Chemical Reaction CO+OH - American Chemical Society

Jun 28, 2016 - NanoScienece Technology Center,. ‡ ... Florida Solar Energy Center, ... Research Parkway, Ste 400, Orlando, Florida 32826, United Sta...
2 downloads 0 Views 1MB Size
Article pubs.acs.org/JPCA

Chemical Reaction CO+OH• → CO2+H• Autocatalyzed by Carbon Dioxide: Quantum Chemical Study of the Potential Energy Surfaces Artem ̈ E. Masunov,*,†,‡,§,∥,⊥ Elizabeth Wait,†,‡ and Subith S. Vasu# †

NanoScienece Technology Center, ‡Department of Chemistry, §Department of Physics, and ∥Florida Solar Energy Center, University of Central Florida, 12424 Research Parkway, Ste 400, Orlando, Florida 32826, United States ⊥ National Research Nuclear University MEPhI, Kashirskoye shosse 31, Moscow, 115409, Russia # Center for Advanced Turbomachinery and Energy Research (CATER), Mechanical and Aerospace Engineering University of Central Florida, Orlando, Florida 32816, United States ABSTRACT: The supercritical carbon dioxide medium, used to increase efficiency in oxy combustion fossil energy technology, may drastically alter both rates and mechanisms of chemical reactions. Here we investigate potential energy surface of the second most important combustion reaction with quantum chemistry methods. Two types of effects are reported: formation of the covalent intermediates and formation of van der Waals complexes by spectator CO2 molecule. While spectator molecule alter the activation barrier only slightly, the covalent bonding opens a new reaction pathway. The mechanism includes sequential covalent binding of CO2 to OH radical and CO molecule, hydrogen transfer from oxygen to carbon atoms, and CH bond dissociation. This reduces the activation barrier by 11 kcal/mol at the rate-determining step and is expected to accelerate the reaction rate. The finding of predicted catalytic effect is expected to play an important role not only in combustion but also in a broad array of chemical processes taking place in supercritical CO2 medium. It may open a new venue for controlling reaction rates for chemical manufacturing.

1. INTRODUCTION Oxy−fuel combustion in supercritical carbon dioxide (sCO2) offers several advantages in fossil energy technology.1 It can significantly increase power plant efficiency2 and drastically reduce pollutants.3 In oxy−fuel combustion, the fuel is burned in pure oxygen (O2) and CO2 diluent instead of nitrogen from the air.4 This yields combustion products consisting mainly of CO2 and H2O, avoids pollution by nitrogen oxides, and opens the possibility for easy carbon capture and sequestration.5 To reduce wasting fuel or oxidant, oxy−fuel systems may operate close to stoichiometry, so that the combustion temperature is controlled by adjusting the amount of diluent.6 The increase in efficiency is determined by the near-liquid density of sCO2 working fluid before entering the turbine.7,8 However, the effects of supercritical CO2 on the combustion kinetics are not well understood. Computational studies can greatly assist in this understanding,9−12 considering instrumental limitations of kinetic experiments at high pressure.13 Here we focus on OH• + CO → H• + CO2 reaction, the second most essential reaction in combustion, which represents the main heat release step.14 Because of its importance in combustion, it had been studied extensively. Rate constants for OH + CO reaction were found to be nearly independent of temperature between 80 and 500 K, while increasing with temperature above 500 K.15 This strong non-Arrhenius © 2016 American Chemical Society

behavior has been attributed to the intricate minimum energy paths (MEP), which have near-isoenergetic barriers in the entrance (OH• + CO) and exit (H• + CO2) channels15,16 and long-lived intermediate species. Several ab initio potential energy surface (PES) studies17−19 report existence of two possible intermediates, HOCO and HCO2, which form in the OH• + CO reaction.17 The minimum energy pathway proceeds via formation of cis-HOCO intermediate (Figure 1). Another intermediate, HCO2, has a lower activation energy for hydrogen dissociation but a relatively high activation barrier for its formation from trans-HOCO via intramolecular isomerization. The previous computational studies were focused on chemical transformations in vacuum or inert diluent.15,16,18,19 The effects of pressure, temperature, and quantum-mechanical tunneling on the kinetics of the reaction were studied.17 To the best of our knowledge, the effects of supercritical CO2 on the rates of product formation had not been analyzed yet. However, it is widely known that the presence of the small molecules (such as water) may alter the reaction pathways20 or molecular conformations,21 and either accelerate22 or slow down23,24 the reactions both in the gas phase and in the condensed state Received: March 30, 2016 Revised: June 1, 2016 Published: June 28, 2016 6023

DOI: 10.1021/acs.jpca.6b03242 J. Phys. Chem. A 2016, 120, 6023−6028

Article

The Journal of Physical Chemistry A

(including main-group atomization energies, barrier heights, and noncovalent interaction energies). However, long-range dispersion is still missing in the M11 functional.32 For this reason, we used Grimme’s three-body dispersion correction (GD3)33 with S8 and SR6 parameters chosen to be S8 = 0.0 and SR6 = 1.619 (values, optimized for a similar M06-2X functional). The 6-311G** basis set34 was selected for compatibility with CBS-QB3 model chemistry.35 Since CBSQB3 method employs optimization at uncorrected DFT level, it is not expected to describe van der Waals (vdW) complexes well. Therefore, we replace B3LYP functional with the dispersion-corrected M11+GD3 method to obtain geometries and frequencies. The modified compound method is dubbed CBS-QM11 and is benchmarked below. The pathway search was made using the following protocol: (1) optimization of the product/reactant van der Waals complex; (2) relaxed potential energy scan along selected bond formation/breaking; (3) transition state (TS) optimization starting form the highest point on that scan; (4) frequency analysis for identification of that TS as first-order critical point; (5) an intrinsic reaction coordinate36 (IRC) search for two (forward and backward) minimum energy pathways (MEP) starting from the transition state; (6) optimization into the local minima to continue both MEPs found in step 5 and then a repeat of step 2 above.

Figure 1. Relative energies (kcal/mol) of the stationary points along the reaction pathways for the HO+CO → H+CO2 reaction. The lowest energy pathway is in green, the intermediate one is in blue, and the higher one is in red (adapted from ref 19).

environment. In this work, we investigate how PES for the OH + CO → H+CO2 reaction is altered in the presence of one CO2 molecule. We hypothesize that trans-HOCO to HCO 2 isomerization may proceed via an intermolecular mechanism with a lower activation barrier. We also investigate the possibility of a spectator CO2 molecule influencing the activation barriers by forming a weak complex with the HOCO intermediate.

3. RESULTS AND DISCUSSION We begin with validation of the M11 exchange-correlation functional30 and the CBS-QM11 model chemistry for the reactive system in question. The energies of the stationary points, predicted in this work at M11+GD3/6-311G** and CBS-QM11 theory level, are reported in Table 1. The highlevel ab initio values from the literature are also shown for comparison. One can see that our M11+GD3 results deviate from the most recent CCSD(T)-F12 ab initio data by less than 3 kcal/mol (less than 2 kcal/mol for CBS-QM11 results). The maximal deviation between two advanced ab initio theory level (CCSD(T) and MRCI//CAS) is somewhat larger (3.8 kcal/ mol). Therefore, the M11+GD3 theory level has sufficient accuracy for this reactive system, and it is used to summarize our results in Table 2, Figures 2 and 3 and in the following discussion.

2. COMPUTATIONAL DETAILS All calculations were performed with the Gaussian 2009 suite of programs.25 Geometry optimizations and normal-mode analysis were performed by using density functional theory (DFT). DFT was shown to be an invaluable tool to predict linear26 and nonlinear spectra,27,28 as well as hyperpolarizability.29 For this work, we selected the M11 exchange-correlation functional.30 This functional belongs to the class of meta-GGA hybrids, where a larger fraction of Hartree−Fock exchange31 ensures the accurate description of the activated complexes, while dependence on the kinetic energy density recovers an accurate description of the stable molecular configurations. Unlike older meta-GGA hybrids, M11 includes range separation, so that fraction of Hartree−Fock exchange varies from 42.8% at short interelectron distances to 100% at the asymptotic limit. M11 was optimized for across-the-board superior performance

Table 1. Relative Energies (kcal/mol) for the Critical Points on PES for the CO + OH• → CO2 + H• Reaction, Predicted at Different Theory Levelsa

a

stationary point

MRCI//CASb

CCSD(T)c

CCSD(T)-F12d

M11/6-311G**+GD3

CBS-QM11

trans-HOCO cis-HOCO OH···CO OH···OC tor-TS cis-TS2 trans-TS4 cis-TS1 trans-TS1 HCO2 H·CO2-TS3 H + CO2 OH + CO

0 1.92 26.98 28.39 9.36 30.67 38.42 29.85 25.38

0 1.86 27.91 28.98 9.33 31.54 37.96 32.72 29.14

0 1.77 27.33 28.37 9.30 32.00 38.37 32.71 29.01 15.95 20.88 6.97 29.59

0.00 1.22 29.49 30.06 9.58 32.08 38.28 34.84 30.63 18.75 20.15 7.13 31.57

0.00 2.25 28.08 28.87 9.41 31.60 37.74 33.85 29.86 17.87 19.86 6.19 30.27

See text for details. bReference 17. cReference 18 dReference 19. 6024

DOI: 10.1021/acs.jpca.6b03242 J. Phys. Chem. A 2016, 120, 6023−6028

Article

The Journal of Physical Chemistry A Table 2. Predicted Relative Energies (kcal/mol) of Stationary Points on Reaction Pathwaysa stationary points, Figure 1

OH···OC

cis-TS1

cis-HOCO

cis-TS2

H + CO2

ΔE

M11/6-311G**+GD3 M11/6-311G**+GD3+ZPE stationary points, Scheme 1

30.06 26.04 RC11

34.84 31.33 TS12

1.22 1.05 IN13

32.08 26.89 TS14

7.13 1.18 IN15

30.86 25.84 ΔE

M11/6-311G**+GD3 M11/6-311G**+GD3+ZPE stationary points, Scheme 1 (contd)

25.16 21.95

44.67 42.61 TS16

24.47 24.11 IN17

36.49 36.27 TS18

17.44 19.20 PC19

19.51 20.66 ΔE

14.27 13.31 IN23

16.82 12.23 TS24

5.07 0.15 PC25

17.03 13.17 ΔE

IN15

M11/6-311G**+GD3 M11/6-311G**+GD3+ZPE stationary points, Scheme 2

17.44 19.20 RC21

34.47 32.37 TS22

M11/6-311G**+GD3 M11/6-311G**+GD3+ZPE stationary points, Scheme 3

25.16 22.47 RC31

25.87 22.77 TS32

−4.98 −4.11 IN33

29.69 25.19 TS34

4.87 −0.20 PC35

34.67 29.30 ΔE

M11/6-311G**+GD3 M11/6-311G**+GD3+ZPE stationary points, Scheme 4

25.52 22.47 RC41

26.79 24.38 TS42

−2.95 −1.95 IN43

47.96 43.45 TS44

6.07 −3.41 PC45

50.91 45.40 ΔE

M11/6-311G**+GD3 M11/6-311G**+GD3+ZPE

28.50 25.04

29.44 26.81

−2.52 −2.08

47.75 43.35

4.86 0.03

50.27 45.43

a For the first two lines, the trans-HOCO intermediate is chosen as the reference point; for all other lines, the trans-HOCO + CO2 system is chosen as the reference point. Both potential energies and enthalpies at 0 K, including the zero-point vibrational energy (ZPE), are reported. Activation barrier heights (ΔE) for the rate-limiting step are shown in the last column.

Scheme 1. Structures of the Complexes, Intermediates, and Transition States, along the Pathway witha Covalently Bound CO2 Molecule

Figure 2. Relative energies (kcal/mol) of the reaction pathway shown on Scheme 1, with one covalently bound CO2 molecule (trans-HOCO + CO2 system is chosen as the reference point).

Figure 3. Relative energies (kcal/mol) of two of the reaction pathways shown shown on Scheme 2 (in black) and Scheme 3 (in red) with spectator CO2 molecule (the trans-HOCO + CO2 system is chosen as the reference point).

search from nonbonded complex IN17 and located two transition states: one for intermolecular transfer of H atom to O atom (TS16), and another one for C−H bond dissociation (TS18). While IRC search from TS18 resulted in products vdW complex PC19, the MEP starting from TS16 lead to the polymeric intermediate IN15. Polymerization of carbon dioxide was studied both experimentally and theoretically under high pressure condition.37,38 It is conceivable that IN15 or similar species may play a role in this polymerization. Next, we investigated several bond dissociation pathways for IN15 and

Next we investigate a possibility that CO2 molecules from the environment will bind covalently and alter the reaction mechanism. Specifically, the H atom transfer from O to C atom in trans-HOCO intermediate may proceed via intermolecular mechanism, thus relieving the steric constraints of 3membered ring in transition state of 1,2-hydrogen shift (transTS4 in Figure 1). Indeed, we discovered the intermolecular mechanism, and it is reported in Scheme 1. We started our 6025

DOI: 10.1021/acs.jpca.6b03242 J. Phys. Chem. A 2016, 120, 6023−6028

Article

The Journal of Physical Chemistry A found CO dissociation via TS14 has the lowest energy. Resulting vdW complex IN13 finally dissociates into the reactants vdW complex RC11 via transition state TS12. The energies of the respective species are reported in Figure 2, and Table 2 (which also lists activation barrier heights). As one can see, the activation barrier for the mechanism shown on Scheme 1 is lower than the one for the mechanism with no additional CO2 molecules (by 11 kcal/mol before and by 5 kcal/mol after zero-point vibrational energy correction). Here we assume collisional thermalization for all the intermediates. It is a reasonable assumption at high pressure limit typical for supercritical fluid. We also investigate the structures of the nonbonded complexes between HOCO intermediate and spectator CO2 molecule. The spectator CO2 molecule forms relatively strong (3−6 kcal/mol) van der Waals complexes with both reactants and intermediates. The geometry of these complexes is determined not only by dispersion interaction, but also by stronger donor−acceptor contacts.39−41 While oxygen atoms may play the role of lone pair donors, both hydrogen and carbon atoms are electron deficient and are able to play the role of acceptors. Therefore, cis-HOCO may form two contacts of this nature (IN23, Scheme 2), while trans-HOCO may form

Scheme 3. Structures of the Complexes, Intermediates, and Transition States along the trans-HOCO Bond Dissociation Pathway with a Spectator CO2 Molecule

Scheme 4. Structures of the Complexes, Intermediates, and Transition States along the trans-HOCO Bond Dissociation Pathway with a Spectator CO2 Molecule

Scheme 2. Structures of the Complexes, Intermediates, and Transition States along the cis-HOCO Bond Dissociation Pathway with a Spectator CO2 Molecule

complexes where trans-HOCO donates electron pair (IN33 and IN43), the OH bond becomes stronger and activation barrier is nearly doubled (50.91 and 50.27 kcal/mol). Since MEP data in Scheme 3 and 4 are quantitatively similar, only Schemes 2 and 3 are illustrated in Figure 3. Finally, in complex IN55, the trans-HOCO intermediate forms a hydrogen bond to an additional CO2 molecule. Breaking the OH covalent bond in this intermediate is accompanied by formation of another covalent bond instead of dissociation. Since this mechanism does not result in formation of products, we do not report associated potential energies. However, the transition complex in the hydrogen transfer process includes the formation of a partially covalent C··C interaction, and this may play role in diffusion of hydrogen atom through the dense CO2 medium.

only one donor−acceptor contact. In the latter case, the hydroxyl oxygen atom (IN33, Scheme 3), or the terminal oxygen atom (IN43, Scheme 4) donates its lone pair. One contact also forms when the hydrogen of trans-HOCO acts as the lone pair acceptor (IN55, Scheme 5). We optimized the three complexes described (IN23, IN33, and IN43) and located the transition states, corresponding to the O−H and OC−OH bond breaking in each of them. The relative energies of the stationary points are collected in Table 2. The pathway without CO2 is also shown for comparison. One can see that in all four mechanisms the rate limiting step is dissociation of OH bond. The activation barrier for this step for the mechanism described on Scheme 2 (34.67 kcal/mol) is comparable with the activation barrier in the absence of additional CO2 molecule (33.30 kcal/mol). In case of

4. CONCLUSIONS We investigated the potential surfaces for the CO + OH• → CO2 + H• reaction in the presence of additional CO2 molecule. Depending on the complex structure, the activation barrier for covalent OH bond dissociation is either comparable or strongly (17−20 kcal/mol) increased. We also found a new reaction 6026

DOI: 10.1021/acs.jpca.6b03242 J. Phys. Chem. A 2016, 120, 6023−6028

Article

The Journal of Physical Chemistry A

Combustion and Implementation in CFD. Proc. Supercritical CO2 Power Cycles Symp. 2016, Paper No. 7. (2) Kotowicz, J.; Michalski, S. Efficiency Analysis of a Hard-CoalFired Supercritical Power Plant with a Four-End High-Temperature Membrane for Air Separation. Energy 2014, 64, 109−119. (3) Zhuang, Y.; Pavlish, J. H. Fate of Hazardous Air Pollutants in Oxygen-Fired Coal Combustion with Different Flue Gas Recycling. Environ. Sci. Technol. 2012, 46, 4657−4665. (4) Andersson, K.; Johnsson, F. Flame and Radiation Characteristics of Gas-Fired O2/CO2 Combustion. Fuel 2007, 86, 656−668. (5) Tan, Y.; Douglas, M. A.; Thambimuthu, K. V. CO2 Capture Using Oxygen Enhanced Combustion Strategies for Natural Gas Power Plants. Fuel 2002, 81, 1007−1016. (6) Bongartz, D.; Ghoniem, A. F. Chemical Kinetics Mechanism for Oxy-Fuel Combustion of Mixtures of Hydrogen Sulfide and Methane. Combust. Flame 2015, 162, 544−553. (7) Utamura, M.; Hasuike, H.; Yamamoto, T. Demonstration Test Plant of Closed Cycle Gas Turbine with Supercritical CO2 as Working Fluid. Strojarstvo 2010, 52, 459−465. (8) Zhang, X. R.; Yamaguchi, H.; Fujima, K.; Enomoto, M.; Sawada, N. Theoretical Analysis of a Thermodynamic Cycle for Power and Heat Production Using Supercritical Carbon Dioxide. Energy 2007, 32, 591−599. (9) Gadzhiev, O. B.; Ignatov, S. K.; Razuvaev, A. G.; Masunov, A. E. Quantum Chemical Study of Trimolecular Reaction Mechanism between Nitric Oxide and Oxygen in the Gas Phase. J. Phys. Chem. A 2009, 113, 9092−9101. (10) Gadzhiev, O. B.; Ignatov, S. K.; Gangopadhyay, S.; Masunov, A. E.; Petrov, A. I. Mechanism of Nitric Oxide Oxidation Reaction (2NO +O2 -> 2NO2) Revisited. J. Chem. Theory Comput. 2011, 7, 2021− 2024. (11) Gadzhiev, O. B.; Ignatov, S. K.; Krisyuk, B. E.; Maiorov, A. V.; Gangopadhyay, S.; Masunov, A. E. Quantum Chemical Study of the Initial Step of Ozone Addition to the Double Bond of Ethylene. J. Phys. Chem. A 2012, 116, 10420−10434. (12) Ignatov, S. K.; Gadzhiev, O. B.; Razuvaev, A. G.; Masunov, A. E.; Schrems, O. Adsorption of Glyoxal (CHOCHO) and Its UV Photolysis Products on the Surface of Atmospheric Ice Nanoparticles. Dft and Density Functional Tight-Binding Study. J. Phys. Chem. C 2014, 118, 7398−7413. (13) Koroglu, B.; Pryor, O.; Lopez, J.; Nash, L.; Vasu, S. S. Shock Tube Ignition Delay Times and Methane Time-Histories Measurements During Excess CO2 Diluted Oxy-Methane Combustion. Combust. Flame 2016, 164, 152−163. (14) Miller, J. A.; Kee, R. J.; Westbrook, C. K. Chemical-Kinetics and Combustion Modeling. Annu. Rev. Phys. Chem. 1990, 41, 345−387. (15) Liu, S.; Chen, J.; Fu, B. N.; Zhang, D. H. State-to-State Quantum Versus Classical Dynamics Study of the OH + CO → H + CO2 Reaction in Full Dimensions (J = 0): Checking the Validity of the Quasi-Classical Trajectory Method. Theor. Chem. Acc. 2014, 133, 1558. (16) Li, J.; Chen, J.; Zhang, D. H.; Guo, H. Quantum and QuasiClassical Dynamics of the OH + CO → H + CO2 Reaction on a New Permutationally Invariant Neural Network Potential Energy Surface. J. Chem. Phys. 2014, 140, 044327. (17) Zhu, R. S.; Diau, E. G. W.; Lin, M. C.; Mebel, A. M. A Computational Study of the OH (OD) + CO Reactions: Effects of Pressure, Temperature, and Quantum-Mechanical Tunneling on Product Formation. J. Phys. Chem. A 2001, 105, 11249−11259. (18) Yu, H. G.; Muckerman, J. T.; Sears, T. J. A Theoretical Study of the Potential Energy Surface for the Reaction OH+CO -> H+CO2. Chem. Phys. Lett. 2001, 349, 547−554. (19) Li, J.; Wang, Y. M.; Jiang, B.; Ma, J. Y.; Dawes, R.; Xie, D. Q.; Bowman, J. M.; Guo, H. Communication: A Chemically Accurate Global Potential Energy Surface for the HO+CO -> H+ CO2 Reaction. J. Chem. Phys. 2012, 136, 041103. (20) Viegas, L. P.; Varandas, A. J. C. Can Water Be a Catalyst on the HO2 + H2O + O3 Reactive Cluster? Chem. Phys. 2012, 399, 17−22. (21) Masunov, A.; Tretiak, S.; Hong, J. W.; Liu, B.; Bazan, G. C. Theoretical Study of the Effects of Solvent Environment on

Scheme 5. Structures of the Complexes, Intermediates, and Transition States along the trans-HOCO Hydrogen Exchange Pathway with Another CO2 Molecule

mechanism, where the additional CO2 molecule covalently binds the OH radical and then to the CO molecule, forming a new intermediate. This intermediate undergoes hydrogen transfer from oxygen to carbon atoms with subsequent CH bond dissociation. All activation barriers along the reaction pathway are less than 20 kcal/mol, considerably lower than activation barrier in the absence of additional CO2 molecule (31 kcal/mol). Hence, the product of CO + OH• → CO2 + H• reaction catalyzes the overall process (which is known as the autocatalytic effect). The new reaction mechanism with predicted autocatalytic effect is expected to play an important role not only in combustion, but also in broad array of chemical processes in supercritical CO2 medium. It may open a new venue for controlling reaction rates for chemical manufacturing. The experimental verification of this discovery is presently under way.



AUTHOR INFORMATION

Corresponding Author

*(A.E.M.) Telephone: 1-407-374-3783. E-mail: amasunov@ucf. edu. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported in part by the Department of Energy (Grant Number DE-FE0025260). The authors acknowledge the National Energy Research Scientific Computing Center (NERSC), and the University of Central Florida Advanced Research Computing Center (https://arcc.ist.ucf.edu) for providing computational resources and support. A.E.M. is grateful to the Russian Science Foundation, Contract No. 1443-00052, administered by Center of Photochemistry, Russian Academy of Science.



REFERENCES

(1) Vasu, S. S.; Pryor, O. M.; Kapat, J. S.; Masunov, A.; Martin, S. M. Developing a Validated Chemical Kinetics Model for sCO 2 6027

DOI: 10.1021/acs.jpca.6b03242 J. Phys. Chem. A 2016, 120, 6023−6028

Article

The Journal of Physical Chemistry A Photophysical Properties and Electronic Structure of Paracyclophane Chromophores. J. Chem. Phys. 2005, 122, 224505. (22) Kircher, C. C.; Sander, S. P. Kinetics and Mechanism of HO2 and DO2 Disproportionations. J. Phys. Chem. 1984, 88, 2082−2091. (23) Zakharov, M.; Masunov, A. E.; Dreuw, A. Water Deficient Environment Accelerates Proton Exchange: Acetone-Water Reaction Catalyzed by Calix 4 Hydroquinone Nanotubes. J. Phys. Chem. C 2009, 113, 10395−10401. (24) Zakharov, A.; Masunov, A. E.; Dreuw, A. Catalytic Role of Calix 4 Hydroquinone in Acetone-Water Proton Exchange: A Quantum Chemical Study of Proton Transfer via Keto-Enol Tautomerism. J. Phys. Chem. A 2008, 112, 10405−10412. (25) Frisch, M. J. et al. Gaussian 09, Rev. D.01; Gaussian, Inc.: Wallingford CT, 2009. (26) De Boni, L.; Toro, C.; Masunov, A. E.; Hernandez, F. E. Untangling the Excited States of DR1 in Solution: An Experimental and Theoretical Study. J. Phys. Chem. A 2008, 112, 3886−3890. (27) Belfield, K. D.; Bondar, M. V.; Frazer, A.; Morales, A. R.; Kachkovsky, O. D.; Mikhailov, I. A.; Masunov, A. E.; Przhonska, O. V. Fluorene-Based Metal-Ion Sensing Probe with High Sensitivity to Zn2+ and Efficient Two-Photon Absorption. J. Phys. Chem. B 2010, 114, 9313−9321. (28) Belfield, K. D.; Bondar, M. V.; Hernandez, F. E.; Masunov, A. E.; Mikhailov, I. A.; Morales, A. R.; Przhonska, O. V.; Yao, S. TwoPhoton Absorption Properties of New Fluorene-Based Singlet Oxygen Photosensitizers. J. Phys. Chem. C 2009, 113, 4706−4711. (29) Suponitsky, K. Y.; Masunov, A. E. Supramolecular Step in Design of Nonlinear Optical Materials: Effect of pi···pi Stacking Aggregation on Hyperpolarizability. J. Chem. Phys. 2013, 139, 094310. (30) Peverati, R.; Truhlar, D. G. Improving the Accuracy of Hybrid meta-GGA Density Functionals by Range Separation. J. Phys. Chem. Lett. 2011, 2, 2810−2817. (31) Mikhailov, I. A.; Bondar, M. V.; Belfield, K. D.; Masunov, A. E. Electronic Properties of a New Two-Photon Absorbing Fluorene Derivative: The Role of Hartree-Fock Exchange in the Density Functional Theory Design of Improved Nonlinear Chromophores. J. Phys. Chem. C 2009, 113, 20719−20724. (32) Goerigk, L. Treating London-Dispersion Effects with the Latest Minnesota Density Functionals: Problems and Possible Solutions. J. Phys. Chem. Lett. 2015, 6, 3891−3896. (33) McLean, A. D.; Chandler, G. S. Contracted Gaussian-Basis Sets for Molecular Calculations 0.1. 2nd Row Atoms, Z = 11−18. J. Chem. Phys. 1980, 72, 5639−5648. (34) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate ab initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. (35) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G. A. A Complete Basis Set Model Chemistry. VI. Use of Density Functional Geometries and Frequencies. J. Chem. Phys. 1999, 110, 2822−2827. (36) Gonzalez, C.; Schlegel, H. B. An Improved Algorithm for Reaction-Path Following. J. Chem. Phys. 1989, 90, 2154−2161. (37) Plasienka, D.; Martonak, R. Structural Evolution in HighPressure Amorphous CO2 from Ab Initio Molecular Dynamics. Phys. Rev. B: Condens. Matter Mater. Phys. 2014, 89, 134105. (38) Datchi, F.; Mallick, B.; Salamat, A.; Rousse, G. E.; Ninet, S.; Garbarino, G.; Bouvier, P.; Mezouar, M. Structure and Compressibility of the High-Pressure Molecular Phase II of Carbon Dioxide. Phys. Rev. B: Condens. Matter Mater. Phys. 2014, 89, 144101. (39) Vyboishchikov, S. F.; Masunov, A. E. Topological Properties of Electron-Density in the H−+H2CO Reaction System. J. Mol. Struct.: THEOCHEM 1994, 311, 161−167. (40) Masunov, A. E.; Zorkii, P. M. Donor-Acceptor Nature of Specific Nonbonded Interactions of Sulfur and Halogen Atoms Influence on the Geometry and Packing of Molecules. J. Struct. Chem. 1992, 33, 423−435.

(41) Masunov, A. E.; Zorkii, P. M. Geometric Characteristics of Halogen-Halogen Intermolecular Contacts in Organic-Crystals. Zh. Fiz. Khim. 1992, 66, 60−69.

6028

DOI: 10.1021/acs.jpca.6b03242 J. Phys. Chem. A 2016, 120, 6023−6028