Next Article in Journal
Partial Denaturation of Double-Stranded DNA on Pristine Graphene under Physiological-like Conditions
Previous Article in Journal
Investigation of the Impact of High Concentration LiTFSI Electrolytes on Silicon Anodes with Reactive Force Field Simulations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Conformational Dependence of the First Hyperpolarizability of the Li@B10H14 in Solution

Instituto de Física, Universidade Federal de Goiás, Goiânia 74690-900, GO, Brazil
*
Author to whom correspondence should be addressed.
Liquids 2023, 3(1), 159-167; https://doi.org/10.3390/liquids3010012
Submission received: 26 November 2022 / Revised: 13 January 2023 / Accepted: 15 February 2023 / Published: 20 February 2023

Abstract

:
Using the ASEC-FEG approach in combination with atomistic simulations, we performed geometry optimizations of a Cs conformer of the lithium decahydroborate (Li@B10H14) complex in chloroform and in water, which has been shown to be the most stable in the gas phase and calculated its first hyperpolarizability. At room temperature, ASEC-FEG calculations show that this conformer is stable only in chloroform. However, it is found that the nonlinear response of the Cs conformer in chloroform is mild, and the result for the hyperpolarizability is markedly decreased in comparison with the result of the C2v conformer.

Graphical Abstract

1. Introduction

Among the molecular systems with nonlinear optical (NLO) responses, electrides have excess electrons that are not located on any specific atom [1,2]. Several theoretical investigations have demonstrated that electride-like structures can be designed by doping an alkali metal atom into organic complexants and can exhibit extraordinary NLO properties relative to the corresponding undoped organic ones [3,4,5,6,7,8,9,10,11]. However, organic electrides may have limited practical applications due to their thermal instability at room temperature and reactivity with air [1,12].
Electrides with relatively good thermal stability and NLO properties have been designed by the addition of lithium atoms to inorganic borane clusters [13,14,15]. In particular, Muhammad et al. [13] showed, in the gas phase, that an inorganic electride obtained from adding a Li atom to the cavity of decaborane (B10H14) can have a remarkable first hyperpolarizability. In that study, the authors considered an arrangement in which an added Li atom interacts with terminal H atoms on the open face of the basket-shaped molecule (C2v symmetry). Based on this geometric structure, we carried out an investigation of the stability of this molecular shape of Li@B10H14 in chloroform and water. We showed, at room temperature, that the geometric changes induced in solution in the C2v conformer of Li@B10H14 result in a stable structure only in chloroform [16]. Recently, Medved et al. [17] have shown, in the gas phase, based on the global energy minimum analysis, that Cs conformers (Li atoms close to the decaborane basket from the side or the side-bottom) of Li@B10H14 can be significantly more stable than the C2v structure, with possible consequences on their NLO responses. Thus, from the most recent theoretical results, there is a possibility that Li@B10H14 conformers containing a loosely bound excess electron may be stable in the solution and present different NLO responses.
In the present study, we report the results on the stability and first hyperpolarizability of a Cs conformer of Li@B10H14 (Figure 1) in chloroform and water solvents, which has been shown to be the most stable in the gas phase [17]. This represents an extension of a previous calculation on the stability and first hyperpolarizability of the C2v conformer of Li@B10H14 (Figure 1) in solution [16]. The absorption spectrum is also presented, considering that the addition of an excess electron can affect the electronic excitation energies of this inorganic electride. The equilibrium structure of the Cs conformer in each medium, at temperature and pressure under ambient conditions, was obtained using the ASEC-FEG scheme [18,19,20,21,22,23,24,25] by employing iteratively the sequential quantum mechanics/molecular mechanics (S-QM/MM) method [26,27]. The S-QM/MM approach has been successfully applied to reliably describe the effects of solvent on the electronic properties of solute molecules [28,29,30,31,32].

2. Computational Details

We have used the ASEC-FEG methodology to optimize the geometry of the Li@B10H14 system in the solution. The methodology is described in detail in previous papers [22,23,24,25]. Briefly, it consists in applying iteratively the Sequential QM/MM methodology [26,27] combined with the Average Solvent Electrostatic Configuration (ASEC) mean field technique [21] and the Free Energy Gradient (FEG) method [18,19,20] to provide free energy gradients and hessians to feed a quasi-newton optimization procedure in the search for the nearest local minimum in the free energy hypersurface.
During the classical simulations, all solvent molecules were kept rigid, and the temperature and pressure were kept at ambient conditions. The molecular properties of the solution were obtained with the unrestricted second-order Møller–Plesset Perturbation Theory (UMP2) using the ASEC mean field [21], where the solvent molecules were treated as point charges. Following previous work [16], boron and hydrogen atoms were described with the cc-pVDZ basis set. For the lithium atom, the aug-cc-pVDZ basis set was used because it is more appropriate to describe the loosely bound valence electron. Thus, for the ASEC-FEG geometry optimizations of the Cs conformer of Li@B10H14, such basis sets were used, denoted as (Li-aug)-cc-pVDZ. For water, the potential used was the extended simple point charge (SPC/E) model [33], because it leads to very good agreement for the critical point of water [34]. For chloroform, we have used the OPLS force field [35]. The atomic charges were obtained by performing an electrostatic potential fit (Merz-Kollman) [36] to the MP2/aug-cc-pVDZ charge density.
Here we report the permanent dipole moment (μ) and the first hyperpolarizability properties that can be compared with experimental results of hyper-Rayleigh scattering (HRS) in the solution [37]. The static first hyperpolarizability components (βijk) were obtained as numerical differentiation of the energy calculated in the presence of different values of the applied electric field, as implemented in the Gaussian 16 package [38]. Both μ and βijk were calculated at the MP2/aug-cc-pVDZ level of theory. The choice of this basis set is supported by previous tests and ensures a good compromise between computational cost and accuracy [39,40,41].
In solution, the NLO quantities that can be compared with HRS measurements are the HRS first hyperpolarizability (βHRS) and the associated depolarization ratio (DR), given respectively by:
β HRS = β Z Z Z 2 + β Z X X 2
and
DR = β Z Z Z 2 β Z X X 2 ,
whose expressions for β Z Z Z 2 and β Z X X 2 are given in Ref. [42].
The vertical electronic excitation energies were calculated using the TD-DFT method with different exchange–correlation functionals based on the generalized gradient approximation (GGA): LC-BLYP [43], CAM-B3LYP [44], M05-2X [45], and M06-2X [46] functionals. The aug-cc-pVDZ basis has been adopted in all TD-DFT calculations. All QM calculations were performed using unrestricted methods, and the spin contamination was found to be very small, less than 2% from the exact expected value. The classical Monte Carlo simulations were performed with the DICE program [47] and the QM calculations with the Gaussian 16 program [38].

3. Results and Discussion

Table 1 shows the MP2/(Li-aug)-cc-pVDZ results for bond lengths between the Li atom and closest B and H atoms of the optimized geometry of Cs conformer of Li@B10H14 in the gas phase, chloroform, and water. The converged values of maximum and RMS forces are below 6.5 × 10−4 and 2.5 × 10−4 hartree/bohr, respectively, which are reasonable values considering fluctuations due to liquid dynamics [22]. For the Cs conformer in chloroform, there is a slight increase in the bond lengths B1–Li, B2–Li, and B3–Li relative to the gas phase results, but these are smaller than the corresponding closest distances of the C2v conformer [16]. This indicates that the Cs conformer of Li@B10H14 is stable at room temperature in chloroform. Differently, the electride-like structure of the Cs conformer is unstable in water at room temperature, and during the simulations, the lithium ion dissociates, as previously shown for the C2v conformer [16]. Nevertheless, we compute the properties of interest using the in-water geometry from the last step of the iterative process.
Having obtained the optimized geometry of the Cs conformer in chloroform, the relative stability of the Cs and C2v conformers is evaluated at the MP2/aug-cc-pVDZ level. Our energy calculations indicate that the Cs conformer in chloroform is more stable than the C2v conformer by an amount of 21 kcal/mol. This finding is in agreement with the previous study by Medved et al. [17], which showed that the Cs conformer is the most stable form of Li@B10H14 in the gas phase. It should be emphasized that the inclusion of the solute-solvent interaction contribution in the energy calculations of the conformers gives a marked decrease in the energy difference between the conformers, which is 35 kcal/mol in the gas phase. Following previous work [16], we have also calculated the vertical ionization potential (VIP) of the Cs conformer in the gas phase and solution at the MP2/aug-ccpVDZ level, used as an indicator of thermal stability. It is found that the formation of excess electrons is slightly more stable in chloroform (6.27 eV) than in the gas phase (6.06 eV). In chloroform, however, our previous result indicated that the excess electron is slightly more stable for the C2v conformer (6.66 eV) [16] than for the Cs one. In water, the VIP value of 6.05 eV for the Cs conformer also suggests relative stability for the excess electron on the decaborane cluster. For comparison, the VIP value for the Cs conformer in chloroform is significantly larger than some electrides formed by doping the lithium atom in organic compounds, such as Li@calyx[4]pyrrole (4.16 eV) [3], Li-[15]aneN5 (2.32 eV) [6], and (Li+@n6adz)Li (2.88 eV) [48].
Figure 2 illustrates the behavior of the dipole moment of the Cs conformer in chloroform solution during the iterative process, indicating a rapid convergence pattern. The MP2/aug-cc-pVDZ results for µ of the Cs conformer in the gas phase and solution are quoted in Table 2. The corresponding results for the C2v conformer are also reported in this table. For the Cs conformer in chloroform, partial negative charges are distributed on the boron and hydrogen atoms closest to the lithium atom, and the polarization effect induces a slight increase in q(Li), with the in-chloroform value of 0.81 e when compared with the gas phase result of 0.77 e. This leads to the in-chloroform result for µ of 2.91 a.u., which is increased by 15% when compared to the gas phase result of 2.53 a.u., but it is 37% smaller than the in-chloroform result for the C2v conformer of 4.59 a.u. [16]. In water, the excess charge of the Cs conformer is delocalized over the decaborane cluster. The in-water q(Li) charge is found to be 0.97 e, close to +1 e, which shows that the Li atom has been ionized to form a cation and the B10H14 anion, and the in-water μ value is increased by 59% relative to the gas phase value. In the gas phase, our MP2/aug-cc-pVDZ results for µ are in good agreement with the MP2/6-31+G(d)//MP2/aug-cc-pVDZ results of Medved et al. [17], which are also quoted in Table 2, whose differences do not reach 4%.
MP2/aug-cc-pVDZ results for the static HRS first hyperpolarizability and depolarization ratio of the Cs conformer obtained in the gas phase and solution are gathered in Table 3. The corresponding results for the C2v conformer are also reported in this table. One can see that the polarization effect of the excess electron for βHRS of the Li@B10H14 complex is particularly affected by the conformational structure, as shown by Medved et al. [17]. In the gas phase, the result for βHRS of the Cs conformer (504 a.u.) is markedly reduced by 92% when compared with that of the C2v conformer (6214 a.u.). As shown in previous work [16], however, the electride-like characteristic gives a very large hyperpolarizability for the C2v conformer.
Our MP2/aug-cc-pVDZ results for βHRS of the Cs conformer are very sensitive to the solvent effects, with significant reductions in chloroform and in water. For this conformer in chloroform, the result of 212 a.u. for βHRS has a significant decrease (58%) relative to the gas phase result of 504 a.u., indicating that the contribution of the excess electron for βHRS is mild. In water, the MP2 model predicts for βHRS of the Cs conformer a result of 154 a.u. which is 27% smaller than the result in chloroform. Despite the structure of the Cs conformer being unstable in water, the first hyperpolarizability is not significantly affected in relation to the value in chloroform. Notice that solvent effects significantly impact the depolarization ratio (Table 3) of the Cs conformer in water, and the DR value of 7.23 indicates a marked dipolar contribution to βHRS. In chloroform, however, there is a less pronounced dipolar contribution to βHRS, whose DR value of 4.02 for the Cs conformer is 38% smaller than the corresponding value of the C2v conformer.
A qualitative interpretation of the solvent dependence of the hyperpolarizability of the Cs conformer can be obtained from the two-level model [49], where βHRS is proportional to the difference between the crucial excited state permanent dipole moment and the ground state permanent dipole moment (Δμ) and the oscillator strength (f0) and inversely proportional to the third power of the transition energy (ΔE):
β HRS Δ μ f 0 / Δ E 3
In this model, the transition energy plays an important role in determining the first hyperpolarizability. For the Cs conformer, the first absorption band is dominated by a highest occupied orbital molecular (HOMO) → lowest unoccupied orbital molecular (LUMO) excitation. However, TD-DFT results indicate that the first transition has a very small oscillator strength, regardless of the environment.
Figure 3 shows the absorption spectra of the Cs conformer in the gas phase and the solution obtained at the CAM-B3LYP/aug-cc-pVDZ level. The corresponding spectra for the C2v conformer are also reported in this figure. TD-DFT/aug-cc-pVDZ results for the excitation energy and oscillator strength of the more intense electronic transition of the Cs conformer are quoted in Table 4. The corresponding results for the first electronic transition C2v conformer are also reported in this table. One can see that the absorption spectra of the Cs conformer, not only in chloroform but also in water, are characterized by more intense transitions with energies larger than 4 eV. Thus, there is a considerable difference between the crucial electronic transition energies for Cs and C2v conformers in both chloroform and water by a factor around 2, and, therefore, the main factor for the marked decrease of βHRS of the Cs conformer is related to the significant increases of the crucial transition energy (Equation (3)). In addition, similar energy values for the more intense transition in chloroform and water, suggest that the βHRS of the Cs conformer does not change significantly in any environment.

4. Conclusions

We have more recently shown that the C2v conformer of the Li@B10H14 complex, which had its NLO properties studied in the gas phase by Muhammad et al., is a stable structure in chloroform at room temperature. However, Medved et al. have recently shown that Cs conformers of the Li@B10H14 complex in the gas phase can be significantly more stable than the C2v conformer, with a marked effect on the second-order NLO properties. In this work, we report an investigation on the stability and first hyperpolarizability of a Cs conformer in aprotic and protic solvents using the sequential QM/MM methodology in combination with the ASEC-FEG model, which has been shown to be the most stable in the gas phase. ASEC-FEG calculations show that the solvent-induced geometric changes in the Cs conformer at room temperature result in a stable structure in chloroform and an unstable structure in water. Our MP2 results show that the Cs conformer in chloroform is more stable than the C2v conformer by an amount of 21 kcal/mol, which confirms the finding of Medved et al. However, the relative stability of the conformers is significantly reduced due to the solute-solvent interaction energy when compared with the gas phase result, which is 35 kcal/mol. Polarization effects due to the electrostatic embedding lead to a reduction of 58% of the hyperpolarizability of the Cs conformer in chloroform in comparison with the gas result of 504 a.u. In chloroform, the contribution of the excess electron to the first hyperpolarizability of the Cs conformer is mild, and the result for βHRS of 212 a.u. is markedly decreased by a factor of 92% in comparison with the corresponding result for the C2v conformer of 2786 a.u. Therefore, the small first hyperpolarizability of the Cs conformer suggests that this inorganic electride-type system should not present significant second-order nonlinear responses in any environment. This finding also illustrates that the influence of the environment must be carefully considered in the design of new molecules with excess electrons that are stable at room temperature for aiming applications in NLO.

Author Contributions

I.B.: data curation, formal analysis, writing—review and editing. T.L.F.: conceptualization, formal analysis, writing—original draft, writing—review and editing. H.C.G.: data curation, formal analysis, writing—review and editing. M.A.C.: formal analysis, writing—review and editing. R.B.P.: formal analysis, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article and bibliography.

Acknowledgments

The authors gratefully acknowledge the financial support of CNPq, CAPES, and FAPEG (PRONEX) agencies (Brazil) as well as the computer resources of the LaMCAD/UFG laboratory.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dye, J.L. Electrons as Anions. Science 2003, 301, 607–608. [Google Scholar] [CrossRef]
  2. Dye, J.L. Electrides: Early Examples of Quantum Confinement. Acc. Chem. Res. 2009, 42, 1564–1572. [Google Scholar] [CrossRef]
  3. Chen, W.; Li, Z.R.; Wu, D.; Li, Y.; Sun, C.C.; Gu, F.L. The Structure and the Large Nonlinear Optical Properties of Li@calix[4]Pyrrole. J. Am. Chem. Soc. 2005, 127, 10977–10981. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, W.; Li, Z.R.; Wu, D.; Li, R.Y.; Sun, C.C. Theoretical Investigation of the Large Nonlinear Optical Properties of (HCN)n Clusters with Li Atom. J. Phys. Chem. B 2005, 109, 601–608. [Google Scholar] [CrossRef] [PubMed]
  5. Xu, H.L.; Li, Z.R.; Wu, D.; Wang, B.Q.; Li, Y.; Gu, F.L.; Aoki, Y. Structures and Large NLO Responses of New Electrides: Li-Doped Fluorocarbon Chain. J. Am. Chem. Soc. 2007, 129, 2967–2970. [Google Scholar] [CrossRef] [PubMed]
  6. Li, Z.J.; Li, Z.R.; Wang, F.F.; Luo, C.; Ma, F.; Wu, D.; Wang, Q.; Huang, X.R. A Dependence on the Petal Number of the Static and Dynamic First Hyperpolarizability for Electride Molecules: Many-Petal-Shaped Li-Doped Cyclic Polyamines. J. Phys. Chem. A 2009, 113, 2961–2966. [Google Scholar] [CrossRef] [PubMed]
  7. Xu, H.-L.; Li, Z.R.; Wu, D.; Ma, F.; Li, Z.J.; Gu, F.L. Lithiation and Li-Doped Effects of [5]Cyclacene on the Static First Hyperpolarizability. J. Phys. Chem. C 2009, 113, 4984–4986. [Google Scholar] [CrossRef]
  8. Liu, Z.B.; Zhou, Z.J.; Li, Y.; Li, Z.R.; Wang, R.; Li, Q.Z.; Li, Y.; Jia, F.Y.; Wang, Y.F.; Li, Z.J.; et al. Push-Pull Electron Effects of the Complexant in a Li Atom Doped Molecule with Electride Character: A New Strategy to Enhance the First Hyperpolarizability. Phys. Chem. Chem. Phys. 2010, 12, 10562–10568. [Google Scholar] [CrossRef]
  9. Xu, H.L.; Sun, S.L.; Muhammad, S.; Su, Z.M. Three-Propeller-Blade-Shaped Electride: Remarkable Alkali-Metal-Doped Effect on the First Hyperpolarizability. Theor. Chem. Acc. 2011, 128, 241–248. [Google Scholar] [CrossRef]
  10. Zhong, R.L.; Xu, H.L.; Li, Z.R.; Su, Z.M. Role of Excess Electrons in Nonlinear Optical Response. J. Phys. Chem. Lett. 2015, 6, 612–619. [Google Scholar] [CrossRef]
  11. Oliveira, I.M.; Castro, M.A.; Leão, S.A.; Fonseca, T.L.; Pontes, R.B. Li4C4H2N2: A Molecule with Large Hyperpolarizabilities and Electride Characteristic. Int. J. Quantum Chem. 2018, 118, e25661. [Google Scholar] [CrossRef]
  12. Matsuishi, S.; Toda, Y.; Miyakawa, M.; Hayashi, K.; Kamiya, T.; Hirano, M.; Tanaka, I.; Hosono, H. High-Density Electron Anions in a Nanoporous Single Crystal: [Ca24Al28O64]4+ (4e). Science 2003, 301, 626–629. [Google Scholar] [CrossRef] [PubMed]
  13. Muhammad, S.; Xu, H.; Liao, Y.; Kan, Y.; Su, Z. Quantum Mechanical Design and Structure of the Li@B10H14 Basket with a Remarkably Enhanced Electro-Optical Response. J. Am. Chem. Soc. 2009, 131, 11833–11840. [Google Scholar] [CrossRef]
  14. Muhammad, S.; Xu, H.; Su, Z. Capturing a Synergistic Effect of a Conical Push and an Inward Pull in Fluoro Derivatives of Li@B10H14 Basket: Toward a Higher Vertical Ionization Potential and Nonlinear Optical Response. J. Phys. Chem. A 2011, 115, 923–931. [Google Scholar] [CrossRef] [PubMed]
  15. Ma, N.; Gong, J.; Li, S.; Zhang, J.; Qiu, Y.; Zhang, G. Second-Order NLO Responses of Two-Cavity Inorganic Electrides Li: N@B20H26 (n = 1, 2): Evolutions with Increasing Excess Electron Number and Various B-B Connection Sites of B20H26. Phys. Chem. Chem. Phys. 2017, 19, 2557–2566. [Google Scholar] [CrossRef] [PubMed]
  16. Brandão, I.; Fonseca, T.L.; Georg, H.C.; Castro, M.A.; Pontes, R.B. Assessing the Structure and First Hyperpolarizability of Li@B10H14 in Solution: A Sequential QM/MM Study Using the ASEC-FEG Method. Phys. Chem. Chem. Phys. 2020, 22, 17314–17324. [Google Scholar] [CrossRef]
  17. Medved, M.; Demissie, T.B.; McKee, M.L.; Hnyk, D. The Behavior of a Paramagnetic System in Electric and Magnetic Fields as Exemplified by Revisiting Li@B10H14. Phys. Chem. Chem. Phys. 2017, 19, 12229–12236. [Google Scholar] [CrossRef]
  18. Okuyama-Yoshida, N.; Nagaoka, M.; Yamabe, T. Transition-State Optimization on Free Energy Surface: Toward Solution Chemical Reaction Ergodography. Int. J. Quantum Chem. 1998, 70, 95–103. [Google Scholar] [CrossRef]
  19. Okuyama-Yoshida, N.; Kataoka, K.; Nagaoka, M.; Yamabe, T. Structure Optimization via Free Energy Gradient Method: Application to Glycine Zwitterion in Aqueous Solution. J. Chem. Phys. 2000, 113, 3519–3524. [Google Scholar] [CrossRef]
  20. Hirao, H.; Nagae, Y.; Nagaoka, M. Transition-State Optimization by the Free Energy Gradient Method: Application to Aqueous-Phase Menshutkin Reaction between Ammonia and Methyl Chloride. Chem. Phys. Lett. 2001, 348, 350–356. [Google Scholar] [CrossRef]
  21. Coutinho, K.; Georg, H.C.; Fonseca, T.L.; Ludwig, V.; Canuto, S. An Efficient Statistically Converged Average Configuration for Solvent Effects. Chem. Phys. Lett. 2007, 437, 148–152. [Google Scholar] [CrossRef]
  22. Georg, H.C.; Canuto, S. Electronic Properties of Water in Liquid Environment: A Sequential QM/MM Study Using the Free Energy Gradient Method. J. Phys. Chem. B 2012, 116, 11247–11254. [Google Scholar] [CrossRef] [PubMed]
  23. Franco, L.R.; Brandão, I.; Fonseca, T.L.; Georg, H.C. Elucidating the Structure of Merocyanine Dyes with the ASEC-FEG Method. Phenol Blue in Solution. J. Chem. Phys. 2016, 145, 194301. [Google Scholar] [CrossRef] [PubMed]
  24. Brandão, I.; Franco, L.R.; Fonseca, T.L.; Castro, M.A.; Georg, H.C. Confirming the Relationship between First Hyperpolarizability and the Bond Length Alternation Coordinate for Merocyanine Dyes. J. Chem. Phys. 2017, 146, 224505. [Google Scholar] [CrossRef] [PubMed]
  25. Valverde, D.; Georg, H.C.; Canuto, S. Free-Energy Landscape of the SN2 Reaction CH3Br + Cl –> CH3Cl + Br- in Different Liquid Environments. J. Phys. Chem. B 2022, 126, 3685–3692. [Google Scholar] [CrossRef] [PubMed]
  26. Coutinho, K.; Canuto, S. Solvent Effects from a Sequential Monte Carlo-Quantum Mechanical Approach. Adv. Quantum Chem. 1997, 28, 89–105. [Google Scholar] [CrossRef]
  27. Coutinho, K.; Canuto, S. Solvent Effects in Emission Spectroscopy: A Monte Carlo Quantum Mechanics Study of the n←π* Shift of Formaldehyde in Water. J. Chem. Phys. 2000, 113, 9132–9139. [Google Scholar] [CrossRef]
  28. Fonseca, T.L.; Coutinho, K.; Canuto, S. Probing Supercritical Water with the n-π* Transition of Acetone: A Monte Carlo/Quantum Mechanics Study. J. Chem. Phys. 2007, 126, 034508. [Google Scholar] [CrossRef]
  29. Fonseca, T.L.; Georg, H.C.; Coutinho, K.; Canuto, S. Polarization and Spectral Shift of Benzophenone in Supercritical Water. J. Phys. Chem. A 2009, 113, 5112–5118. [Google Scholar] [CrossRef]
  30. Fonseca, T.L.; Coutinho, K.; Canuto, S. Hydrogen Bond Interactions between Acetone and Supercritical Water. Phys. Chem. Chem. Phys. 2010, 12, 6660–6665. [Google Scholar] [CrossRef]
  31. Junior, L.A.; Colherinhas, G.; Fonseca, T.L.; Castro, M.A. Solvent Effects on the First Hyperpolarizability of Retinal Derivatives. Chem. Phys. Lett. 2014, 598, 43–47. [Google Scholar] [CrossRef]
  32. Oliveira, L.B.A.; Colherinhas, G.; Fonseca, T.L.; Castro, M.A. Spectroscopic Properties of Vitamin E Models in Solution. Chem. Phys. Lett. 2015, 628, 49–53. [Google Scholar] [CrossRef]
  33. Berendsen, H.J.C.; Grigera, J.R.; Straatsma, T.P. The missing term in effective pair potentials. J. Phys. Chem. 1987, 91, 6269–6271. [Google Scholar] [CrossRef]
  34. Guissani, Y.; Guillot, B. A computer simulation study of the liquid– vapor coexistence curve of water. J. Chem. Phys. 1993, 98, 8221–8235. [Google Scholar] [CrossRef]
  35. McDonald, N.A.; Carlson, H.A.; Jorgensen, W.L. Free energies of solvation in chloroform and water from a linear response approach. J. Phys. Org. Chem. 1997, 10, 563–576. [Google Scholar] [CrossRef]
  36. Besler, B.H.; Merz, K.M.; Kollman, P.A. Atomic charges derived from semiempirical methods. J. Comput. Chem. 1990, 11, 431–439. [Google Scholar] [CrossRef]
  37. Clays, K.; Persoons, A. Hyper-Rayleigh Scattering in Solution. Phys. Rev. Lett. 1991, 66, 2980–2983. [Google Scholar] [CrossRef] [PubMed]
  38. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B. GAUSSIAN 16, Revision B.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  39. Naves, E.S.; Castro, M.A.; Fonseca, T.L. Dynamic (Hyper)Polarizabilities of the Ozone Molecule: Coupled Cluster Calculations Including Vibrational Corrections. J. Chem. Phys. 2011, 134, 054315. [Google Scholar] [CrossRef]
  40. Silveira, O.; Castro, M.A.; Fonseca, T.L. Vibrational Corrections to the First Hyperpolarizability of the Lithium Salt of Pyridazine Li-H3C4N2. J. Chem. Phys. 2013, 138, 074312. [Google Scholar] [CrossRef]
  41. Silveira, O.; Castro, M.A.; Leão, S.A.; Fonseca, T.L. Second Hyperpolarizabilities of the Lithium Salt of Pyridazine Li-H3C4N2 and Lithium Salt Electride Li-H3C4N2. Chem. Phys. Lett. 2015, 633, 241–246. [Google Scholar] [CrossRef]
  42. Castet, F.; Bogdan, E.; Plaquet, A.; Ducasse, L.; Champagne, B.; Rodriguez, V. Reference Molecules for Nonlinear Optics: A Joint Experimental and Theoretical Investigation. J. Chem. Phys. 2012, 136, 024506. [Google Scholar] [CrossRef]
  43. Iikura, H.; Tsuneda, T.; Yanai, T.; Hirao, K. A Long-Range Correction Scheme for Generalized-Gradient-Approximation Exchange Functionals. J. Chem. Phys. 2001, 115, 3540–3544. [Google Scholar] [CrossRef]
  44. Yanai, T.; Tew, D.P.; Handy, N.C. A New Hybrid Exchange-Correlation Functional Using the Coulomb-Attenuating Method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef] [Green Version]
  45. Zhao, Y.; Schultz, N.E.; Truhlar, D.G. Design of Density Functionals by Combining the Method of Constraint Satisfaction with Parametrization for Thermochemistry, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2006, 2, 364–382. [Google Scholar] [CrossRef]
  46. Zhao, Y.; Truhlar, D.G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Function. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef] [Green Version]
  47. Cezar, H.M.; Canuto, S.; Coutinho, K. DICE: A Monte Carlo Code for Molecular Simulation Including the Configurational Bias Monte Carlo Method. J. Chem. Inf. Model. 2020, 60, 3472–3488. [Google Scholar] [CrossRef]
  48. Wang, F.F.; Li, Z.R.; Wu, D.; Wang, B.Q.; Li, Y.; Li, Z.J.; Chen, W.; Yu, G.T.; Gu, F.L.; Aoki, Y. Structures and Considerable Static First Hyperpolarizabilities: New Organic Alkalides (M+@n6adz)M′- (M, M′ = Li, Na, K; n = 2, 3) with Cation inside and Anion Outside of the Cage Complexants. J. Phys. Chem. B 2008, 112, 1090–1094. [Google Scholar] [CrossRef]
  49. Oudar, J.L.; Chemla, D.S.; Batifol, E. Optical Nonlinearities of Various Substituted Benzene Molecules in the Liquid State, and Comparison with Solid State Nonlinear Susceptibilities. J. Chem. Phys. 1977, 67, 1626–1635. [Google Scholar] [CrossRef]
Figure 1. Stable conformers of Li@B10H14 (left, Cs conformer and right, C2v conformer).
Figure 1. Stable conformers of Li@B10H14 (left, Cs conformer and right, C2v conformer).
Liquids 03 00012 g001
Figure 2. Behavior of the dipole moment of the Cs conformer in chloroform during the iterative process. For comparison, the corresponding results for the C2v conformer (taken from Ref. [16]) are included.
Figure 2. Behavior of the dipole moment of the Cs conformer in chloroform during the iterative process. For comparison, the corresponding results for the C2v conformer (taken from Ref. [16]) are included.
Liquids 03 00012 g002
Figure 3. Absorption spectra of the Cs conformer in the gas phase and solution. For comparison, the corresponding results for the C2v conformer (taken from Ref. [16]) are included.
Figure 3. Absorption spectra of the Cs conformer in the gas phase and solution. For comparison, the corresponding results for the C2v conformer (taken from Ref. [16]) are included.
Liquids 03 00012 g003
Table 1. MP2/(Li-aug)-cc-pVDZ results for bond lengths (Å) of the Cs conformer in the gas phase and solution. In the case of chloroform, the values are convergent within an error bar of 0.001 Å.
Table 1. MP2/(Li-aug)-cc-pVDZ results for bond lengths (Å) of the Cs conformer in the gas phase and solution. In the case of chloroform, the values are convergent within an error bar of 0.001 Å.
LengthGas-PhaseChloroform
(Converged Value)
Water
(Last Step Value)
B1-Li2.2562.2953.150
B2-Li2.2442.2813.230
B3-Li2.2562.2953.228
H9-Li2.0212.0512.679
H10-Li1.9571.9802.764
H11-Li2.0212.0512.981
H23–Li2.0432.1692.819
Table 2. MP2/aug-cc-pVDZ results for the dipole moment (in a.u.) of the Cs conformer in the gas phase and solution. For comparison, the corresponding results for the Cs and C2v conformers are included.
Table 2. MP2/aug-cc-pVDZ results for the dipole moment (in a.u.) of the Cs conformer in the gas phase and solution. For comparison, the corresponding results for the Cs and C2v conformers are included.
ConformerGas-PhaseChloroform
(Converged Value)
Water
(Last Step Value)
Cs2.53 (2.55 2)2.916.56
C2v3.59 1 (3.72 2)4.59 19.42 1
1 Results taken from Ref. [16]. 2 Results taken from Ref. [17].
Table 3. MP2/aug-cc-pVDZ results for the static HRS first hyperpolarizability (in a.u.) and depolarization ratio of the Cs conformer. For comparison, the corresponding results for the C2v conformer are included.
Table 3. MP2/aug-cc-pVDZ results for the static HRS first hyperpolarizability (in a.u.) and depolarization ratio of the Cs conformer. For comparison, the corresponding results for the C2v conformer are included.
ConformerβHRS (a.u.)DR
Gas-PhaseChloroform (Converged Value)Water
(Last Step Value)
Gas-PhaseChloroform (Converged Value)Water
(Last Step Value)
Cs5042121544.944.027.23
C2v 162142786646.206.523.57
1 Results taken from Ref. [16].
Table 4. TD-DFT results for excitation energy (in eV) and (oscillator strength) of the more intense electronic transition (considering the first 10 excited states) of the Cs conformer in gas phase and in solution. For comparison, the results for the first electronic transition of the C2v conformer are included. All results were obtained using the aug-cc-pVDZ basis set.
Table 4. TD-DFT results for excitation energy (in eV) and (oscillator strength) of the more intense electronic transition (considering the first 10 excited states) of the Cs conformer in gas phase and in solution. For comparison, the results for the first electronic transition of the C2v conformer are included. All results were obtained using the aug-cc-pVDZ basis set.
Cs ConformerC2v Conformer 1
MethodGas-PhaseChloroform (Converged Value)Water (Last Step Value)Gas-PhaseChloroform (Converged Value)Water (Last Step Value)
LC-BLYP4.58 (0.0302)4.48 (0.0439)4.90 (0.0723)2.23 (0.0504)2.59 (0.0505)3.81 (0.0069)
M05-2X4.10 (0.0300)4.21 (0.0331)4.58 (0.0773)2.28 (0.0503)2.64 (0.0494)3.55 (0.0065)
M06-2X3.65 (0.0211)3.95 (0.0225)4.30 (0.0351)1.92 (0.0573)2.43 (0.0522)3.44 (0.0011)
CAM-B3LYP4.38 (0.0319)4.22 (0.0201)4.64 (0.0744)1.87 (0.0578)2.25 (0.0553)3.63 (0.0061)
1 Results taken from Ref. [16].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Brandão, I.; Fonseca, T.L.; Georg, H.C.; Castro, M.A.; Pontes, R.B. Conformational Dependence of the First Hyperpolarizability of the Li@B10H14 in Solution. Liquids 2023, 3, 159-167. https://doi.org/10.3390/liquids3010012

AMA Style

Brandão I, Fonseca TL, Georg HC, Castro MA, Pontes RB. Conformational Dependence of the First Hyperpolarizability of the Li@B10H14 in Solution. Liquids. 2023; 3(1):159-167. https://doi.org/10.3390/liquids3010012

Chicago/Turabian Style

Brandão, Idney, Tertius L. Fonseca, Herbert C. Georg, Marcos A. Castro, and Renato B. Pontes. 2023. "Conformational Dependence of the First Hyperpolarizability of the Li@B10H14 in Solution" Liquids 3, no. 1: 159-167. https://doi.org/10.3390/liquids3010012

Article Metrics

Back to TopTop