Ardita
Kilaj§
a,
Silvan
Käser¶
a,
Jia
Wang¶||
b,
Patrik
Straňák¶**
a,
Max
Schwilk
a,
Lei
Xu
a,
O. Anatole
von Lilienfeld
acdef,
Jochen
Küpper
*bghi,
Markus
Meuwly
*aj and
Stefan
Willitsch
*a
aDepartment of Chemistry, University of Basel, Klingelbergstrasse 80, 4056, Basel, Switzerland. E-mail: m.meuwly@unibas.ch; stefan.willitsch@unibas.ch
bCenter for Free-Electron Laser Science CFEL, Deutsches Elektronen-Synchrotron DESY, Notkestr. 85, 22607 Hamburg, Germany. E-mail: jochen.kuepper@cfel.de
cVector Institute for Artificial Intelligence, Toronto, ON M5S 1M1, Canada
dDepartments of Chemistry, Materials Science and Engineering, and Physics, University of Toronto, St. George Campus, Toronto, ON M5S 3H6, Canada
eMachine Learning Group, Technische Universität Berlin, 10587 Berlin, Germany
fBerlin Institute for the Foundations of Learning and Data - BIFOLD, Germany
gDepartment of Physics, Universität Hamburg, Luruper Chaussee 149, 22761 Hamburg, Germany
hDepartment of Chemistry, Universität Hamburg, Martin-Luther-King-Platz 6, 20146 Hamburg, Germany
iCenter for Ultrafast Imaging, Universität Hamburg, Luruper Chaussee 149, 22761 Hamburg, Germany
jDepartment of Chemistry, Brown University, Providence, RI 02912, USA
First published on 28th April 2023
Recent advances in experimental methodology enabled studies of the quantum-state- and conformational dependence of chemical reactions under precisely controlled conditions in the gas phase. Here, we generated samples of selected gauche and s-trans 2,3-dibromobutadiene (DBB) by electrostatic deflection in a molecular beam and studied their reaction with Coulomb crystals of laser-cooled Ca+ ions in an ion trap. The rate coefficients for the total reaction were found to strongly depend on both the conformation of DBB and the electronic state of Ca+. In the (4p)2P1/2 and (3d)2D3/2 excited states of Ca+, the reaction is capture-limited and faster for the gauche conformer due to long-range ion–dipole interactions. In the (4s)2S1/2 ground state of Ca+, the reaction rate for s-trans DBB still conforms with the capture limit, while that for gauche DBB is strongly suppressed. The experimental observations were analysed with the help of adiabatic capture theory, ab initio calculations and reactive molecular dynamics simulations on a machine-learned full-dimensional potential energy surface of the system. The theory yields near-quantitative agreement for s-trans-DBB, but overestimates the reactivity of the gauche-conformer compared to the experiment. The present study points to the important role of molecular geometry even in strongly reactive exothermic systems and illustrates striking differences in the reactivity of individual conformers in gas-phase ion–molecule reactions.
To rationalize the experimental findings, we modelled the kinetics and dynamics using adiabatic capture theory30 and reactive molecular dynamics simulations on a full-dimensional potential energy surface (PES) of the system trained on the results of ab initio calculations by a neural network (NN). The theory yields excellent agreement with experiment for the reaction of s-trans-DBB, but overestimates the reactivity of the gauche-conformer by almost an order of magnitude suggesting the presence of an as yet unaccounted dynamic bottleneck along the gauche reaction pathway. Possible explanations are discussed. The present study highlights the important role of molecular conformation in gas-phase ion–molecule reactions of polyatomic species and the role of both long-range and short-range effects in determining conformational differences in chemical reactivity.
The reaction of Ca+ with the gauche and s-trans DBB conformers proceeded via individual pathways with bimolecular reaction rate coefficients kg and kt, respectively (Fig. 1b). Throughout the reaction, the fluorescence of Ca+ ions due to laser cooling on the 2S1/2 → 2P1/2 transition at 397 nm was imaged onto a camera (Fig. 1b). As the reactions progressed, the Coulomb crystals changed shape due to loss of Ca+ ions and accumulation of heavier product ions around the Ca+ core. Quantitative analysis of the reaction kinetics and products was performed by ejecting the trapped ions into a time-of-flight mass-spectrometer (TOF-MS) radially coupled to the ion trap.31
![]() | (1) |
![]() | (2) |
![]() | (3) |
![]() | (4) |
nCa+(t) = nCa+(0)![]() ![]() | (5) |
![]() | (6) |
The Monte Carlo simulation was used to determine the conformer populations as a function of deflection coordinate (Fig. 3b). Three beam positions, marked I–III in Fig. 3a and b, corresponding to the pure s-trans conformer (I), a thermal mixture (II) in the undeflected beam, and the pure gauche conformer (III), were chosen for reaction-rate measurements. In each position, the Ca+ ion count was measured as a function of reaction time (Fig. 3c). The background loss rate of Ca+ was determined separately by adjusting the molecular beam such that it did not hit the Coulomb crystal (grey data points in Fig. 3c). All traces exhibit an exponential decay of the number of Ca+ ions which confirms the validity of using a pseudo-first-order rate law for the bimolecular reaction of DBB + Ca+ with a constant DBB density. Pseudo first-order rate coefficients tot,i (i = I–III) were obtained by fitting exponential-decay models to the data and subtracting the corresponding background rate. From the
tot,i, the bimolecular rate coefficients ktot,i =
tot,i/ni were calculated using the DBB beam densities ni at each position i = I–III in the deflection profile.
Fig. 3d shows the measured bimolecular rate coefficients ki as a function of the s-trans population pt obtained from the Monte–Carlo simulation. The bimolecular rate coefficient for the depletion of Ca+ was modeled as the linear combination ktot,i = pg,ikg + pt,ikt of the conformer-specific rate coefficients kg/t (gauche/s-trans). The weighting factors pg/t,i are the respective conformer populations at location i (Fig. 3b). A least-squares fit (solid line in Fig. 3d) with this model was applied to the data and yielded the bimolecular reaction rate coefficients kg = 0.92(11) × 10−9 cm3 s−1 for the gauche-conformer and kt = 1.30(13) × 10−9 cm3 s−1 for the s-trans-conformer. This implies a relative difference rexp = 2(kg − kt)/(kg + kt) = −0.35(7) by which the s-trans-conformer is observed to react faster than the gauche-conformer.
To measure the reaction rate coefficients as a function of the excited-state population of Ca+, a set of detunings Δ397 of the 397 nm laser labelled i–iii in Fig. 4a was chosen that samples a combined 2P1/2 + 2D3/2 population between 0.25 and 0.7. At each detuning, one rate measurement was conducted with pure s-trans DBB (molecular beam position I in Fig. 3a) and three measurements for pure gauche DBB (molecular beam position III). Results of the state- and conformer-resolved rate measurements are presented in Fig. 4b. A linear model kg/t,x = (1 − px)kg/t;S + pxkg/t;P+D was fitted to the measured rate coefficients kg/t,x as a function of the 2P1/2 + 2D3/2 excited state population px (x = i, ii, iii) to retrieve the ground-state (kg/t;S) and excited-state (kg/t;P+D) rate coefficients. Due to the strong correlation between the populations and rate coefficients in the 2P1/2 and 2D3/2 states obtained in the fit, no statements about their individual reaction rates could be made and only the effective rate coefficient averaged over both excited states is given.
Strikingly, the observed dependence of the rate coefficient on the Ca+ excited state population differs strongly between the two conformers of DBB. While the rate coefficient remains nearly constant for s-trans-DBB, a clear increase with excited-state population is observed for the gauche conformer, pointing to a considerably reduced reaction rate for the gauche species with ground-state Ca+. The fitted bimolecular rate coefficients of s-trans DBB with Ca+ in ground and excited states, kt;S = 1.4(3) × 10−9 cm3 s−1 and kt;P+D = 1.5(4) × 10−9 cm3 s−1, respectively, lie within one standard deviation from each other. For the gauche conformer, however, the fitted bimolecular rate coefficients are kg;S = 0.3(2) × 10−9 cm3 s−1 for Ca+ in its ground state and kg;P+D = 2.2(3) × 10−9 cm3 s−1 for the excited states.
The measured rate coefficient for the s-trans conformer agrees with the calculated value within the uncertainty limits. By contrast, for the reaction of gauche-DBB with Ca+ in its ground electronic state, adiabatic capture theory predicts a reaction rate coefficient about an order of magnitude larger than the experimental results. For reactions with electronically excited Ca+ ions, however, capture theory and experiment again agree within the experimental uncertainties. It can thus be concluded that all but one combinations of both conformers of DBB and the electronic states of Ca+ considered, the kinetics and dynamics of the title reaction can adequately be described by capture theory and are thus governed by long-range intermolecular forces, in this case ion–dipole interactions. The exception is the reaction of gauche-DBB with Ca+ in its ground state, which was found to be considerably slower than the capture limit. Thus, in order to gain a more in-depth understanding of the reaction dynamics, additional computational studies were carried out including electronic structure calculations and reactive molecular dynamics simulations.
Fig. 5 depicts the energetically lowest reaction channels found by a scan of possible chemically distinct reaction pathways, as well as products, based on the assumption that the formation of the Ca–Br ionic bond is the driving force of the overall reaction. Fig. S6 of the ESI‡ shows all reaction pathways identified including ones with higher barriers. Intrinsic reaction coordinates (IRCs) and zero-point-energy corrections have been obtained at the spin-unrestricted B3LYP/TZ level of theory. Electronic energies of the stationary points have been recomputed at the CCSD(T)-F12/DZ level of theory. All computational details, including quality checks for the IRCs and single-reference-character checks of the PES are described in the Methods section and the ESI.‡ Spin-restricted Hartree–Fock and B3LYP orbitals of the stationary-point structures were used for a chemical analysis of the reaction channels (the most important valence orbitals for every stationary-state geometry are depicted in Tables S4 and S5 of the ESI‡).
The reaction of Ca+ (4s) 2S1/2 with DBB can initiate via the addition of Ca+ to either the bromine moieties or to the π-electron system of DBB. The relevant reaction intermediates are shown in Fig. 5 (I2 and I1t/g). For both s-trans-DBB and gauche-DBB, an η4-coordination of a singly occupied orbital of the π4 system of DBB− to Ca2+ emerges as a stable intermediate (I1t/g) with similar energies. In these intermediates, a single electron transfer from Ca+ to DBB has occurred and the Ca2+ moiety is stabilized by strong σ-back-donation from the bromine atoms. Our results suggest that the s-trans adduct is energetically slightly favoured over the gauche adduct (energy difference ≈3 kcal mol−1 (0.13 eV)). For gauche-DBB, we also found an adduct intermediate I2 similar in energy to I1g in which both bromine groups coordinate to Ca+. In this case, the calcium moiety has not undergone an electron transfer. These intermediates in the three entrance channels stabilize by 23–25 kcal mol−1 (1.00–1.08 eV) with respect to their reactant state via a barrier-less reaction path.
From all three entrance-channel intermediates (I1t/g and I2), a Br− migration from the DBB moiety to calcium over a low-lying transition state (activation energy <3 kcal mol−1 [0.13 eV]) was found forming intermediate I3. This intermediate displays a strong energetic stabilisation (≈46.5 kcal mol−1 [2.02 eV] with respect to the gauche-DBB reactant state) and exhibits an allene radical structure with the remaining DBB bromine coordinatively binding to CaBr+. CaBr+ dissociation from I3 yields the product P1 (CaBr+, C4H4 Br) which is energetically stabilized by ≈15.7 kcal mol−1 (0.68 eV) with respect to the gauche-DBB reactant state. The ionic product of this pathway, CaBr+, is observed in the experimental mass spectra (Fig. 2(a)).
Another pathway originating from I3via its conformer I4 was found to feature a second Br− migration to CaBr+viaTS4 to I5. In I5, perpendicular π4 and π2 systems have formed with the singly occupied molecular orbital (SOMO) being the HOMO of the π4 system (for depictions of the orbitals see Table S4 of the ESI‡). Strong σ-back-donation from CaBr2 stabilizes the uncommon electronic structure of the C4H4+ species. Dissociation of CaBr2 then results in the energetically most favoured product P2 (CaBr2, C4H4+) where the SOMO exhibits a spiral π-system, typical for cumulenes. P2 is energetically stabilized with respect to the gauche-DBB reactant state by ≈23.6 kcal mol−1 (1.02 eV). The ionic product of this pathway, C4H4+, and fragmentation products thereof are observed in the experimental mass spectra (Fig. 2(a)). I5 was also found to connect to a 1,2-migration of Br− to the terminal carbon center, yielding I6 which also exhibits an allene-radical electronic structure. This low-energy structure (stabilized by 52.1 kcal mol−1 (2.26 eV) with respect to the gauche-DBB reactant state) can then evolve via CaBr+ dissociation to P3 (CaBr+, C4H4Br) which is close in energy to P1.
The initial conformer-specific reaction channels connect to the same intermediates after the first Br− migration in the formation of I3. All barriers found are clearly submerged below the energies of the reactants, rendering all studied reaction pathways effectively barrierless. For the reaction of the s-trans conformer, this finding rationalizes the good agreement between the experimental and theoretical rate coefficient which was computed assuming barrierless capture dynamics. However, the barrierless minimum-energy paths displayed in Fig. 5 do not provide an explanation for the significantly lower reaction-rate coefficient observed for gauche-DBB.
Opacity functions for reactive collisions of the gauche and s-trans conformers are shown in Fig. 6. The results were obtained starting from 1000 initial conditions per impact parameter interval. To assess the role of rotational excitation in the dynamics, opacity functions were obtained for a range of rotational energies of the molecule corresponding to different rotational angular momentum quantum numbers j. The data suggest that the capture range for gauche-DBB is larger by ≈1 Å compared to the s-trans conformer irrespective of the molecular rotation. For both isomers, however, rotational excitation (j > 0) reduces the capture range leading to a smaller maximum impact parameter bmax and thus a reduced reaction rate. Note that j = 20 corresponds to rotational energies Erot ≈ 0.09 and 0.18 kcal mol−1 (3.90 and 7.81 meV) for gauche-DBB and s-trans-DBB, respectively, while for j = 29, the corresponding values are Erot ≈ 0.19 and 0.38 kcal mol−1 (8.24 and 16.48 meV). Under the present experimental conditions, both gauche- and s-trans-DBB populate considerably lower rotational levels with the largest population found in j = 4.17 Thus, in the experiment Erot < 0.01 kcal mol−1 (0.43 meV). Consequently, based on the present MD simulations, it can be surmised that the small rotational excitation of the molecules has a small influence on the observed reaction rates. We note, however, that these MD simulations, due to their purely classical-electrodynamics description, cannot reproduce effects due to weak-field-seeking behaviour in the approach of the molecule and ion (see Discussion below).
The reaction rate coefficients computed from the opacity function in Fig. 6 are listed in Table 1. Irrespective of rotation, the computed rate coefficients for gauche-DBB are larger by about 20% compared to the s-trans conformer. The rate coefficients obtained from the simulations are in good agreement with the capture theory and with experiment for the s-trans conformer, but fail to reproduce the experimental rate coefficient for reactions with gauche-DBB.
gauche | s-trans | |
---|---|---|
PhysNet (j = 0) | 2.5 × 10−9 | 2.0 × 10−9 |
PhysNet (j = 20) | 2.2 × 10−9 | 1.5 × 10−9 |
PhysNet (j = 29) | 1.7 × 10−9 | 1.5 × 10−9 |
Capture theory | 2.4 × 10−9 | 1.3 × 10−9 |
Exp. | 3(2) × 10−10 | 1.4(3) × 10−9 |
One difference between the two DBB isomers is that gauche-DBB features an ion–dipole barrier for certain directions of approach in the entrance channel. If the dynamics was sensitive to such a barrier in the entrance channel, this could explain the reduction of the rate for gauche- vs. s-trans-DBB. In order to test this hypothesis explicitly and to better characterize the underlying dynamics, 1000 trajectories with impact parameter b = 0 Å were analysed for both, gauche- and s-trans-DBB. For the polar gauche conformer, the long-range ion–dipole interaction with Ca+ causes about 50% of the area of the molecule to be shielded by the repulsive ion–dipole barrier that rises up to ∼3 kcal mol−1 (0.13 eV). This might suggest that collisions along these directions lead to unproductive collisions and therefore reduce the total rate. For trans-DBB, lower repulsive barriers with heights <1 kcal mol−1 (0.043 eV) covering about ∼30% of the circumference were found along the approach of Ca+ towards the two methylene groups.
The dynamics simulations, however, indicate that the ion–dipole barrier present for gauche-DBB does not reduce the reaction rate because during the collisions the reactants reorient towards the most favourable approach, i.e., Ca+ attacking the negative end of the molecular dipole, and thus avoid the barrier (Fig. 7). An alignment effect also occurs for s-trans-DBB as shown in Fig. 8. Because of these dynamics, it can be surmised that the barriers in the entrance channels of the collisions of both conformers do not play a significant role in the dynamics under the present conditions. As an additional verification that the barriers in the entrance channel are not responsible for the disagreement between observed and computed rates, additional simulations with collision energies decreased by up to a factor of ∼3 were carried out. The result was an increase of the rate coefficient for the gauche-conformer in line with capture theory, which is, however, still at variance with the much smaller value obtained from experiment. Hence, the barriers in the entrance channel are not decisive for the differences in the reaction rates because the long-range intermolecular interactions orient gauche-DBB into a reaction-competent fashion and the barrier is never sampled for the collision energies relevant and considered here.
![]() | ||
Fig. 7 Reorientation dynamics of the gauche-DBB + Ca+ system over the course of reactive trajectories for impact parameter b = 0 Å. The figure shows the position of Ca+ with respect to DBB 4, 6 and 8 ps after the start of the simulation for 1000 different initial conditions (collision angles). The orientation of DBB was fixed in the illustration. Intermolecular forces steer the Ca+ ion towards the negative end of the molecular dipole, favouring the reaction pathway via intermediate I2 in Fig. 5. The distances between the centre of mass of DBB and Ca+ are ∼41/25/7 Å after 4/6/8 ps. The full distributions of the distances is given in Fig. S9 of the ESI.‡ |
![]() | ||
Fig. 8 Alignment dynamics of the s-trans-DBB + Ca+ system over the course of reactive trajectories for impact parameter b = 0 Å. The figure shows the position of Ca+ with respect to DBB 4, 6, and 8 ps after the start of the simulation for 1000 different initial conditions (collision angles). The molecular frame of DBB was fixed in the illustration. Intermolecular forces steer the Ca+ ion toward a planar delocalization, or antialignment. After 8 ps, the system has transitioned toward a preferential reaction pathway passing through (the two symmetrically equivalent) I1t states in Fig. 5. The full distributions of the distances of the collision partners are given in Fig. S10 of the ESI.‡ |
The most apparent difference between the two conformers is that s-trans-DBB only features one reaction pathway whereas for gauche-DBB, two paths are in principle accessible, see Fig. 5. However, a closer evaluation of the trajectories in Fig. 9 reveals that for the gauche species, the majority of reactive trajectories exclusively pass through intermediate I2 in which the Ca+ moiety is coordinated to the bromine atoms of DBB. While the small energetic differences between the two pathways are unlikely to account for this dynamic bias, this finding is in line with the reorientation dynamics discussed above: during the approach of Ca+, the negative end of the molecular dipole formed by the bromine atoms of DBB orients towards the ion which directly leads to intermediate I2 from where the reaction proceeds further, usually in a direct abstraction process. Conversely, for s-trans-DBB the reaction is found to proceed via two symmetric I1t intermediates in which the Ca+ ion is coordinated to the π electron system of the molecule as shown in Fig. S13 of the ESI.‡
![]() | ||
Fig. 9 Evaluation of the reaction pathway for the gauche-DBB simulations for impact parameter b = 0 Å and j = 0. Note that 62 out of the 1000 trajectories were excluded from the analysis, because they also transitioned towards trans-DBB before reaction. The analysis shows that the reaction happens exclusively via intermediate I2 and does not sample the path viaI1g (see Fig. 5). The trajectories are shown up to reaction, i.e. up to the point for which a C–Br bond is broken and exceeds 3.0 Å (the equilibrium bond distance is 1.92 Å). The turquoise points mark the snapshot at which (one of the) Br–Ca ≤ 2.5 Å for the first time (the equilibrium distance of Br–Ca in I3 is ∼2.5 Å). The analysis of all 1000 trajectories is given in Fig. S12 of the ESI.‡ |
The lifetimes of the collision complexes were found to be generally short (<1 ps) for both DBB isomers. Thus, a possible slowdown of the reaction due to formation of a long-lived reaction intermediate cannot be the origin for the discrepancy between experiment and simulations for the gauche species. Typically, the simulations yielded direct abstraction reactions. Only for the minority of trajectories, lifetimes of the reaction complex longer than several vibrational periods (i.e., longer than 1 ps) were observed. As discussed above, rotational excitation was found to change the reaction rates for both isomers in a similar fashion.
Concerning the quality of the PES, there are two main factors that influence the computed rates, first the quality of the PhysNet fit (i.e. how accurately the NN reproduces the ab initio reference data) and second, the chosen level of theory and basis set of the reference data.
First, the PhysNet representation of the global PES is accurate. Fig. S14 of the ESI‡ compares the ab initio with the PhysNet energies along two representative trajectories. The NN-based PES faithfully reproduces the DFT energies, which was also found when predicting energies for a test set that was not used during training. The out-of-sample mean absolute and root mean squared errors (MAE and RMSE) for predicting energies for ∼10 000 structures and covering an energy range of 230 kcal mol−1 (9.97 eV) amount to MAE(E) = 0.06 and RMSE(E) = 0.3 kcal mol−1 (2.6 and 13.0 meV), respectively. Thus, the quality of the fit can be judged excellent and fitting errors can be ruled out as a source of the discrepancy.
Second, an insufficient level of theory of the ab initio PES may be another reason for the discrepancy. A comparison between Fig. S6 of the ESI‡ and Fig. 5 illustrates the overall similarity of the B3LYP PES used for training the NN-PES and running the reactive MD simulations with the one obtained at the higher level of theory CCSD(T)-F12. However, as the energies of some stationary points differ by up to a few kcal mol−1, an effect of the level of theory cannot be completely ruled out. For example, transition state TS1g is higher in energy by 1.4 kcal mol−1 (0.06 eV) than TS2 at the CCSD(T) level but in the B3LYP calculations, the ordering of the two states is reversed and the energy difference increases to 3.3 kcal mol−1 (0.14 eV). Notwithstanding the fact that the relevant barriers here are strongly submerged with respect to the energy of the reactants, we note that within transition state theory one order of magnitude difference in the rate corresponds to an energy scale of ∼1.5 kcal mol−1 (0.065 eV) which can, e.g., be compared with an uncertainty of ∼4 kcal mol−1 (0.17 eV) in the ordering of TS1g and TS2 mentioned above.
In addition, the use of classical MD simulations as opposed to a full quantum treatment of the nuclear dynamics may be questioned. However, the masses of the atoms involved in the reaction as well as the energies along the reaction pathway are both large which make quantum effects as a source for the disagreement unlikely. Moreover, it was recently demonstrated that quasi-classical trajectory simulations yield energy-dependent rates that compare favourably with quantum wavepacket calculations for the C + O2 → O + CO reaction in its electronic ground state down to collision energies of a few 10 K.36
On the other hand, molecules in low rotational states can display significant weak-field-seeking behaviour which leads to anti-orientation of the dipole in the electric field.37 This effect is entirely quantum mechanical in nature and in the present case it is expected to be particularly important for intramolecular geometries between the long-range electrostatically dominated regime and chemical-bonding distances. From simple estimates and for the relevant ion–molecule distances of a few nanometer to a few 100 picometer, both the translation time and the re-orientation time of the molecular dipole in the field of the ion are on the order of some picoseconds. In the most obvious way, this would lead to “inverted” approach geometries and thus reduce the reaction rate for gauche DBB, but more generally this effect indicates that intricate details of the PES could be very relevant with respect to modified entrance-channel geometries and dynamics. In the present experiment, low rotational states for DBB are most prevalent and neglect of this in the MD simulations is potentially critical for correctly describing the reaction dynamics. Full dimensional quantum dynamics simulations for systems of this size are currently unfeasible, but a possible alternative is to precompute geometry-dependent Stark-effect-corrected effective interaction energies and add those to the energy function used in the MD simulations. This, however, is beyond the scope of the present work.
Lastly, multi-reference, non-adiabatic and spin–orbit effects could influence the dynamics in the present system. Checks for a possible spin contamination of the stationary points displayed in Fig. 5 remained inconspicuous, see Tables S2 and S3 of the ESI.‡ Thus, while these tests did not reveal any obvious multireference effects in the present system, a possible role of non-adiabatic dynamics cannot be ruled out until the lowest excited states of the system and potential crossings between them have been thoroughly explored which, however, is outside the scope of the present study. Along similar lines, the NN-based PES used in the trajectory calculations is unable to account for long-range charge transfer. All electronic effects are “only” included within the concept of a Born–Oppenheimer PES. It is conceivable that electronic structure calculations at almost all levels of theory are incapable of capturing such effects. As yet another possibility, it is also imaginable that spin–orbit interactions in the two conformers differ such as to stabilize a gauche-DBB intermediate to lower the rates.
Finally, we note that an isomeric effect was also observed in ref. 38 where the reaction of N+ ions with s-cis-dichloroethylene (C2H2Cl2) was found to be considerably slower than predicted by capture theory while the rate coefficient of the s-trans species agreed very well with the theoretical predictions. In that study, it was speculated that the reason for this discrepancy for s-cis-C2H2Cl2 is a mismatch between the direction of most favoured attack for short-range charge transfer and the direction of the long-range approach of the collision partners along the axis of the molecular dipole. While this explanation is in line with the finding in the present system that the attack of Ca+ occurs exclusively along the direction of the molecular dipole of the gauche conformer of DBB leading to the formation of intermediate I2, it does not rationalize the reduced reactivity in the present case according to our trajectory calculations.
The experimental findings were analysed using adiabatic-capture-theory calculations and reactive molecular dynamics simulations on a full-dimensional, machine-learned PES. For both isomers the abstraction reaction for forming CaBr+ is direct with rare formation of an intermediate with lifetimes longer than 1 ps. For the s-trans conformer, the simulations yielded near-quantitative agreement with experiments for the rate coefficient. This is consistent with findings for reactive atom + diatom systems using similar computational approaches.36 On the other hand, the computations overestimated the reaction rate for the gauche conformer, which features two possible reaction pathways, by an order of magnitude compared with experiment. Again, this has been found, for example, for the MgO+ + CH4 reaction, for which the computed rate differed from experiments by one order of magnitude but showing the correct temperature dependence.39 The reason for the discrepancies in both cases may be limitations in the electronic structure methods that can be applied to the systems of interest given their size, omitting non-adiabatic effects, or neglect of purely quantum-mechanical re-orientation effects of the cold polar molecule in the electric field of the ion. Further studies are required to clarify these possibilities.
The present study highlights the important role of molecular conformation in gas-phase ion–molecule reactions already in moderately complex polyatomic systems. While previous studies of conformer-selected ion–molecule reactions14,18 mainly uncovered conformational dependencies due to differences in long-range interactions, the current results suggest that short-range conformational effects strongly suppress the reactivity of the gauche-conformer, which still needs to be explained.
Moreover, although the actual – time-resolved – elementary dynamics of similarly controlled molecular systems were experimentally observed,40 unravelling the underlying dynamics of bimolecular reactions such as the present one for now seems a formidable task, especially regarding problem of defining the starting time of the individual reactions but also the low occurrence statistics of ongoing reactions. The current combination of measuring time-averaged rates and molecular dynamics simulations consistent with experimental findings is a meaningful starting point for disentangling the chemical dynamics.
The ion trap was radially coupled to a TOF-MS orthogonal to the molecular-beam propagation axis for the mass and quantitative analysis of reactant and product ions.31 Two different modes of operation were used for the TOF-MS. For the determination of mass spectra of the reaction products, a low-resolution mode was used to extract ions into the TOF-MS by applying a 1 μs long pulse of 4.0 kV to the repeller electrode. For the rate measurements, additional high-voltage pulses, delayed by 0.45 μs, were applied to the extractor electrodes to selectively enhance the resolution for the Ca+ and C4Hn+ species.31 Ions were detected using a microchannel plate detector (MCP, Photonis USA) operated at a voltage of typically 2.3 kV placed at the end of the flight tube.
Intrinsic bond orbitals (IBOs)52 of restricted open-shell B3LYP/def2TZVPP and HF/VDZ-F12 wave functions of the stationary point structures were computed for the chemical analysis. The orbitals depicted in Tables S4 and S5 of the ESI‡ have been visualized by the IboView program package.53 The spin contamination of the wavefunctions for the stationary points were found to be low (see Tables S2 and S3 of the ESI‡).
Reference data including energies, forces and dipole moments was calculated at the B3LYP/def2-TZVPP using Orca.54 Relevant geometries for the collisions of Ca+ with both gauche- and s-trans-DBB were obtained by running Langevin dynamics using the semiempirical tight binding GFN2-xTB method.55 Random momenta drawn from a Maxwell–Boltzmann distribution corresponding to 300 K were assigned to DBB (note that the resulting DBB configurations also contain configurations relevant to lower temperatures56). Diverse collision angles for the reaction were obtained by choosing the Ca+ position randomly on a sphere around DBB before accelerating it towards DBB. After training PhysNet on this initial data set, new geometries were generated using adaptive sampling.57,58 Adaptive sampling is used to detect regions on the PES which are underrepresented in the reference data set using an ensemble of PhysNet models. The final data set contained 191982 structures covering the separated fragments and reaction products.
PhysNet was trained following the procedure outlined in ref. 35 and using the standard hyperparameters as suggested therein. The resulting PES of the studied system is given by
![]() | (7) |
Simulations were run in the NVE ensemble using the velocity Verlet integrator as implemented in the atomic simulation environment60 with a time step of Δt = 0.25 fs. The simulations were terminated based on two distance criteria, i.e., for r(C–Br) > 3 Å and r(Ca–Br) < 3 Å a reaction was considered to have occurred. Additionally, the simulations were stopped and considered “nonreactive” if, after initial approach, DBB and Ca+ were separated by more than 30 Å without satisfying the criteria for a “reactive” trajectory. At the beginning of the simulation, the structure of DBB was optimized and remained vibrationally cold.
Two sets of simulations were carried out: The first was run without assigning any rotational energy to the DBB molecules. This is justified by the low temperature (few K) of DBB in the experiment. The second set included rotation of the DBB molecule to study the influence of classical rotation in general. For this, the moments of inertia IA, IB, IC together with the three principal axes of rotation were determined and a rotation was assigned to the principal axis with lowest IA. Simulations were carried out at different rotational energies ER = j(j + 1)B where B is the rotational constant around the principal axis with lowest IA.
![]() | (8) |
k = σ·vrel | (9) |
Footnotes |
† In memoriam Max Schwilk (1986–2022). |
‡ Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cp01416a |
§ Present address: Novartis Institutes for Biomedical Research, Fabrikstrasse 2, 4056 Basel, Switzerland. |
¶ These authors contributed equally to the present work. |
|| Present address: Department of Chemistry, University of Hawaii at Manoa, Honolulu, HI 96822, USA. |
** Present address: Fraunhofer Institute for Applied Solid State Physics IAF, Tullastrasse 72, 79108 Freiburg, Germany. |
This journal is © the Owner Societies 2023 |