Chinese Chemical Letters  2022, Vol. 33 Issue (2): 1058-1064   PDF    
A comparative study on the reactivity of cationic niobium clusters with nitrogen and oxygen
Benben Huanga,b, Mengzhou Yanga,b, Xin Leib, Wen Gana,b, Zhixun Luoa,*     
a Beijing National Laboratory for Molecular Sciences (BNLMS), State Key Laboratory for Structural Chemistry of Unstable and Stable Species, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China;
b University of Chinese Academy of Sciences, Beijing 100049, China
Abstract: We have prepared well-resolved Nbn+ (n = 1–10) clusters and report here an in-depth study on the essentially different reactivity with N2 and O2, by utilizing a multiple-ion laminar flow tube reactor in tandem with a customized triple quadrupole mass spectrometer (MIFT-TQMS). As results, the Nbn+ clusters are found to readily react with N2 and form adsorption products NbnN2m+; in contrast, the reactions with O2 give rise to NbnO1−4+Oproducts, and the odd-oxygen products indicate O-O bond dissociation, as well as increased mass abundance of NbO+ pertaining to oxygen-etching reactions. We illustrate how N2 prefers a physical adsorption on Nbn+ clusters with an end-on orientation for all the products, and allow for size-selective Nbn+ clusters to act as electron donor or acceptor in forming NbnN2m+. In contrast to these nitrides, the dioxides NbnO2+ display much larger binding energies, with O2 always as an electron acceptor, corresponding to superoxide or peroxide states in the initial reactions. Density-of-states and orbital analyses show that the interactions between Nbn+ and O2 are dominated by strong π-backdonation indicative of incidental electron transfer; whereas weak π-backdonation and simultaneous σ donation interactions exist in NbnN2+. Further, reaction dynamics analysis illustrates the different interactions for N2 and O2 in approaching the Nbn+ clusters, showing the energy diagrams for N2 adsorption and O-O bond dissociation in producing odd-oxygen products. Fragment analyses with orbital correlation and donor-acceptor charge transfer are also performed, giving rise to full insights into the reactivities and interactions of such transition metal clusters with typical diatomic molecules.
Keywords: Gas phase reaction    Nb cluster    Chemical adsorption    Orbital analysis    Energy decomposition analysis    

Gas adsorption is one of the most basic interfacial phenomena and closely related to diverse applications in industrial and agricultural production and daily life. People may take it for granted that the adsorption of gas molecules on the metals is simply divided into physical adsorption and chemical adsorption; however, the underlying mechanism that determines the diverse interactions and surface reactivity at reduced sizes are illusive to be fully understood. The study of gas adsorption on solid surfaces is helpful to understand the micro-mechanism of various physicochemical processes; also, adsorption of gas molecules on nanometals is a key step in catalytic processes and important for various gas phase reactions. While physisorption is often dominated by molecular van-der Waals force and generally emits rare energies, chemisorption can be largely exothermic and may involve covalent interactions and stepwise reactions [1]. In this regard, molecular adsorption can be associated with three processes, i.e., inactivated adsorption, activated adsorption and precursor-mediated adsorptions [2]. The adsorbate may reach a chemisorption state through a barrierless process directly from physisorption state; or it needs to overcome a potential barrier simply by heating or other reaction conditions; or undergoes a precursor-mediated adsorption, where the chemisorbed state could be attained through a precursor intermediate.

On this basis, the reactivities of transition metal clusters with small molecules have been extensively studied, illustrating the distinction of physical and chemical adsorptions on surfaces [2]. Also, the interaction mechanisms and properties of metal cluster complexes have been widely analyzed by joint experimental and theoretical studies. Among them, niobium clusters have been an important research object unveiling size-dependent chemical and physical properties of transition metals at reduced sizes [3-10]. Combining mass spectrometry and gas flow reactors [11, 12], the reactivities of niobium clusters with many chemicals have been studied, such as H2 [12-15], O2 [16, 17], N2 [18, 19], COx [20-22], NOx [23, 24] and hydrocarbons [25-29]. The reactions of Nb clusters with N2 and H2 (/D2) have similar adsorptive patterns; while the reaction with O2 shows a difference due to the differences of atomic valence configuration and bonding nature of these diatomic molecules. In general, N2 is inert in most reactions due to a very high dissociation energy (945 kcal/mol) of the N≡N triple bond and a high HOMO-LUMO gap (~10.82 eV); in comparison, ground-state O2 is in triplet with two spin single electrons pertaining to its high activity [30]. However, it is not fully unveiled how the stability of such metal clusters and atomic valence configuration and bonding nature cooperatively determine the experimental observation on their reactivity.

On the other hand, previously published studies have also estimated the reaction rate constants for niobium clusters with N2 and O2 showing a likely difference of three orders of magnitude [18, 19]. Further studies reported the similar adsorption of N2 on niobium clusters [18, 19, 31], and there are adsorption products of polynitrogen molecules depending on the pressure and concentration of the reaction gas at room temperature [32]. For example, Pillai et al. [33] generated Nb+(N2)n complexes (n = 3–16) with nitrogen as buffer gas and investigated the vibration in forming such complexes by infrared photodissociation spectroscopy (IRPD), showing a preferred coordination of six ligands. Meanwhile, the reactions of small Nbn+ (n = 1–3) with O2 have been studied by Loh et al. [16] who measured the cross section as a function of kinetic energy and proposed dissociation pathways and threshold values. It is anticipated that these experimental cluster physics studies of such small Nb clusters can be connected with theorical energy calculations in determining the thermodynamics and their reaction dynamics in gas phase [34].

On these bases, here we have performed a further in-depth study of the thermalized gas-phase Nbn+ (n = 1–10) clusters in reacting with sufficient N2 and O2 gas in a customized multiple ions laminar flow tube in tandem with triple quadrupole mass spectrometer (MIFT-TQMS). It is found that the Nbn+ clusters, produced by the home-made magnetron sputtering (MagS) source, readily react with both N2 and O2, but form different series of products respectively, seen as NbnN2m+ (m = 1–3) and NbnOx+ (x = 1–4). Theoretical calculations results reveal that the NbnN2m+ clusters all adopt end-on orientation adsorption for N2 on Nbn+ with much smaller binding energy than that of correlative NbnO2+ series which correspond to strong hollow-site adsorption allowing for occasional dissociative adsorption. Further, we conducted a comprehensive study thermodynamic energetics, HOMO-LUMO gaps, charge transfer, Wiberg bond index (WBI), density of state (DOS), natural atomic orbital (NAO) and energy decomposition analyses based on natural orbitals for chemical valence (EDA-NOCV). In addition, we carried out potential scan for N2 and O2 in approaching an Nbn+ cluster until the formation of Nb3N2+ and Nb3O2+, illustrating how cluster-molecule interactions initiate their reaction. Finally, we explore the reaction pathways of both N2 and O2 with Nb3+ as a representative, demonstrate the N2-adsorption behavior and the reaction dynamics in producing Nb3O2+ from O2-adsorption, to dissociative intermediate, and to ultimately exhaust of NbO (/NbO+).

The experiments in this study are conducted on our customized multi-ions laminar flow tube reactor in tandem with a triple quadrupole mass spectrometer (MIFT-TQMS). A self-designed MagS source with a DC power supply of 5 kW was used to obtain the clean mass distributions of small niobium clusters [35]. A niobium disk (99.95% purity, 50.8 mm diameter, 4 mm thickness) was used as the sputtering target. High-purity helium (> 99.999%) and high-purity argon (> 99.995%) were used as carrier gas and sputtering work gas, introduced from the rare of the MagS source and the magnetron head respectively, with the gas flow rates controlled by two mass flowmeters (Alicat, a range of 0–100 sccm and 0–100 slm, respectively). Typical parameters for the MagS source to produce the niobium clusters with n = 1–10 are: 170 V DC voltage, 4 A current, a background pressure of source chamber at ~7 Torr, He carrier gas at ~12 slm helium, while Argon is 60–120 sccm. Different concentrations of N2 and O2 (ca., 0.5%–30% in He) are injected into the flow tube (60 mm diameter, 1 m long) which is maintained at 0.9 Torr pressure for stable laminar flow and sufficient collisional reaction. The detailed values for these parameters may differ with dependence on the target situation and vacuum status, etc.

All the DFT calculations were performed using Gaussian 09 software package [36]. B3LYP exchange-correlation functional was used to optimize the geometric structures of all the niobium clusters and products, the basis sets of Lanl2dz was used for Nb with an extra d polarization function, while 6–311G(d) for N and O atoms [36, 37]. Vibrational frequency calculations were performed for each of the optimized structures and zero-point vibrations were corrected for all the energy calculations. Thermodynamic data, Wiberg bond index (WBI), partial density of states (PDOS), natural atomic orbital (NAO) and electron configuration are analyzed by using the Multiwfn software [38]. The geometric structures, frontier orbitals, charge distributions by natural population analysis (NPA), and the PDOS patterns of the Nbn+, NbnN2+, NbnO2+ (n = 1–10) are plotted by visual molecular dynamics (VMD) [39]. Energy decomposition analysis was conducted based on natural orbitals for chemical valence (EDA-NOCV), calculated with Amsterdam Density Functional (ADF) program [40, 41] at the B3LYP/TZP level of theory.

Fig. 1 presents the mass spectra of the Nbn+ (n = 1–21) clusters in the absence and presence of N2 and O2 gas reactants respectively, where the concentrations and gas flowrates are controlled by a mass flowmeter (ALICAT-LD12). As is shown in Fig. 1a, most of the Nbn+ clusters react with N2 just to form molecular adsorption products NbnN2m+ (m = 1–3) with size dependence, which is in contrast to the sequence reactions between Nbn+ and O2 in forming NbnOx+ (x = 1–4). From mass spectra in Fig. 1a and the variation of mass abundances of the nascent Nbn+ clusters and NbnN2m+ products (Fig. S1 in Supporting information), it is seen that Nb3−6+ clusters react faster than the other clusters and convert to NbnN2+products. In comparison, the Nb7+ is seen to prefer an adsorption of several nitrogen molecules to form Nb7N4, 6+ (also seen for Nb4+); The cation Nb+, also Nb8+ and Nb9+ slowly react with N2, but Nb2+and Nb10+display rare reaction products with N2 in the experimental condition of this study. The variation of Nbn+ clusters reacting with N2 reflects their size-dependent stability and reactivity, assuming they undergo identical collision annihilation with the wall of the flow tube. We have plotted the changes in relative intensity of Nbn+ (n = 3–9) and –ln(I/I0) of Nbn+ as a function of the gas flow rate (Fig. S3 in Supporting information). The experimental results display that the N2 addition on Nbn+ (n = 1–10) closely follow the quasi first-order reactions for which the rate constants have been estimated simply by using the equation ln(IA/I0) = -k1(PeΔt/kBT) [42], where kB is the Boltzmann constant, T is the temperature, Pe and Δt correspond to the effective pressure and reaction time. As results, the size-dependent reaction rates of Nbn+ (n = 1–10) exhibit a wide diversity, and Nb3+ is found to display a maximal value of ~10−9 cm3 molecule−1 s−1.

Download:
Fig. 1. The mass spectra of Nbn+ clusters in the absence and presence of (a) 30% N2/He and (b) 0.5% O2/He, respectively. The weak mass peaks marked with stars (*) in the nascent Nbn+ cluster distributions correspond to trace amount of oxygen contamination.

There is different case. As shown in Fig. 1b, the Nbn+ (n = 1–10) clusters are highly reactive with O2 even at a very low concentration (e.g., 0.8 sccm of 0.5% O2 in He gas). Even in the presence of a small amount of oxygen, the dissociative products NbnO+ are observed. This is reasonable considering the large Nb-O bond energy up to ~770 kJ/mol which is even larger than the O-O bond strength (~400 kJ/mol); and electron transfer readily occur between niobium and oxygen [43, 44], with a variety of exothermic reaction channels as given in Table 1. Among the thermodynamically favorable channels, the O2-dissociaed equation to produce a NbO neutral suggests the maximal energy release for all the Nbn+ (n = 2–10) clusters. This could well explain the experimental observation of dissociative products NbnO+ which was also found in previous experimental study [17]. It is worth noting that, such a favorable reaction channel "Nbn+ + O2 → Nbn-1O+ + NbO" largely differs from the previously determined reactions of "Aln+ + O2 → Alm++ Aln-mO2" [45], "Aln + O2 → Aln−4 + 2Al2O" [46] and "Con + O2 → Con−1 + CoO1" [47]. Due to the large Nb-O bond energy, here the reaction channels with Nb0/+or Nb20/+removal are much less favorable from the thermodynamics, which is well consistent with the experimental observation that Nb+ and Nb2+ do not display increased mass abundance in the varied reaction conditions. Also, this is consistent with the previous study by Radi et al. [17], where the dominant reaction pathway of mass-selected small niobium clusters with O2 in a drift reactor was proposed to be the one in producing primary products of NbO and NbO+.

Table 1
DFT-calculated energies (eV) for the likely diverse reaction pathways of Nbn+ (n = 1–10) with O2. All the data are calculated at the B3LYP/Lanl2dz level for Nbn+ and B3LYP/6-311G(d) level for N/O.

Utilizing DFT calculations we have provided further insights into the diverse reactivities of niobium clusters (for details of the lowest-energy structures, spin multiplicity, bond lengths, and point group see Fig. S6 in Supporting information). For the adsorption complexes, N2 adopts an end-on adsorption orientation on the Nbn+ clusters of which the nascent structures and spin multiplicities almost do not change, except Nb2N2+ which shows a V-shape structure with a high spin multiplicity. In contract, NbnO2+ (n > 2) clusters all adopt hollow-site adsorption mode with elongated O-O bond length ranging from 1.2 Å to 1.5 Å, although the structures of Nbn+ in NbnO2+do not show much changes. Moreover, the spin multiplicities of the odd-number small niobium clusters (i.e., Nb2n+1O2+) reduced in the formation of NbnO2+ (except Nb9O2+); whereas, for the even-number niobium clusters (i.e., Nb2n+), the spin of the cluster can be aligned opposite to that of the 3O2 molecule thus spin conservation in the Nb2nO2+products [48, 49].

Fig. 2 plots the DFT-calculated HOMO-LUMO gaps, binding energies (BE), and natural population analysis (NPA) of charge distributions of the Nbn+, NbnN2+ and NbO2+ (n = 1–10) clusters. More energetics data such as the VSE, ASE, binding energy per atom (BE) and Δ2E of the Nbn+ are given in Fig. S9 (Supporting information). While no surprise that the NbnN2+clusters show much smaller cluster-molecule binding energies and charge transfer comparing with that of the NbnO2+products, the HOMO-LUMO gaps of the nascent Nbn+clusters and the subsequent NbnN2+ and NbnO2+ are comparable with each other except a jump at n = 2, 7, 8 pertaining to altered spin multiplicity (Nb7O2+), symmetry (Nb2O2+) or structure changes (Nb8O2+). Note that, O2 always serves as an electron acceptor in interacting with the Nbn+ clusters, with a total electron transfer of 0.6–0.9 |e|, which is in sharp contrast to that of the NbnN2+ complexes of which the σ donations from N≡N to the Nbn+clusters are enabled especially for the larger ones at n ≥ 6. Detailed information about NPA charge and natural electron configuration of Nb3/6+, N2, O2, Nb3/6N2+ and Nb3/6O2+ is shown in Tables S3 and S4 (Supporting information).

Download:
Fig. 2. (a) DFT-calculated HOMO-LUMO gaps of Nbn+, NbnN2+ and NbO2+ (n = 1–10). (b) Binding energy of NbnN2+ and NbnO2+. (c) NPA charge transfer of Nbn+ to N2 in NbnN2+ compared with Nbn+ to O2 in NbnO2+. All the data are calculated at the B3LYP/Lanl2dz level for Nbn+ and B3LYP/6-311G(d) level for N/O.

Table 2 lists the Wiberg bond indexes (WBI) between the atoms in NbnN2+ and NbnO2+ (n = 1–10). As is shown, every WBI of the N(1)-N(2) bond in NbnN2+ changes to 2.69-2.90 from the primary N≡N triple bond in dinitrogen. Meanwhile, the WBI of the linked N(1)-Nb bond is about 0.3–0.5, while the WBI of N(2)-Nb is as small as 0.05–0.15 indicating weak interaction between the Nbn+ clusters and N2 corresponding to physical adsorption. By contrast, every WBI of O-O bond changes to about 1 from O=O double bond in the ground state oxygen molecule, and the bond length elongated from 1.2 Å to 1.4~1.5 Å), which is similar to O-O bond length and bonding mode in H2O2, as named peroxy bond (or peroxide state) after a chemical adsorption. Moreover, every O atom links to 2 Nb atoms in the 3Nb-2O elementary units when n ≥ 3, showing two types of O-Nb bonds with a minor different WBI, that is, O(1)-Nb(1, 2) and O(2)-Nb(1, 3) bond. It is notable that Nb2O2+ possesses 4 equal Nb-O bonds (WBI = 0.48) and the WBI of O-O bond is 1.02. In all, the WBI of every Nb-N and Nb-O provides detailed information of their bonding strengths.

Table 2
The WBI analysis between the atoms in NbnN2+ and NbnO2+ (n = 1–10).

We have conducted orbital analysis to reveal the essential interaction of Nbn+ with N2 and O2. Fig. 3 presents a comparison for the partial density of states (PDOS) and corresponding orbitals of 2 N(/O) atoms and linked 3 Nb atoms typically in Nb3N2+ and Nb3O2+ respectively. Firstly, the overlap area in PDOS of Nb3O2+ is much more than Nb3N2+. For Nb3N2+, this is displayed by the reconstituted Nb3N2+ orbital which possesses a pattern involving both the cluster and nitrogen molecule. The PDOS of non-dissociative adsorption of N2 gives rise to seen as σ2s, σ2s*, σ2pz which are associated with the dominant N2→Nb3+ σ donation interactions, and two equivalent orbitals π2px and π2py involving very few interactions, vacant π2px* and π2py* of nitrogen are associated with the Nb3+→N2 π backdonation interactions (detailed orbital information including the natural atomic orbital (NAO) contributions is seen in Table S5 in Supporting information) [50]. While in Nb3O2+, the interacting orbitals also possess character of free O2 molecule, where corresponding to σ2s, σ2s*, σ2pz of oxygen refer to O2→Nb3+ σ donation, and two equivalent orbital π2px and π2py both contribute to the O2→Nb3+ π donation interaction; meanwhile, the strong DOS correlative to π2px* and π2py* of oxygen accounts for the Nb3+→O2 dominant π backdonation (Table S5) [51]. Similarly, the NAO analyses for Nb6N2+ and Nb6O2+ are given in Table S6 (Supporting information).

Download:
Fig. 3. Energy level of Nb3N2+ and Nb3O2+ together with PDOS of 2 N atoms and 3 Nb atoms in Nb3N2+ and PDOS of 2 O atoms and 3 Nb atoms in Nb3O2+. Insets show the corresponding orbitals.

Further, we conducted energy decomposition analysis based on natural orbitals for chemical valence (EDA-NOCV) [52-55]. The typical results of Nb3N2+ and Nb3O2+ are shown in Table 3, where the strength of the intrinsic interaction energies (ΔEint) are divided into three components (ΔEint = ΔEpauli + ΔEelstat + ΔEorb) corresponding to pauli exclusion (ΔEpauli), electrostatic interaction (ΔEelstat) and orbital interaction (ΔEorb) which can be further divided into σ donation, π donation, π backdonation, polarization and the rest interactions. For Nb3N2+, ΔEelstat and ΔEorb each contributes a half of the attractive interactions; while in Nb3O2+, the ΔEorb contributes about 60%. In total orbital interactions, the N2→Nb3+ σ donation interaction and polarization interaction contribute ~48%; Nb3+→N2 π backdonation contributes ~40%. In contrast for Nb3O2+, ΔEorb is much larger than ΔEelstat in total interactions, among which and the two O2→Nb3+π donation interactions contribute only ~13%, reversely, Nb3+→O2 π backdonation contributes ~83% of the interactions. It is worth mentioning that the interacting fragments 3Nb3+ and 3O2 are both in spin triplet, but 1Nb3O2+ in singlet state, which results in a larger value of ΔEint than the BE in forming Nb3O2+ because of the differences between the unactivated fragments and the activated complex (which involves an additional energy to activate the precursor).

Table 3
EDA results for 3Nb3N2+ and 1Nb3O2+ at the B3LYP/TZP level of theory using ADF, taking Nb3+ and N2 (/O2) as interacting fragments. ΔEint is the intrinsic interaction energies. ΔEpauli is the repulsion energy caused by the Pauli exclusion principle. ΔEelstst and ΔEorb are the attraction energies due to electrostatic and orbital interactions, respectively. The values in parentheses show the contribution to ΔEint and the total orbital interaction ΔEorb.

Fig. 4 displays the principal orbitals correlation diagrams of Nb3N2+ and Nb3O2+ based on the corresponding fragments, with only α-orbitals being shown as the correlation of β-orbitals is similar. In Nb3N2+ (Fig. 4a), N2→Nb3+ σ donation interaction dominantly forms compound orbital HOMO-8, which is mainly contributed (about 83%) by HOMO of N2 (σ2pz) and vacant LUMO+1 orbital of Nb3+; and Nb3+→N2 π backdonation forms new compound orbitals (HOMO, HOMO-1, HOMO-6) which are composed by the occupied orbitals of Nb3+ and LUMO of N2 (π2px/y*). In Nb3O2+ as a comparison (Fig. 4b), two O2→Nb3+ π donation interactions form HOMO-8, HOMO-9 orbitals of the complex, which are mainly from the occupied antibonding HOMO (π2px*) and one of the degenerate orbitals (π2py) of O2, coupled with vacant orbitals (LUMO, LUMO+1) of Nb3+. Here π backdonation interactions sourcing from bonding orbitals of Nb3+ and LUMO (π2py*) of O2 (Nb3+→O2) play a dominant contribution and form HOMO-7 orbital of the Nb3O2+ complex. The electrons on O2 fragment π2py* orbital (Fig. S14 in Supporting information) stem from charge transfer and orbital interactions [56, 57]. In addition, NOCV analysis including detailed values of interaction energies and the shapes of electron deformation densities (Δρ) are shown in Figs. S17 and S18 (Supporting information), which display the weak interaction/bonding between Nb3+ and N2, while much larger donor-acceptor interaction energies for Nb3O2+ in the single state.

Download:
Fig. 4. Kohn–Sham orbital correlation diagrams of (a) Nb3N2+ in the triplet state. (b) Nb3O2+ in the stable triplet state. Blue lines are for orbitals composited by one part, and red lines are for orbitals interaction of two fragments with percentage which solid lines are for occupied orbitals and dotted line are for vacant orbitals. All the data are calculated at the B3LYP/TZP level of theory using ADF.

Figs. 5a and b present the potential scan curves of Nb3+ in reacting with N2 and O2 respectively to form the products (more details of varying directions are given in Figs. S20 and S21 in Supporting information), where the NPA charge distribution is displayed in color with dependence on the varying distances of the reactants. For "Nb3+···N2" (Fig. 5a), the two overlap curves display minor vibration of the N-N bond distance ranging from 1.09 Å to 1.11 Å. Considering that the product 1Nb3O2+has an altered spin multiplicity and oxygen state comparing with the nascent 3Nb3+ and 3O2, we have performed the potential scan from two directions, allowing the geometrical configuration and spin multiplicity to be equal to the reactants and the products respectively. As is shown in Fig. 5b, the two scanning curves for "Nb3+···O2" display a crossing at ~2.4 Å, which indicates a critical value for the interaction distance of chemical reaction, and the energy minima point at a Nb3+···O2 distance of 1.74 Å corresponds to the oxygen-activated state (i.e., 1Δg state, with O-O bond distance at 1.44 Å). Furthermore, Kohn–Sham orbital correlation diagrams, charge transfer diagrams along with orbital interactions and detailed NOCV analysis of Nb3O2+ in imaginary quintuplet state at a Nb3+···O2 distance of 1.80 Å and O-O bond distance of 1.20 Å are shown in Figs. S13, S15 and S19 in Supporting information).

Download:
Fig. 5. Potential scan curves of the Nb3+ cluster in approaching (a) N2 and (b) O2 from two directions with the geometrical configuration and spin multiplicity equal to the reactants and the products respectively. Red lines refer to real approaching process due to energy crossing point, while light blue/green lines refer to imaginary scanning processes. The insets show the NPA charge of every atom as a function of different interaction distance.

Having unveiled the different interactions involved in "Nb3+ + N2" and "Nb3+ + O2", we then plotted the reaction coordinates for the reaction dynamics of the niobium clusters with oxygen and nitrogen, as shown in Fig. 6. Two reaction pathways are considered for Nb3+ with N2 in view of the likely varying molecular orientation of dinitrogen in approaching the cluster. As is shown, when the dinitrogen takes an end-on adsorption on a vertex atom of the Nb3+ cluster (black line), it gets a 0.62 eV energy gain but subsequent path toward the final dissociative Nb3N2+ needs to overcome a few endothermic steps and the rate-determine transition state TS3 is 0.27 eV higher than the relative energy of the reactants. Alternatively, if the dinitrogen molecule could approach the cluster via an orientation orthogonal to one of the triangular edges of the Nb3+, the reaction path (green line) needs to overcome a relatively smaller energy barrier but the rate-determine TS1 is still 0.24 eV higher in energy than the reactants. This energy barrier is small especially when considering the N≡N bond dissociation energy of ~9.79 eV, but it could not proceed spontaneously in the thermalized flow tube reactors (where previous studies suggest a threshold value could be ~0.1 eV by taking into consideration of vibrational, rotational and translational kinetic energy) [56, 58].

Download:
Fig. 6. The calculated reaction pathways for (a) "Nb3+ + N2" and (b) "Nb3+ + O2". All the thermodynamic data are calculated at the B3LYP/Lanl2dz for Nb and B3LYP/6-311G (d) for O and N.

In sharp contrast, the reaction of "Nb3+ + O2" is largely exothermic (8.10 eV in our calculation) to form an oxygen-dissociative products with two bridge oxygen atoms bonding to two edges of the triangular Nb3+ cluster, and the spin multiplicity turns to be singlet indicative of spin accommodation in the reaction process of 3Nb3+ with 3O2 [48, 59]. Note that, this reaction to form oxygen-dissociated product is highly thermodynamic and also kinetics-favorable, with the TS1 being 2.33 eV (or 3.2 eV for the yellow line) lower than the total energy of the reactants. Therefore, the mass spectrometry observation of Nb3O2+ (and the other NbnO2+) could be completely dissociated products. This is largely due to the large Nb-O bond energy (~770 kJ/mol) in contrast to the bond strengths of O-O (~400 kJ/mol) and Nb-Nb (~513 kJ/mol). The large amount of energy release by dissociative adsorption makes it feasible to produce odd-oxygen produts (e.g., Nb2O+/0, NbnO1, 3+) along with the NbO0/+ removal.

In summary, utilizing our customized instrument MIFT-TQMS we report here an in-depth study to compare the gas-phase reactivity of Nbn+ clusters with N2 and O2. Under sufficient backing gas pressure (up to 1 Torr), the Nbn+ clusters readily react with N2 and form a series of adsorption products NbnN2m+. In comparison, the Nbn+ clusters react with O2 in a comparable rate but produce NbnO1−4+along with a rapidly growing product NbO+, showing the O-O bond dissociation and metal cluster etching reactions akin to the previous findings on anionic aluminum and cobalt clusters. Energetics calculations show that the adsorption energies of NbnO2+ (2.5~4 eV) are significantly larger than that of NbnN2+ (0.5~0.9 eV). While the dinitrogen all follow an end-on orientation adsorption on the Nbn+ clusters, the chemisorptive NbnO2+ species conform to hollow-site adsorption along with spin-flip and perfect accommodation to be singlet. PDOS and frontier orbital analyses reveal strong π backdonation from Nbn+ to O2; which is in contrast to the NbnN2+ system which involves both σ donation from N2 to Nbn+ and simultaneously weak π backdonation from Nbn+ to N2. Potential scan plots a van der Waals curve for the N2 molecule in approaching the Nb3+ cluster, whereas the potential scan for "Nb3+···O2" displays a crossing from two directions with the spin multiplicity equal to the reactants and the products respectively, shedding light on the spin crossing effect. Further, reaction dynamics calculation unveils feasible O-O bond dissociation and finally NbO (/NbO+) removal from such NbnO2m+ clusters.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgments

This work was financially supported by the CAS Instrument Development Project (No. Y5294512C1), the National Natural Science Foundation of China (No. 21722308) and Key Research Program of Frontier Sciences (CAS No. QYZDBSSWSLH024).

Supplementary materials

Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2021.04.020.

References
[1]
M. Schmidt, A. Masson, H.P. Cheng, C. Bréchignac, ChemPhysChem 16 (2015) 855-865. DOI:10.1002/cphc.201402726
[2]
M.B. Knickelbein, Annu. Rev. Phys. Chem. 50 (1999) 79-115. DOI:10.1146/annurev.physchem.50.1.79
[3]
S.K. Loh, L. Lian, P.B. Armentrout, J. Am. Chem. Soc. 111 (1989) 3167-3176. DOI:10.1021/ja00191a010
[4]
M.B. Knickelbein, S. Yang, J. Chem. Phys. 93 (1990) 5760-5767. DOI:10.1063/1.459570
[5]
W.J.C. Menezes, M.B. Knickelbein, J. Chem. Phys. 98 (1993) 1856-1866. DOI:10.1063/1.464220
[6]
V. Kumar, Y. Kawazoe, Phys. Rev. B 65 (2002) 125403. DOI:10.1103/PhysRevB.65.125403
[7]
W. Fa, C. Luo, J. Dong, Phys. Rev. B 71 (2005) 245415. DOI:10.1103/PhysRevB.71.245415
[8]
P.V. Nhat, V.T. Ngan, M.T. Nguyen, J. Phys. Chem. C 114 (2010) 13210-13218. DOI:10.1021/jp103484k
[9]
N.P. Vu, N.V. Thi, T.T. Ba, N.M. Tho, J. Phys. Chem. A 115 (2011) 3523-3535. DOI:10.1021/jp110758p
[10]
C. Berkdemir, S.-B. Cheng, A.W. Castleman, Int. J. Mass. Spectrom. 365-366 (2014) 222-224. DOI:10.1016/j.ijms.2014.03.011
[11]
C.Q. Jiao, B.S. Freiser, J. Phys. Chem. 99 (1995) 10723-10730. DOI:10.1021/j100027a008
[12]
J.L. Elkind, F.D. Weiss, J.M. Alford, R.T. Laaksonen, R.E. Smalley, J. Chem. Phys. 88 (1988) 5215-5224. DOI:10.1063/1.454596
[13]
M.E. Geusic, M.D. Morse, R.E. Smalley, J. Chem. Phys. 82 (1985) 590-591. DOI:10.1063/1.448732
[14]
A.B. Vakhtin, K. Sugawara, Chem. Phys. Lett. 299 (1999) 553-560. DOI:10.1016/S0009-2614(98)01320-7
[15]
A.B. Vakhtin, K. Sugawara, J. Chem. Phys. 115 (2001) 3629-3639. DOI:10.1063/1.1389302
[16]
S.K. Loh, L. Lian, P.B. Armentrout, J. Chem. Phys. 91 (1989) 6148-6163. DOI:10.1063/1.457434
[17]
P.P. Radi, G. von Helden, M.T. Hsu, P.R. Kemper, M.T. Bowers, Int. J. Mass. Spectrom. 109 (1991) 49-73. DOI:10.1016/0168-1176(91)85096-5
[18]
M.D. Morse, M.E. Geusic, J.R. Heath, R.E. Smalley, J. Chem. Phys. 83 (1985) 2293-2304. DOI:10.1063/1.449321
[19]
A. Bérces, P.A. Hackett, L. Lian, S.A. Mitchell, D.M. Rayner, J. Chem. Phys. 108 (1998) 5476-5490. DOI:10.1063/1.475936
[20]
L. Song, A. Eychmueller, R.J. St. Pierre, M.A. El-Sayed, J. Chem. Phys. 93 (1989) 2485-2490. DOI:10.1021/j100343a050
[21]
L. Holmgren, M. Andersson, A. Rosén, Surf. Sci. 331-333 (1995) 231-236. DOI:10.1016/0039-6028(95)00099-2
[22]
H. Gronbeck, A. Rosen, Phys. Rev. B 54 (1996) 1549-1552. DOI:10.1103/PhysRevB.54.1549
[23]
Q.F. Wu, S.H. Yang, Int. J. Mass. Spectrom. 184 (1999) 57-65. DOI:10.1016/S1387-3806(98)14252-5
[24]
D.J. Harding, T.A. Oliver, T.R. Walsh, et al., Eur. J. Mass. Spectrom. 15 (2009) 83-90. DOI:10.1255/ejms.945
[25]
L. Song, M.A. Elsayed, J. Chem. Phys. 94 (1990) 7907-7913. DOI:10.1021/j100383a031
[26]
C. Berg, T. Schindler, G. Niedner-Schatteburg, V.E. Bondybey, J. Chem. Phys. 102 (1995) 4870-4884. DOI:10.1063/1.469535
[27]
J.M. Parnis, E. Escobar-Cabrera, M.G. Thompson, et al., J. Phys. Chem. A 109 (2005) 7046-7056. DOI:10.1021/jp0506944
[28]
S.R. Miller, T.P. Marcy, E.L. Millam, D.G. Leopold, J. Am. Chem. Soc. 129 (2007) 3482-3483. DOI:10.1021/ja068568g
[29]
B. Pfeffer, S. Jaberg, G. Niedner-Schatteburg, J. Chem. Phys. 131 (2009) 194305. DOI:10.1063/1.3264575
[30]
X. Tang, Y. Hou, C.Y. Ng, B. Ruscic, J. Chem. Phys. 123 (2005) 074330. DOI:10.1063/1.1995699
[31]
M.R. Zakin, R.O. Brickman, J. Chem. Phys. 127 (1988) 3555-3560.
[32]
J. Mwakapumba, K.M. Ervin, Int. J. Mass Spectrom. Ion Processes 161 (1997) 161-174. DOI:10.1016/S0168-1176(96)04469-2
[33]
E.D. Pillai, T.D. Jaeger, M.A. Duncan, J. Am. Chem. Soc. 129 (2007) 2297-2307. DOI:10.1021/ja0661008
[34]
M.A. Addicoat, K.F. Lim, G.F. Metha, J. Chem. Phys. 137 (2012) 034301. DOI:10.1063/1.4733287
[35]
Z. Luo, W.H. Woodward, J.C. Smith, A.W. Castleman Jr., Int. J. Mass. Spectrom. 309 (2012) 176-181. DOI:10.1016/j.ijms.2011.09.016
[36]
J.P. Perdew, Phys. Rev. B 34 (1986) 7406. DOI:10.1103/PhysRevB.33.8822
[37]
C. Lee, W. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785-789. DOI:10.1103/PhysRevB.37.785
[38]
T. Lu, F. Chen, J. Comput. Chem. 33 (2012) 580-592. DOI:10.1002/jcc.22885
[39]
W. Humphrey, A. Dalke, K. Schulten, J. Mol. Graphics Modell. 14 (1996) 33-38. DOI:10.1016/0263-7855(96)00018-5
[40]
E.J. Baerends, T. Ziegler, J. Autschbach, et al., ADF2018, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands. http://www.scm.com.
[41]
G.G. te Velde, F.M. Bickelhaupt, E.J. Baerends, et al., J. Comput. Chem. 22 (2001) 931-967. DOI:10.1002/jcc.1056
[42]
H. Zhang, H. Wu, Y. Jia, et al., Rev. Sci. Instrum. 90 (2019) 073101. DOI:10.1063/1.5108994
[43]
K.A. Zemski, D.R. Justes, A.W. Castleman, J. Phys. Chem. B 106 (2002) 6136-6148. DOI:10.1021/jp0142334
[44]
J.R. Sambrano, L. Gracia, J. Andrés, S. Berski, A. Beltrán, J. Phys. Chem. A 108 (2004) 10850-10860. DOI:10.1021/jp047229b
[45]
J.E. Bower, J. Chem. Phys. 85 (1986) 5373-5375. DOI:10.1063/1.451157
[46]
R.E. Leuchtner, A.C. Harms, A.W. Castleman Jr., J. Chem. Phys. 91 (1989) 2753-2754. DOI:10.1063/1.456988
[47]
E. Kapiloff, K.M. Ervin, J. Phys. Chem. A 101 (1997) 8460-8469.
[48]
Z. Luo, G.U. Gamboa, J.C. Smith, et al., J. Am. Chem. Soc. 134 (2012) 18973-18978. DOI:10.1021/ja303268w
[49]
R. Burgert, H. Schnöckel, A. Grubisic, X. Li, S.T. Stokes, Science 319 (2008) 438-442. DOI:10.1126/science.1148643
[50]
C. Geng, J. Li, T. Weiske, H. Schwarz, Proc. Natl. Acad. Sci. U. S. A. 115 (2018) 11680-11687. DOI:10.1073/pnas.1814610115
[51]
N. Chen, R.T. Yang, Ind. Eng. Chem. Res. 35 (1996) 4020-4027. DOI:10.1021/ie960299n
[52]
M.P. Mitoraj, A. Michalak, T. Ziegler, J. Chem. Theory. Comput. 5 (2009) 962-975. DOI:10.1021/ct800503d
[53]
T. Ziegler, A. Rauk, Theor. Chim. Acta 46 (1977) 1-10. DOI:10.1007/BF02401406
[54]
F.M. Bickelhaupt, E.J. Baerends, Rev. Comput. Chem. 15 (2000) 1-86.
[55]
M.V. Hopffgarten, G. Frenking, Wiley Interdiscip. Rev. Comput. Mol. Sci. 2 (2012) 43-62. DOI:10.1002/wcms.71
[56]
J. Chen, H. Zhang, X. Liu, et al., Phys. Chem. Chem. Phys. 18 (2016) 7190-7196. DOI:10.1039/C5CP06892D
[57]
J. Chen, Z. Luo, J. Yao, Phys. Chem. Chem. Phys. 19 (2017) 21777-21782. DOI:10.1039/C7CP02538F
[58]
Z. Luo, S. Khanna, Springer Nature Singapore Pte Ltd., Springer, 2020, p. 267.
[59]
A.C. Reber, S.N. Khanna, P.J. Roach, W.H. Woodward, A.W. Castleman, J. Am. Chem. Soc. 129 (2007) 16098-16101. DOI:10.1021/ja075998d