In the theoretical calculations reported here, we show that the hydration of the Criegee intermediate within the sub-nanospace of fullerene cages occurs differently in different fullerenes, thereby providing evidence for the size-dependent reactivity inside these exotic carbon cages. Upon C70 or C84 encapsulation, the Criegee hydration occurs instantaneously without any activation barrier, whereas inside the C120 cage, the hydration involves a small barrier of 4.4 kcal/mol. Our Born-Oppenheimer molecular dynamics simulations suggest that the Criegee intermediate and the product of its hydration, α-hydroxy methyl hydroperoxide, remain dynamically stable over 20 ps time scale at the 300 K, implying that fullerene cages provide a robust framework for reactivity.
I. INTRODUCTION
Fullerenes are promising candidates for catalysis. The inner spherical radii of the fullerenes C60 and C70 are 3.7 and 4.6 Å, which are large enough to encapsulate atoms and molecules, and provide potential environments for endohedral chemistry.1,2 In recent years, the arc-discharge methods and the emergence of a unique synthetic approach, molecular surgery,3,4 have allowed the synthesis of fullerene complexes that can encapsulate a wide variety of species, including metal ions,5,6 noble gases,7,8 molecular hydrogen,9,10 water molecules,11,12 hydrogen fluoride (HF),13 and nitrogen atoms.14 These complexes have attracted a lot of attention and have been the focal point of numerous experimental and theoretical studies. The synthetic feasibility of encaging small molecules inside the sub-nanospace of fullerenes raises a mechanistically appealing fundamental question: how does the nanoscale confinement influence the properties and reactivity of the enclosed species?
The available limited experimental and theoretical data suggest that the cavity inside fullerenes provides a unique environment for chemical reactivity.10,15–24 The confined molecules display the quantization of their coupled translational and rotational degrees of freedom15 and exhibit phenomena such as nuclear spin isomerism16 and ortho-para conversion.10,17 The nuclear spin conversion of the caged water molecules in H2O@C60 impacts the dielectric constant of the material.18 The rotational and vibrational constants of the encapsulated HF molecules are red-shifted relative to free HF.13 The 13C resonance for Xe@C60 is shifted downfield by 0.95 ppm.19 129Xe NMR showed a line 179.2 ppm downfield from xenon gas. The free-energy barrier at 298 K for the hydrogen exchange reaction 3H2 → 3H2 is reduced from 88.8 kcal/mol for the free reaction to 36.2 kcal/mol for the reaction within C60, reflecting at least 3 orders of magnitude rate enhancement due to the encagement.17 The presence of noble gas atoms or molecular hydrogen influences the equilibrium constant for a Diels-Alder (DA) reaction between 9,10-dimethylanthracene and C60 or C70.20
The insertion of metal ions into the fullerene cage is also found to greatly promote the DA cycloadditions to C60. Matsuo experimentally studied the DA cycloaddition of 1,3-cyclopentadiene and 1,3-cyclohexadiene to Li+@C60.21,22 These reactions were found to be 1000- and 2700-fold faster than the corresponding additions to the 6–6 bond of empty C60, respectively.21,22 The size of a fullerene cage also plays an important role in modifying chemical behavior inside a given cage. For example, the shape of the inner metal carbide (M2C2) cluster in endohedral fullerenes is found to be influenced by the size of a fullerene cage.23 Dorn and co-workers suggested that decreasing the size of the fullerene would compress M2C2 clusters from a nearly linear, stretched geometry to a constrained butterfly structure.23 Deng and Popov also predicted similar cage-dependent behavior of M2C2 units.24 Though the properties and intramolecular chemistry of the encapsulated single atoms or molecules have been studied previously, the bimolecular reactivity inside the fullerene cavity is yet to be explored.
We report the results of quantum chemical calculations and Born-Oppenheimer molecular dynamics (BOMD) simulations on the reaction between Criegee intermediates and water inside fullerene cages of different sizes. Criegee intermediates are carbonyl oxides that are key branching points in the ozone-olefin cycloadditions25 and are important because of their role in the nighttime chemistry of the OH radical in the troposphere. Ever since Criegee intermediates have directly been detected in the gas-phase,26,27 interest in their chemistry has grown exponentially.28 Criegee intermediates are not only atmospherically important but are also relevant in general synthetic organic reactions en route to value-added chemicals.29,30 In the results reported here, we show that the hydration of a Criegee intermediate in different sized fullerene cages occurs differently and provides the compelling evidence of a size-dependent reactivity within the sub-nanospace of a fullerene. These results not only offer deeper mechanistic insights into the effect of fullerene-cage confinement on the Criegee chemistry but should also open up new avenues of experimental and theoretical studies for studying other bimolecular processes inside fullerene cages.
II. COMPUTATIONAL DETAILS
In the first step, the gas-phase reaction of the simplest Criegee intermediate, CH2OO with H2O inside fullerenes C60, C70, C84, and C120 was examined at the density functional theory (DFT) level. The geometries of all the stationary points on the potential energy surfaces were fully optimized using the M06-2X31 density functional theory method and cc-pVDZ and cc-pVTZ32 basis sets. All fullerene calculations were done at the M06-2X/cc-pVTZ level except for the C120 ones that were done at the M06-2X/cc-pVDZ level. These gas-phase calculations were performed using Gaussian 0933 software and at a standard state of 298.15 K and 1 atm. The Cartesian coordinates of key optimized species are provided in the supplementary material document.
In the second step, the BOMD simulations were performed on the basis of DFT methods as implemented in the CP2K34 code. In the BOMD simulations, the fullerenes C70, C84, and C120 and one CH2OO molecule were considered. The dimensions of the simulation box are x = 35 Å, y = 35 Å, and z = 35 Å, which are large enough to avoid interactions between adjacent periodic images of fullerene cages. Prior to the BOMD simulations, the system was fully relaxed using the DFT method. For optimizing the system at the DFT level, the exchange and correlation interactions of electrons are treated with the Becke-Lee-Yang-Parr (BLYP)35,36 functional, and Grimme’s dispersion correction method37 is applied to account for the weak dispersion interaction. A double-ζ Gaussian basis set combined with an auxiliary basis set38 and the Goedecker-Teter-Hutter (GTH) norm-conserved pseudopotentials39,40 is adopted to treat the valence electrons and the core electrons, respectively. An energy cutoff of 280 Ry was set for the plane wave basis set and 40 Ry for the Gaussian basis set. All the BOMD simulations were performed in the constant volume and temperature (NVT) ensemble, with the Nose-Hoover chain method for controlling the temperature of the system. The integration step was set as 1 fs, which had been proven to achieve sufficient energy conservation for the water system. All the BOMD simulations were performed at a constant temperature of 300 K.
III. RESULTS AND DISCUSSION
A. Energetics and implications for broader endohedral fullerene chemistry
First, we examined the calculated structure and energetics of the C60 C70, C84, and C120 encapsulated CH2OO (CH2OO@C60, CH2OO@C70, CH2OO@C84, and CH2OO@C120). The equilibrium geometries of the encapsulated CH2OO are shown in Fig. 1. The comparative analysis of CH2OO with and without a fullerene cage (Table S1 of the supplementary material) reveals that the encapsulation does not induce any major structural change in the Criegee intermediate. The calculated interaction energy for CH2OO@C60 is +15.8 kcal/mol (Fig. 1) suggests that the encaging of CH2OO inside C60 is energetically not favorable. On the other hand, the calculated interaction energies for CH2OO@C70, CH2OO@C84, and CH2OO@C120 are −27.6, −24.8, and −21.0 kcal/mol, respectively. These interaction energies are at least 36.8 kcal/mol more favorable than that for CH2OO@C60, indicating that the encapsulation in a larger fullerene cage (>C60) stabilizes CH2OO, and the laboratory synthesis of the endohedral complexes of C70, C84, and C120 with Criegee intermediates might be achievable. We also examined the effect of C70 confinement on the larger C2 Criegee intermediates. Interestingly, the binding energy of anti-CH3CHOO@C70 is +25.3 kcal/mol (Fig. S1 of the supplementary material), suggesting that the encapsulation of anti-CH3CHOO inside the C70 cage is strongly disfavored. On the other hand, the encapsulation energy of syn-CH3CHOO is only 6.1 kcal/mol above the thermoneutral zero (Fig. S1 of the supplementary material), suggesting that selecting a larger fullerene cage will make the endohedral chemistry of syn-CH3CHOO feasible. This stark difference in the binding preference of two C2 conformers toward fullerene C70 is due to the presence of an intramolecular hydrogen bond between the terminal Criegee oxygen and the —C—H fragment of the syn-methyl group in syn-CH3CHOO@C70 that avoids steric clashes between Criegee intermediates and the walls of the fullerene cage. It is important to mention here that the presence of an intramolecular hydrogen bonding in syn-CH3CHOO makes it more stable in the gas-phase than anti-CH3CHOO. Specifically, Anglada et al. predicted syn-CH3CHOO to be 3.49 kcal/mol more stable than anti-CH3CHOO in the gas-phase at the CCSD(T)/aug-cc-pVTZ//B3LYP/6-311 + G(2df,2p) level of theory.41 Though this intramolecular hydrogen bond has traditionally been used to explain the slower gas-phase reactivity of syn-CH3CHOO relative to anti-CH3CHOO,41,42 the results here suggest that this interaction could be utilized to guide the endohedral chemistry of larger Criegee intermediates inside the fullerene cages.
Considering that the C70 or larger fullerene cage encapsulation of the simplest Criegee intermediate CH2OO is favorable, we next examined the plausible gas-phase reaction between CH2OO and H2O inside C70, C84, and C120 cages. Despite our attempts to optimize the (CH2OO⋯H2O)@C70 and (CH2OO⋯H2O)@C84 complexes using various starting geometries, the global minimum for the caged CH2OO⋯H2O complex could not be characterized. Instead, all the starting geometries converged to the final product of the reaction, α-hydroxy methyl hydroperoxide (HMHP, (HO)CH2(OOH)). The formation of HMHP from CH2OO and H2O inside C70 and C84 cages has reaction energies of −61.3 and −82.6 kcal/mol, respectively. That was quite surprising, yet an interesting outcome considering the fact that the CH2OO⋯H2O → (HO)CH2(OOH) reaction in the gas-phase involves an appreciable barrier of ∼6.9 kcal/mol at the M06-2X/cc-pVTZ level of theory and has a reaction energy of −55.9 kcal/mol (Fig. 2). It is important to mention here that though the previously calculated data for the gas-phase CH2OO⋯H2O → (HO)CH2(OOH) reaction is available at relatively higher levels of theory, we herein have calculated the gas-phase reaction profile at the M06-2X/cc-pVTZ level of theory to allow comparison between the gas-phase and caged reactions at the same theoretical footing. For example, Anglada et al. have studied the same reaction at the CCSD(T)/aug-cc-pVTZ//B3LYP/6-311 + G(2df,2p) level of theory.41 Though the M06-2X/cc-pVTZ calculated energies of individual stationary points for the gas-phase CH2OO⋯H2O reaction are significantly different from those previously calculated at the CCSD(T)/aug-cc-pVTZ//B3LYP/6-311 + G(2df,2p) theoretical level, the M06-2X/cc-pVTZ calculated barrier of 6.9 for the gas-phase CH2OO⋯H2O reaction is in reasonable agreement with the CCSD(T)/aug-cc-pVTZ//B3LYP/6-311 + G(2df,2p) calculated value of 7.5 kcal/mol. On the other hand, the (CH2OO⋯H2O)@C120 complex could be characterized at the M06-2X/cc-pVDZ level (Fig. 3), suggesting that the facile formation of HMHP inside C70 and C84 cages is due to the confinement effects inside fullerene cages. Note that we have used a slightly smaller basis set of cc-pVDZ for the C120 system due to its relatively larger size as compared to C60, C70, and C84 systems.
The (CH2OO⋯H2O)@C120 complex is 18.8 kcal/mol more stable relative to CH2OO@C120 and H2O. The transition state lies 4.4 kcal/mol above the (CH2OO⋯H2O)@C120 complex. The formation of HMHP@C120 has a reaction energy of −63.0 kcal/mol relative to CH2OO@C120 and H2O and has a reaction energy of −84.0 kcal/mol relative to separated CH2OO, H2O, and C120, respectively. The HMHP@C120 complex is 24.3 kcal/mol more stable than the separated C120 and HMHP. It is interesting to compare the results of the Criegee hydration in the gas-phase with those inside the C120 cage. The CH2OO⋯H2O complex in the gas-phase is 13.6 kcal/mol more stable than separated CH2OO and H2O. Inside the C120 cage, the same complex is 39.8 kcal/mol more favorable than the starting precursors, which translates into a significant stabilization of 26.2 kcal/mol because of the nanoconfinement. The M06-2X/cc-pVDZ calculated barrier for the Criegee hydration in the gas-phase is 5.3 kcal/mol, which is 0.9 kcal/mol larger than that inside the C120 cage. Although the barrier lowering due to the fullerene-cage confinement is not dramatic, the encapsulation energy of CH2OO@C120 or (CH2OO⋯H2O)@C120 complex is significantly favorable, which should make this reaction occur rapidly under fullerene confinement. It is also important to point out that the product of Criegee hydration, HMHP, is better stabilized inside a larger cage, i.e., the HMHP@C70 complex is only 5.4 kcal/mol more stable than free HMHP, whereas the HMHP@C84 and the HMHP@C120 complexes are 26.7 and 24.3 kcal/mol, respectively, more stable than free HMHP.
B. Spatial distribution and dynamic properties of CH2OO and HMHP inside fullerene cages
In order to deeply understand the Criegee reactivity inside a fullerene cage (>C60), we performed BOMD simulations at 300 K. The BOMD simulations are a powerful computational approach that offers useful insights into the dynamic behavior of a molecular system at a given temperature and allows the study of a chemical reaction as a dynamic time-dependent process. Specifically, the BOMD simulations were performed to better understand the dynamic behavior of Criegee intermediates inside C70, C84, and C120 cages. The simulation results show that CH2OO inside these fullerene cages remain dynamically stable for the 20 ps duration (Movies S1–S3 of the supplementary material). Compared to the gas-phase, the key C—O and O—O bond distances of CH2OO inside the C70, C84, and C120 cages undergo minor elongations of 0.05-0.10 Å during the simulated time (Fig. 4). No major deformation in the structure of the fullerene cage occurs during simulations, suggesting that the fullerene cage provides a stable framework for reactivity.
The calculated radial distribution functions (RDFs) for the encapsulated CH2OO provide useful information about its spatial distributions inside C70, C84, and C120 cages. Figure 5 displays the RDFs(R) for CH2OO@C70, CH2OO@C84, and CH2OO@C120 from the BOMD simulations; R here is the distance from the center of mass of CH2OO to the carbons of a fullerene cage. The results show that the CH2OO RDFs inside C70 and C84 cages have relatively narrower distributions and the peaks for these RDFs are centered around 4.0 Å, which is close to the interior radius of these cages, whereas the CH2OO RDF inside the C120 cage has a relatively broad distribution with a peak around 5.0 Å. Overall, the most probable distribution of Criegee intermediates inside a fullerene cage resides at a distance of 4.0–5.0 Å from the nearest carbons of the fullerene cage. We next calculated the mean square displacements (MSDs) of CH2OO inside these cages at 300 K. The calculated MSDs are shown in Fig. 5. The results suggest that the Criegee intermediates inside C70 or C84 is less mobile, whereas it exhibits relatively higher mobility inside the C120 cage, which is consistent with the size of a fullerene cage.
We also examined the dynamics of HMHP inside C70, C84, and C120 fullerene cages. The simulation results show that HMHP inside these cages remain dynamically stable for the 20 ps duration (Movies S4–S6 of the supplementary material). Figure 6 displays the RDFs(R) for HMHP@C70, HMHP@C84, and HMHP@C120 from the BOMD simulations. The dynamics behavior of HMHP inside these cages is similar to that of Criegee intermediates, i.e., the RDFs for HMHP inside C70 and C84 cages have relatively narrower distributions and the peaks for these RDFs are centered around 3.7 and 4.3 Å, respectively, whereas the HMHP RDF inside the C120 cage has a broader distribution with a peak around 5.0 Å. We also calculated the MSDs of HMHP inside these cages at 300 K. The calculated MSDs are shown in Fig. 6. Just like Criegee intermediates, HMHP inside the C70 or C84 cage is less mobile, whereas it exhibits relatively higher mobility inside the C120 cage, which is consistent with the size of a fullerene cage. Interestingly, HMHP inside C70 and C84 cages is relatively less mobile than CH2OO. But inside the larger cage of C120, HMHP is relatively more mobile than CH2OO. This is because HMHP is larger than CH2OO, which makes it less mobile inside smaller cages of C70 and C84. The mobility of HMHP inside C120 is not adversely impacted because C120 is larger in size than C70 or C84. Overall, these new results may open up new avenues of experimental and theoretical research for trapping and harnessing the chemical reactivity inside fullerene cages.
IV. SUMMARY
In conclusion, our density functional theory calculations show that the hydration of CH2OO within a fullerene cage (>C60) occurs favorably, suggesting the role of space confinement in chemical reactivity. The Born-Oppenheimer molecular dynamics simulations indicate that CH2OO and the product of its hydration, α-hydroxy methyl hydroperoxide, remain dynamically stable for the 20 ps time scale at 300 K and their self-diffusion coefficients inside fullerene cages correlate with the size of a cage. These results reveal that the confinement of the fullerene cage provides a unique environment for executing bimolecular chemistries. The conclusions from this study should serve as useful and new leads for studying other bimolecular chemical processes inside fullerene cages.
SUPPLEMENTARY MATERIAL
See supplementary material for the calculated equilibrium geometries and interaction energies of anti-CH3CHOO and syn-CH3CHOO encapsulated C70 fullerene, calculated structural parameters of the uncaged and caged Criegee intermediates, Cartesian coordinates of key species optimized at M06-2X/cc-pVTZ and M06-2X/cc-pVDZ levels of theory, and trajectories of the BOMD simulations for the simplest Criegee intermediate as well as for α-hydroxy methyl hydroperoxide inside C70, C84, and C120, respectively.
ACKNOWLEDGMENTS
This work is supported by the National Science Foundation (Grant No. CHE-1500217) and by the University of Nebraska Holland Computing Center.