 CrossRef CAS PubMed
. - H. J. Zhai, B. Kiran, J. Li and L. S. Wang, Nat. Mater., 2003, 2, 827 CrossRef CAS PubMed
. - M. P. Johansson, J. Phys. Chem. C, 2009, 113, 524 CrossRef CAS
.
|
This journal is © the Owner Societies 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.