Using the Supermolecule Approach to Predict the Nonlinear Optics

Aquino, G. L. B.; Baseia, B.; Napolitano, H. B. The Solid State Structure and Environmental. 402. Polarization Effect of a Novel Asymmetric Azine. New...
0 downloads 0 Views 862KB Size
Subscriber access provided by University of South Dakota

A: Spectroscopy, Molecular Structure, and Quantum Chemistry

Using the Supermolecule Approach to Predict the Nonlinear Optics Potential of a Novel Asymmetric Azine Jean M. F. Custodio, Ricardo R. Ternavisk, Cristino J. S. Ferreira, Andreza S. Figueredo, Gilberto Lucio Benedito de Aquino, Hamilton B. Napolitano, Clodoaldo Valverde, and Basilio Baseia J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b07872 • Publication Date (Web): 18 Dec 2018 Downloaded from http://pubs.acs.org on December 20, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28 1 2 3 1 4 2 5 6 3 7 4 8 5 9 10 6 11 7 12 8 13 9 1410 15 1611 17 1812 19 20 2113 22 23 2414 25 26 15 27 28 2916 30 3117 32 3318 34 35 3619 37 3820 39 4021 41 42 4322 44 4523 46 4724 48 49 25 50 51 5226 53 54 5527 56 57 5828 59 60

The Journal of Physical Chemistry

Using the Supermolecule Approach to Predict the Nonlinear Optics Potential of a Novel Asymmetric Azine Jean M. F. Custodio1, Ricardo R. Ternavisk2,3, Cristino J. S. Ferreira4, Andreza S. Figueredo2, Gilberto L. B. Aquino2, Hamilton B. Napolitano2, Clodoaldo Valverde2,3, and Basílio Baseia3,4 Instituto de Quimica, Universidade Federal de Goiás, Goiânia, GO, 74.690-900, Brazil. Ciencias Moleculares, Universidade Estadual de Goiás, Anápolis, GO, 75.132-903, Brazil. 3 Instituto de Fisica, Universidade Paulista, Goiânia, GO, 74.845-090, Brazil. 4 Instituto de Fisica, Universidade Federal da Paraíba, João Pessoa, PB, 58.051-970, Brazil. 1 2

Emails: [email protected]; [email protected]; [email protected]

Abstract Organic molecules with electron acceptors or withdrawal substituents terminal at π-conjugated system are promising candidates to be explored as materials with high linear and nonlinear optical properties. Based on these features, a novel asymmetric azine (7E, 8E)-2-(3-methoxy-4-hydroxybenzylidene)-1-(4-Nitrobenzylidene)-hydrazineC15H13N3O4 (NMZ) was synthesized. The molecular structure of NMZ was elucidated by X-ray crystallography and the supramolecular arrangement was analyzed from Hirshfeld surface methodology. An iterative electrostatic scheme using a super molecule approach, where neighboring molecules are represented by charge points, was employed to investigate optical dipole moment (𝜇), the linear polarization (𝛼) and the first (𝛽) and second (𝛾) hyperpolarizabilities. The NMZ crystalized in the centrosymetric space group P21/n and packs via combined O-H⋯O, C-H⋯O and N⋯π interactions. The macroscopic property of third order 𝜒(3) found for the NMZ is 298.62 times greater

than

values

reported

for

for

chalcone

derivative

(2E)-1-(3-bromophenyl)-3-[4

(methylsulfanyl)phenyl]prop-2-en-1-one. The results for NMZ indicate a good nonlinear optical effect.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 229 3 4 530 6 731 8 932 10 11 1233 13 1434 15 1635 17 18 36 19 20 2137 22 2338 24 25 39 26 27 2840 29 3041 31 32 3342 34 3543 36 37 44 38 39 4045 41 4246 43 4447 45 46 4748 48 4949 50 5150 52 53 51 54 55 5652 57 5853 59 60

Page 2 of 28

1) Introduction Organic compounds have attracted the attention of researchers in recent years due their large applicability in various areas from their potential as nonlinear optical materials (NLO), e.g., as solar cell materials, photonic materials, photonic devices, optical devices, electrochemical sensors in ultra-fast optical signal processing, etc

1–10.

The fact that organic crystals present a great architectural flexibility and easy

manipulation in NLO dispositives, by simply changing their substituents and functional groups in the initial reagents, increases the attention of researchers in view of their great NLO susceptibility. So, to find new organic crystals that present NLO efficient properties is a great challenge of the present days11–13. Intermolecular interactions and electrostatic molecular behavior have significant influences on linear and nonlinear optical properties of molecules including high order nonlinear polarizations. The addition of electron donors or withdrawal groups to a π-conjugated system form a so-called push-pull mechanism (Dπ-A), which favors the formation of a charge transfer complex through the phenyl rings and conjugated system modifying considerably the values of hyperpolarizability. Among these organic compounds with potential as NLO materials, azines are highlighted in the literature. They are products of condensation between a hydrazine and two carbonyl compounds which can be designated as symmetrical or asymmetrical azines based on the similarity of the carbonyl compounds. Depending on some structural features such as planarity, electronic delocalization and the addition of electron acceptors or withdrawal groups at the ends of the π-conjugated system, a macroscopic dipole moment can be created, which explains the NLO properties14–16. Symmetrical and asymmetrical azines are important intermediates in synthesis of drugs and substances with pharmacological activities

17.

Their

biological applications as: antibacterial, anticonvulsant, anti-inflamatory, antiparasitic, insecticidal antioxidant and antitumor are some of the many of possible applications

14–16.

Some authors have

highlighted the ability of this class of compounds to form a complexation with metals. They have been recently used as ligands in coordination chemistry, due their great reactivity18. Azines have potential applications in the design liquid crystals 19, organic photovoltaic material, pigments and dyes.19,20. Azines

ACS Paragon Plus Environment

Page 3 of 28 1 1 2 3 4 2 5 6 7 3 8 9 4 10 11 5 12 13 6 14 15 16 7 17 18 8 19 20 9 21 22 2310 24 25 11 26 27 28 12 29 30 31 13 32 33 3414 35 36 3715 38 3916 40 4117 42 43 18 44 45 4619 47 48 4920 50 5121 52 5322 54 55 23 56 57 5824 59 60

The Journal of Physical Chemistry

have a good electronic linear and non-linear response and demonstrate remarkable NLO properties. NLO materials are at the forefront of light-based technology for communication and computing 21,22. Following this direction, a novel asymmetric azine (7E, 8E)-2-(3-methoxy-4-hydroxybenzylidene)-1-(4-nitrobenzylidene)hydrazine) C15H13N3O4 (NMZ) was synthetized, and its molecular structure was elucidated from X Ray Diffraction and its supramolecular arrangement was investigated by Hirshfeld surface (HS) analyses. Furthermore, considering the crystalline environment, ab initio calculations were used to estimate the dipole momentum, linear polarizability and second and third hyperpolarizabilities

11,12

in the study of the NLO properties of NMZ. The supermolecule approach

simulated the effect of the crystalline environment, considering the effects of neighboring of the molecule to improve the theoretical description of NMZ 23–25.

2) Experimental and computational procedures 2.1) Synthesis and Crystallization The microwave-assisted synthesis of NMZ (Scheme 1) was conducted by dissolving 4-hydroxy-3methoxybenzaldehyde (0.152 g, 1.0 mmol) 1 and (E)-1-(4-nitrobenzylidene) hydrazine (0.179 g, 1.0 mmol) in anhydrous ethanol (1 mL) into a microwave flask equipped with a stir bar. Then, the mixture was stirred and heated at 373K in two cycles of 90s [Discover reactor, DC-7196 CEM Corp. 2.45 GHz, adjusted power of 200 W, and an IR temperature detector]. The reaction was followed by TLC and after cooling, the precipitation of a thin yellow solid was observed. The product was vacuum filtered, and its yield was verified by GC/MS (100%). The structure was confirmed by spectrometric methods (Figure S1). FTIR spectra (Figure S2). (KBr pellet) 3500 cm-1 (free OH, stretch); 3460 cm-1 ( Hydrogen-bonded OH strech, hydroxyl), 3000 cm-1. (Csp2-H, stretch). 2932 cm-1 (Csp3-H, stretch); 1610 cm-1 (C=N, stretch); 1519 cm-1 (NO2, bend); MS analysis m/z calcd. for C15H13N3O4 299, found 299 (Figure S3). 1H NMR (500 MHz, CDCl3) δ: 8.72 (s, 1H, H-8), 8.63 (s, 1H,

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 25 2 3 426 5 627 7 8 28 9 10 1129 12 13 14 15 16 17 18 19 20 21 22 2330 24 2531 26 2732 28 29 3033 31 3234 33 3435 35 36 3736 38 39 4037 41 42 4338 44 4539 46 4740 48 49 5041 51 5242 53 5443 55 56 44 57 58 5945 60

Page 4 of 28

H-7), 8.32 (d, J = 8.7 Hz, 2H, H-11, H-13), 8.02 (d, J = 8.7 Hz, 2H, H-10, H-14), 7.57 (d, J = 1.7 Hz, 1H, H-6), 7.29 (dd, J = 8.1, 1.7 Hz, 2H, H-3), 7.02 (d, J = 8.1 Hz, 1H, H-4), 6.09 (s, 1H, H-16), 4.02 (s, J = 5.0 Hz, 3H, H-15); and

13C

NMR (125 MHz, CDCl3) δ: 163.58 (C-7); 158.27 (C-8); 128.91 (C-10, C-14);

123.9 (C-11, C-13); 123.9 (C-6); 115,30 (C-4); 108,86 (C-3); 56.58 (C-15) (Figure S4 and S5), see Scheme 1. O OH N O2N

NH2

O

O

+

microwave 373K 300 psi 250 W

HO

N

N

O2N

Scheme 1: Representation of the synthesis reaction of NMZ The crystallization has occurred by slow evaporation of the non-polar solvent dichloromethane at room temperature. The flasks were submitted to ultrasound heating for 310K in three periods of 10 min. with the purpose of assisting the solubilization of the compound. The solution was filtered and after a period of 7 min. one observes a formation of small red colored single crystals. The crystals presented regular faces, prismatic shape and polarization of light; these characteristics enabled the crystals to be used for structural analysis by x-ray diffraction methodology.

2.2) Crystallographic characterization Experimental X-ray diffraction data of NMZ were collected using graphite monochromated MoKα radiation (λ=0.71073 Å), at a Bruker APEX II CCD diffractometer. All diffraction measurements were performed at room temperature [293(2) K] and 2406 total reflections were measured (2069 independent). The acquisition of the X-ray diffraction frames and the raw dataset were treated with SAINT v8.34A. The structure was solved by direct methods using SHELXS-2016. The initial model was refined by the fullmatrix least squares method on F2 with SHELXL-2016, and non-hydrogen atoms were refined with anisotropic thermal parameters 26. The hydrogen atoms were stereo-chemically positioned and fixed with individual isotropic displacement parameters [Uiso(H) = 1.2 or 1.5 Ueq] according to the riding model with

ACS Paragon Plus Environment

Page 5 of 28 1 46 2 3 447 5 648 7 8 49 9 10 1150 12 13 51 14 15 1652 17 18 1953 20 21 54 22 23 2455 25 2656 27 2857 29 30 31 32 33 34 35 3658 37 3859 39 4060 41 42 4361 44 4562 46 4763 48 49 64 50 51 5265 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

C-H bond lengths of 0.93 Å for aromatic groups and 0.96 Å for methyl groups. The molecular representations were generated by Mercury 3.8 and Crystal Explorer v3.1 softwares. All interactions and contacts were checked from PARST7 software and Mercury 3.8 and analyzed from Hirshfeld surface 27–30. The crystallographic information files of NMZ were deposited in the Cambridge Structural Data Base under the code CCDC 1838229. Copies of the data can be obtained, free of charge, via www.ccdc.cam.ac.uk.

2.3) Hirshfeld surface analysis Hirshfeld surfaces (HS) analysis is a powerful tool to qualify and quantify the intermolecular interactions and contacts in molecular crystals. It was developed to partition the space in molecular crystals for electron density integration purposes. The HS of a molecule in a crystal is defined by the weight function W(r), dependent on the atomic electron densities where the contribution to the electron density from all molecules are equal [23]

𝑊(𝑟) =

∑𝑖 ∈ 𝑚𝑜𝑙𝑒𝑐𝑢𝑙𝑒𝜌𝑖(𝑟) ∑ 𝜌 (𝑟), 𝑖 ∈ 𝑐𝑟𝑦𝑠𝑡𝑎𝑙 𝑖

(1)

where the numerator is a summation of the atoms in the molecule of interest and the denominator is an analogous sum of the crystal. The weight function W(r) can assume the values 0 < W(r) < 1 but when the weight function is equal to 0.5, the contour surface gets the maximum electron density region. Thus, the HS can be considered as the frontier between regions where the electron distribution is dominated by the contribution of the reference molecule (interior) and of the neighboring molecules in the crystal (exterior). The normalized contact distances, dnorm, are defined in terms of di (the distance to the nearest nucleus internal to the surface), de (distance from the point to the nearest external nucleus to the surface) and the van der Waals radii of the atoms. dnorm is given by,

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 766 8 9 1067 11 1268 13 1469 15 16 1770 18 1971 20 2172 22 23 73 24 25 2674 27 28 2975 30 31 3276 33 3477 35 36 78 37 38 3979 40 4180 42 43 81 44 45 4682 47 48 4983 50 5184 52 5385 54 55 86 56 57 5887 59 60

𝑑𝑛𝑜𝑟𝑚 =

Page 6 of 28

(𝑑𝑖 ― 𝑟𝑣𝑑𝑊 ) (𝑑𝑒 ― 𝑟𝑣𝑑𝑊 ) 𝑖 𝑒 +

𝑟𝑣𝑑𝑊 𝑖

(2)

𝑟𝑣𝑑𝑊 𝑒

where 𝑟𝑣𝑑𝑊 and 𝑟𝑣𝑑𝑊 are the van der Waals radii. The interactions with distances shorter, equal or longer 𝑖 𝑒 to the sum of the van der Waals radii are represented by red, white and blue areas, respectively. HS analysis was used to investigate the intermolecular interactions in the crystal packing of NMZ and the associated two dimensional fingerprint plots, which is the combination of de and di in two dimensions32, were generated using the software Crystal Explorer 3.128. This fingerprint plot is a visual manner to investigate the nature of the interactions and quantify each kind of interaction present at a crystal molecule 33. Finally, hydrophobic interactions were studied from the shape index surface, which identify regions where two molecular surfaces touch one another30.

2.4) Theoretical Calculations The supermolecule approach is an iterative process in which the polarizing effect of the crystalline environment simulates the electrical properties of the crystal. The applicability of this approach to electrostatic incorporation is supported by the rapid convergence of the dipole moment of the crystal throughout the iterative process. Beyond the dipole moment (𝜇), the linear polarization (𝛼) and the second hyperpolarizability (𝛾) are also calculated. Since the crystal is centrosymmetric it presents no first hyperpolarizability (𝛽), 𝛽 being null. The polarization of the crystalline environment was simulated by the supermolecule (SM) approach where the atoms of the molecules within the asymmetric unit are considered as point charges. This approach has been used in several works

23–25,34–39.

In Ref.

40

the authors showed that SM approach simulates the

dipole moment with results close to the experimental values and in Refs.

41,42

the authors simulate the

macroscopic properties 𝜒(1) and 𝜒(2) of the crystal, with results close to those of experiments. Other techniques to calculate the macroscopic properties 𝜒(1) and 𝜒(2) can also be cited, e.g., the works by Seidler

ACS Paragon Plus Environment

Page 7 of 28 1 88 2 3 489 5 6 790 8 991 10 1192 12 13 93 14 15 1694 17 18 1995 20 2196 22 2397 24 25 2698 27 2899 29 30 100 31 32 101 33 34 102 35 36 37 103 38 39 104 40 41 105 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

and Champagne

43,44.

However, although the SM has provided good results, it is worth mentioning that

other factors can also affect the nonlinear properties of the crystal. To study the linear and nonlinear optical properties of the NMZ crystal we used the supermolecule method (SM) to simulate the polarization of the crystalline environment in a single NMZ molecule. For the NMZ we have used a set of 11×11×11 unit cell, with 4 asymmetric units in each unit cell, totaling 1331 unit cell with 186340 atoms; see an outline in Figure S1 the isolated NMZ is in the middle of the bulk with a blue outline. The atoms surrounding the isolated molecule were considered as point charges. The iterative process of the SM approach is carried out in several steps: firstly, we determined the electric charge of the isolated molecule by adjusting the molecular electrostatic potential (𝐶ℎ𝑒𝑙𝑝𝐺) considering the electric charges distribution in vacuum through the MP2 method. The partial atomic charges of the single isolated molecule of an asymmetric unit are calculated (𝐶ℎ𝑒𝑙𝑝𝐺). Then we replace each corresponding atom in the generated unit cells by the partial atomic charge previously obtained and the static electric properties (dipole moment (𝜇), linear polarizability (𝛼) and second (𝛾) hyperpolarizability), including the calculation of the new partial atomic charges of the asymmetric unit. The iterative process continues with the substitution of the partial atomic charges in each step of the calculation, until the convergence of the electric dipole moment be reached. The applicability of the SM approach is advantageous in view of the rapid convergence of the dipole moment of the NMZ, in which nine iterations were considered. The rapid convergence of the dipole moment can be seen in Figure 1.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 106 23 24 25 107 26 27 108 28 29 30 109 31 32 110 33 34 111 35 36 37 38 39 40 41 42 112 43 44 45 46 47 48 113 49 50 114 51 52 115 53 54 55 56 57 58 59 60

Page 8 of 28

Figure 1: Convergence of dipole moment of supermolecule approach used the set of base functions 6311+G(d) in the level Møller-Plesset of second order - MP2

In the calculations of electrical charge, dipole moment, and the linear polarizability we have used the set of base functions 6-311+G(d) in the level Møller-Plesset of second order - MP2 and for the second hyperpolarizability we employed the set of base functions 6-311+G(d) in the level CAM-B3LYP (DFT), since in the level MP2 the calculation is not implemented. In this numerical evaluation, the total dipole moment 𝜇 is defined as, 1

𝜇=(

𝜇2𝑥

+

𝜇2𝑦

+

(3)

)

𝜇2𝑧 2,

whereas the linear average polarizabilty ⟨𝛼⟩ is defined in the form, ⟨𝛼⟩ =

𝛼𝑥𝑥 + 𝛼𝑦𝑦 + 𝛼𝑧𝑧 3

,.

(4)

The average linear polarizability 〈𝛼〉 can be related to the linear refractive index (n) of the crystal by the Clausius- Mossotti relation, given by45, 𝑛2 ― 1 𝑛2 + 2

=

4𝜋𝑁 〈𝛼〉, 3

ACS Paragon Plus Environment

(5)

Page 9 of 28

The Journal of Physical Chemistry

1

116 2 3

117 4

5 6 118 7 8 9 10 11 12 13 119 14 15 16 120 17 18 19 20 21 22 121 23 24 122 25 26 27 28 29 30 123 31 32 124 33 34 35 125 36 37 126 38 39 40 127 41 42 43 128 44 45 46 129 47 48 49 130 50 51 131 52 53 132 54 55 133 56 57 58 134 59 60

here N is the number of molecules per unit cell volume. For the average of second hyperpolarizability one has 46,

⟨𝛾⟩ =

1 15



(𝛾𝑖𝑖𝑗𝑗 + 𝛾𝑖𝑗𝑖𝑗 + 𝛾𝑖𝑗𝑗𝑖)

(6)

𝑖,𝑗 = 𝑥,𝑦,𝑧

In case of static average, the second hyperpolarizability can be simplified using the symmetry relation γxxyy = γyyxx = γxyyx = γyxxy as proposed by Kleinmann approach 47 and calculated via the expression,

[

]

𝛾𝑥𝑥𝑥𝑥 + 𝛾𝑦𝑦𝑦𝑦 + 𝛾𝑧𝑧𝑧𝑧 1 . 〈𝛾〉 = 5 +2(𝛾 𝑥𝑥𝑦𝑦 + 𝛾𝑥𝑥𝑧𝑧 + 𝛾𝑦𝑦𝑧𝑧)

(7)

The experimental quantity, third-order electric susceptibility 𝜒(3), is related to the second hyperpolarizability by the expression 45,48, 𝜒(3) =

𝑁〈γ〉 , 𝜖𝑜 𝑉𝑢𝑛𝑖𝑡 𝑐𝑒𝑙𝑙

(8)

where 𝜖𝑜 is the vacuum permittivity, N is number of molecules in the unit cell. All the numerical results for the polarizability and hyperpolarizabilities tensors were obtained from the Gaussian 09 49 output file and converted by the electronic units (𝑒𝑠𝑢).

3) Results and Discussion 3.1) Crystallographic structure The asymmetric azine NMZ crystallized in the centrosymmetric monoclinic space group P21/n with four asymmetric units per unit cell and one independent molecule in asymmetric unit. The structure is composed by one para-substituted nitro group on aromatic ring A and hydroxy/methoxy groups at para/meta positions on aromatic ring B, respectively. There is an open chain conjugated to azomethine system [-C2=N2-N3=C12-] connecting the aromatic rings A and B, which provides the electronic delocalization along the π bond between the aromatic rings and the conjugated system. Figure 2 shows the

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1

29

135 2

ORTEP

136 4

compound, while the main data structure are presented in Table 1.

3

5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 137 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 138 47 48 49 139 50 140 51 141 52 53 54 142 55 56 143 57 58 59 60

Page 10 of 28

diagram of ellipsoids at 50% probability level with the atomic numbering scheme for title

Table 1: Crystal and refinement data of NMZ. Empirical formula C15H13N3O4 Formula weight 299.28 Temperature 293(2) K Wavelength 0.71073 Å Crystal system, space group Monoclinic, P21/n a = 6.9601(4) Å b = 25.5503(14) Å c = 7.9209(4) Å Unit cell dimensions α = 90º β = 103.367(2)º γ = 90º Volume 1370.43(13) Å3 Z, Calculated density 4, 1.451 Mg/m3 Absorption coefficient 0.108 mm-1 F(000) 624 Crystal size 0.436 x 0.264 x 0.058 mm Reflections collected / unique 19702 / 2621 [R(int) = 0.0357] Refinement method Full-matrix least-squares on F2 2 Goodness-of-fit on F 1.070 Final R indices [I > 2 σ(I)] R1 = 0.0450, wR2 = 0.1069 R indices (all data) R1 = 0.0560, wR2 = 0.1136

Figure 2: ORTEP diagram of ellipsoids at 50% probability level with the atomic numbering scheme for NMZ.

The geometric parameters of NMZ are given in the supplementary material (Table S1) and were validated by searching in the Cambridge Crystallographic Data Centre (CCDC) through Mogul 50,51. The

ACS Paragon Plus Environment

Page 11 of 28 1

The Journal of Physical Chemistry

144 2

main dihedral angles are presented in Table 2. The nitro group is coplanar to ring A (τ1 = 3,3(2)º), while

145 4

both methoxy and hydroxy groups are coplanar to ring B (τ5 = 179,5(1)º τ6 = 177.9(2)º). NMZ presents

3

5 6 146 7 8 147 9 10 148 11 12 13 149 14 15 150 16 17 151 18 19 20 21 22 23 24 25 26 27 28 29 152 30 31 32 153 33 34 35 154 36 37 155 38 39 40 156 41 42 157 43 44 158 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

predominantly planar geometry. Consequently, there is a parallelism of the sp2 orbitals between the ring atoms A, B and the azomethine group, resulting in the electronic delocalization of the π bond, which contributes to the structural stability. Throughout the analysis of the angle formed between the aromatic ring planes (θ = 6.85º) and the main dihedral angles (Table 2) it is possible to observe the quasi planarity of NMZ, with the main discrepancies close to the nitro group (τ1) and aromatic ring A (τ2). Table 2: Main dihedral angles of NMZ. Angle Atoms Torsion

τ1 τ2 τ3 τ4 τ5 τ6

O2-N1-C3-C2

3.3(2)

C1-C6-C7-N2

175.5(2)

C7-N2-N3-C8

178.9(1)

N3-C8-C9-C10

179.7(1)

C12-C11-O4-C15

179.5(1)

O3-C12-C13-C14

177.9(2)

Infinite zigzag 1D chains of NMZ along [010] with the 𝐶22(12)[𝑅22(10)] motif are formed via nonclassical intermolecular hydrogen bond C-H⋯O (𝑑𝐷⋯𝐴 = 3.352 Å) and a classical intermolecular hydrogen bond O3-H3⋯O2 (𝑑𝐷⋯𝐴 = 3.084 Å) involving the nitro, methoxy and hydroxy groups, as can be seen in Figure 3a (cyan color). A layer parallel to (100) is formed by the association of these chains throughout C-H⋯O (𝑑𝐷⋯𝐴 = 3.394 Å and 3.512 Å) interactions in a 𝐶12(6)[𝑅12(6)] chain motif, also shown in Figure 3a (violet color).

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 159 36 160 37 38 161 39 40 41 162 42 43 163 44 45 164 46 47 48 165 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 28

Figure 3: Intermolecular interactions (a) and crystal packing (b) of NMZ.

Finally, the layers formed are stacked through one C-H⋯O interaction in C(6) motif involving the methoxy and hydroxyl groups, stabilizing the crystal packing of NMZ Figure 3b. More information about intermolecular interactions is given in Table 3. Table 3: Intermolecular interactions involved in the NMZ crystal packing. D–H···A C2–H2⋯O4 O3–H3⋯O2 C1–H1⋯O1 C7–H7⋯O1 C15–H15B⋯O3

dD–H (Å) 0.930 0.820 0.930 0.930 0.960

dH···A (Å) 2.524 2.522 2.585 2.719 2.662

dD···A (Å) 3.352 3.084 3.394 3.512 3.358

∡D–H···A (°) 148.55 (17) 126.86 (2) 145.84 (3) 143.82 (3) 129.57 (2)

ACS Paragon Plus Environment

Symmetry code 3/2-x, -1/2+y,3/2-z 3/2-x, 1/2+y,3/2-z x, y, 1+z x, y, 1+z 1/2+x,3/2-y,-1/2+z

Page 13 of 28 1

The Journal of Physical Chemistry

166 2

HS and its fingerprint plots calculated by Crystal Explorer software have their initial atomic

167 4

parameters derived from the Crystal Structure Database 33. Figures 4a and 4b present the dnorm map of NMZ

3

5 6 168 7 8 169 9 10 170 11 12 13 171 14 15 172 16 17 173 18 19 20 174 21 22 175 23 24 25 176 26 27 177 28 29 178 30 31 32 179 33 34 180 35 36 181 37 38 39 182 40 41 183 42 43 184 44 45 46 185 47 48 186 49 50 51 187 52 53 188 54 55 189 56 57 58 59 60

whose red regions represent donor and acceptor locals with the highest incidence of intermolecular interactions, followed by the shape index mapping in Figure 4c. The regions (2), (4), (5), (6) and (12) represent di regions where the molecule acts as contact donor and (1), (7), (9), (10) and (11) represent de regions where the molecule acts as contact acceptor. Red regions (2) and (7) represent the donor and acceptor sites of the classical O3-H3⋯O2 hydrogen interaction, while (6) and (1) represent donor and acceptor sites of non-classical hydrogen bond C2-H2⋯O4. The bifurcated interaction C1-H1⋯O1 and C7H7⋯O1 with donor sites (5) and (4) and acceptor sites (10) and (9) respectively, are less intensive than C2H2⋯O4. In addition, hydrophobic contacts are represented in Figure 4c. The two-dimensional mapping of HS is shown in Figures 4d (total contacts) and 4e (O⋯H contacts), characterizing the nature and percentage of each individual contact in the crystal packing . This mapping, named fingerprint, is a two-dimensional plot of di against de, which summarizes the intermolecular contacts of NMZ molecule52. The two sharp peaks in the fingerprint plot Figure 4e correspond to reciprocal O⋯H and H⋯O interactions, representing 29.5% of the total Hirshfeld surface. Moreover, Figures 4a and 4b show the evidence of non-conventional hydrophobic intermolecular interaction of the type C5-H5⋯H14C14 viewed in detailed regions (3) and (8) with symmetry code x, y, 1+z, distance d(H⋯H) of 2.27Å (8) and angle (∡D–H···A) of 133.84º(16). The non-conventional interactions of type C-H⋯H-C predominantly with distance d(H⋯H) between 2.18 – 2.896Å agree with past studies53,54. Geometric and energetic parameters of this type of interaction in azines were analyzed in the literature55. The 2D fingerprint plot shows that H⋯H contacts constitute a high percentage with 36.3% of the surface occurring in greater proportion on the aromatic hydrogens including the region of non-conventional hydrophobic interaction C5-H5⋯H14-C14. The O⋯H and H⋯O interactions are responsible for 29.5% of the surface related to C-H⋯O and O-H⋯O interactions described previously Table 2. The hydrophobic

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1

Page 14 of 28

190 2

interactions C⋯C and C⋯H represent 5.2% and 11.5% of the total HS surface. The lower percentage of C⋯

191 4

C indicates that spatial arrangement of NMZ is not stabilized by π⋯π interactions.

3

5 6 192 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 193 42 43 194 44 195 45 196 46 47 197 48 49 198 50 51 199 52 53 54 200 55 56 201 57 58 202 59 60

Figure 4: The Hirshfeld surface dnorm mapped (A), (B) and shape index mapped (C) of NMZ visualized donor and acceptor interaction regions and showing intercontacts of the type O-H⋯O and C-H⋯O. Fingerprint plot of NMZ representing the total interaction (D) and (E), shows O⋯H interactions including reciprocal contacts.

Figure 5a shows a N⋯π interaction forming a dimer with distance of N1⋯Cg2 of 3.48(3) Å. The electron cloud of the aromatic ring acts as a π-electron donor interacting with electron deficient nitrogen N1 acting as a π-electron acceptor, as shown in details in Figure 5b. The presence of para-substituent hydroxy and meta-substituent methoxy increases the electron-donor capability of the aromatic ring B

ACS Paragon Plus Environment

Page 15 of 28 1

The Journal of Physical Chemistry

203 2

favoring the interaction with the electron-deficient N1 56–58. Figure 5c shows reciprocal contacts N⋯H with

204 4

8.6% of the surface which are related to regions where occurs N⋯.π interaction, N1⋯Cg2, contributing to

3

5 6 205 7 8 9 206 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 207 30 31 208 32 33 209 34 210 35 36 211 37 38 39 212 40 41 42 213 43 44 214 45 46 215 47 48 49 216 50 51 217 52 53 218 54 55 219 56 57 58 220 59 60

the formation of parallel layers.

Figure 5: (A) Mercury diagram of the dimer N1⋯Cg2 interaction. (B) Shape index Hirshfeld surface of NMZ highlighting close N⋯C contacts. (C) Two-dimensional fingerprint plots included reciprocal contacts C⋯H with 8.6% of the total intermolecular interaction area.

3.2) Dipole Moment The rapid convergence provided by the SM approach allows us to obtain the dipole moment (𝜇) of the crystal throughout the iterative process of electric polarization; the process begins with the atomic charges of the molecule alone and it leads to dipole moment results close to the experimental ones 40. The calculation of the dipole moment converges at 5.67D for the isolated molecule and at 6.50D for the embedded molecule, as shown in Table 4, where the latter presents a 14.64% increase in the total dipole moment and that there is a slight rotation of the polarization vector of the isolated molecule in respect to the crystal , since the shortest component μx show larger variation than the longest due to the mentioned polarization effect of the environment.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1

221 2

3 4 5 6 7 8 222 9 10 11 223 12 13 14 224 15 16 225 17 18 226 19 20 21 227 22 23 228 24 25 26 229 27 28 29 230 30 31 32 231 33 34 232 35 36 233 37 38 39 234 40 235 41 42 236 43 44 45 46 47 48 49 237 50 51 52 238 53 54 239 55 56 240 57 58 59 60

Page 16 of 28

Table 4: Results MP2/6-311+G(d) of the components and the total value of dipole moment (D). 𝝁𝒙 𝝁𝒚 𝝁𝒛 ⟨𝝁⟩ Isolated -1.02 -5.03 -2.42 5.67 Embedded -1.24 -5.77 -2.73 6.50 21.57 14.71 12.81 14.64 𝜟%

Comparing each component with the total value of the dipole moment of the isolated molecule we obtain a difference between 𝜇𝑥 and 𝜇 of 82.01%, between 𝜇𝑦 and 𝜇 of 11.29%. This clearly shows the component 𝜇𝑦 giving the most contribution for the total dipole moment of the isolated molecule. On the other hand, for the embedded molecule one obtains a difference between 𝜇𝑥 and between 𝜇𝑦with 𝜇 of 11.23% and between 𝜇𝑧 and 𝜇 of 58.00%. In the same way, this shows that the component 𝜇𝑦 gives the most contribution to the total dipole moment of the embedded molecule.

3.3) Linear polarizability The results of calculation for the linear polarization converges at 35.44 × 10-24 esu for the isolated molecule and at 35.67 × 10-24 esu for the embedded molecule, see Table 5, the latter showing an increase of 0.65%.

Table 5: ResultsMP2/6-311+G(d) of the components and of average value of the linear polarizability (in 𝟏𝟎 ―𝟐𝟒𝒆𝒔𝒖). 𝜶𝒙𝒙 𝜶𝒙𝒚 𝜶𝒚𝒚 𝜶𝒙𝒛 𝜶𝒚𝒛 𝜶𝒛𝒛 ⟨𝜶⟩ Isolated 16.76 4.53 45.03 -0.01 16.26 44.53 35.44 Embedded 16.80 4.52 45.20 0.06 16.61 45.01 35.67 0.24 -0.22 0.38 500 2.15 1.08 0.65 𝜟%

Again, a comparison of each component with the total value of the linear polarization of the isolated molecule reveals a difference between 𝛼𝑦𝑦 and 𝛼 of -27.06%, 𝛼𝑥𝑧 and 𝛼 of 99.97%. This clearly shows the component 𝛼𝑦𝑦 giving the most contribution to the total linear polarizability of the isolated molecule.

ACS Paragon Plus Environment

Page 17 of 28 1

The Journal of Physical Chemistry

241 2

On the other hand, for the embedded molecule one obtains a difference between 𝛼𝑦𝑧 and 𝛼 of 99.83%

4 242

and between 𝛼𝑧𝑧 and 𝛼 of -26.18%. In the same way, this shows the component 𝛼𝑦𝑦 being the one that most

3

5 6 243 7 8 244 9 10 11 245 12 13 246 14 15 16 247 17 18 19 248 20 21 249 22 23 24 250 25 251 26 27 28 29 30 31 32 252 33 34 253 35 36 37 254 38 39 255 40 41 42 256 43 44 257 45 46 47 258 48 49 259 50 51 52 260 53 54 55 261 56 57 58 59 60

contributes to the total linear polarizability of the embedded molecule. The 𝛼𝑥𝑧 component of the involved molecule presented a 500% increase in relation to the isolated molecule, this difference occurs due to the polarization of the crystalline environment. 3.4) Second Hiperpolarizability The value of the second hyperpolarizability converges at 262.75 × 10-36 esu for the isolated molecule and at 276.36 × 10-36 esu for the embedded molecule, the latter exhibiting an increase of 18% in comparison with the isolated molecule (Table 6). Table 6: Results DFT/6-311+G(d) for hiperpolarizability (in𝟏𝟎 ―𝟑𝟔𝒆𝒔𝒖). 𝜸𝒙𝒙𝒙𝒙 𝜸𝒚𝒚𝒚𝒚 Isolated 10.12 239.46 Embedded 10.41 245.90 2.87 2.69 𝜟%

the components and the average value of the second 𝜸𝒛𝒛𝒛𝒛 407.27 437.14 7.33

𝜸𝒙𝒙𝒛𝒛 𝜸𝒚𝒚𝒛𝒛 𝜸𝒙𝒙𝒚𝒚 5.69 317.38 5.38 5.58 333.21 5.39 -1.93 4.99 0.19

⟨𝜸⟩ 262.75 276.36 5.18

Comparing once more the value of each component with the total value of the second hyper polarizability of the isolated molecule we find a difference between 𝛾𝑥𝑥𝑥𝑥 and ⟨𝛾⟩ of 96.15%, between𝛾𝑦𝑦𝑦𝑦 and ⟨𝛾⟩ of 8.86%. This shows that component 𝛾𝑧𝑧𝑧𝑧 is the one giving the most contribution to the total value of the second hiper polarizability of the isolated molecule. On the other hand, for the embedded molecule we find a difference between 𝛾𝑥𝑥𝑦𝑦 and ⟨𝛾⟩ of 98.05%, between 𝛾𝑦𝑦𝑦𝑦 and ⟨𝛾⟩ of 11.02%. Similarly, this reveals the component 𝛾𝑧𝑧𝑧𝑧 giving the most contribution to the total value of the second hiperpolarizability of the embedded molecule.

3.5) Relating the dynamic second hyperpolarizability

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1

Page 18 of 28

262 2

In this section the second hyperpolarizability and linear polarization as a function of the vibrational

263 4

frequency are presented. The second hyperpolarizabilities ⟨𝛾( ―𝜔;𝜔,0,0)⟩ and ⟨𝛾( ―2𝜔;𝜔,𝜔,0)⟩ of the

3

5 6 264 7 8 265 9 10 266 11 12 13 267 14 15 268 16 17 269 18 19 20 270 21 22 271 23 24 25 272 26 27 273 28 29 30 31 32 33 34 35 274 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 275 53 54 276 55 277 56 57 278 58 59 60

isolated and embedded molecule, for two frequency patterns: 𝜔 = 0.0239 𝑎.𝑢. and 𝜔 = 0.0428 𝑎.𝑢, are shown in the Table 7. When we compare the value of the second hyperpolarizability of the isolated NMZ molecule in the static case ⟨𝛾(0;0,0,0)⟩, with the dynamic cases ⟨𝛾( ―𝜔;𝜔,0,0)⟩ and ⟨𝛾( ―2𝜔;𝜔,𝜔,0)⟩, for the values of frequencies in the range 0.0239-0.0428 a.u., we observe an increase of 8.25% -30.35% and [28.03% - 153.892%]. Meanwhile, the results for the crystal (embedded molecule) are 8.47% - 31.27% and [28.88% - 161.55%], as displayed in the Table 7. The dispersion curves of the average linear polarizability and average of second hyperpolarizability of the isolated molecule and the crystal are shown in Figures 6a and 6b. In both figures the linear polarizability and second hyperpolarizability present a similar behavior as a function of frequency. Table 7: CAM-B3LYP/6-311+G(d) Results of components of the second hyperpolarizability, in 10-36esu. NMZ Isolated Embedded

𝜔 = 0.0239𝑎.𝑢. and 𝜆 = 1907𝑛𝑚 ⟨𝛾( ―𝜔;𝜔,0,0)⟩ ⟨𝛾( ―2𝜔;𝜔,𝜔,0)⟩ 284.00 336.39 299.75 356.15

𝜔 = 0.0428𝑎.𝑢. and 𝜆 = 1064𝑛𝑚 ⟨𝛾( ―𝜔;𝜔,0,0)⟩ ⟨𝛾( ―2𝜔;𝜔,𝜔,0)⟩ 342.00 667.05 362.76 722.79

Figure 6: Dynamic evolution of the calculated values of the linear polarizability (in 10-24 esu) (a) and second hyperpolarizability in 10-36esu of NMZ with respective values of frequencies (b).

ACS Paragon Plus Environment

Page 19 of 28 1

The Journal of Physical Chemistry

279 2

3.6) Calculations of macroscopic properties

280 4

The SM approach was used to calculate the linear microscopic quantity, 𝜒(1) of the crystals leads to

3

5 6 281 7 8 282 9 10 283 11 12 13 284 14 15 285 16 17 286 18 19 20 21 22 287 23 24 288 25 26 289 27 28 29 290 30 31 291 32 292 33 34 35 36 37 38 39 40 41 293 42 43 44 294 45 46 295 47 48 49 296 50 51 297 52 53 298 54 55 56 299 57 58 300 59 60

results close to those of experiments

41,42.

The SM approach was also used to calculate the refraction

index n of the Crystal NMZ in the case static through the Eq. (5) and the use of data Table (6) and Table (1), the calculation resulting n = 1.82. We have obtained a first estimation of the frequency-dependent second hyperpolarizability (〈𝛾( ―𝜔;𝜔,𝜔, ― 𝜔)〉) associated to an nonlinear optical process

59

of the

intensity dependent refractive index (IDRI) from dc-K results. Following a previous work 60, for small frequencies 61, 〈𝛾( ―𝜔;𝜔,𝜔, ― 𝜔)〉 can be written as 〈𝛾( ―𝜔;𝜔,𝜔, ― 𝜔)〉≅2〈𝛾( ―𝜔;𝜔,0,0)〉 ― 〈𝛾(0;0,0,0)〉,

(9)

The values of 〈𝛾( ―𝜔;𝜔,0,0)〉 and 〈𝛼( ―𝜔;𝜔)〉 in the frequency of 0.0856 a.u. are 〈𝛾( ―𝜔;𝜔,0,0)〉 = 1066.22 × 10 ―36𝑒𝑠𝑢 〈𝛼( ―𝜔;𝜔)〉 = 45.28 × 10 ―24𝑒𝑠𝑢. Using Eqs. (5), (6), (8) and (9) and the values of Tables (1) and (6) allow us to obtain the third-order electric susceptibility of the NMZ crystal (Table 8). Table 8: DFT/CAM-B3LYP results for the linear refractive index and for the static and dynamic third2 order nonlinear susceptibility (𝑝𝑚 𝑉) . NMZ

𝜒(3)( ― ω;ω,𝜔, ― 𝜔)

𝑛(𝜔)

CAM-B3LYP

𝜔=0

𝜔 = 0.0856 𝑎.𝑢.

𝜔=0

𝜔 = 0.0856 𝑎.𝑢.

1.820

2.173

11262

75645

2 The value of 𝜒(3)( ― ω;ω,𝜔, ― 𝜔)=75645 (𝑝𝑚 𝑉) , a value 1.8 times greater than the Novel

Asymmetric Azine

39,

5.0 times greater than the novel amino chalcone

37,

8.91 times greater than the

Fluoro-N-Acylhydrazide derivative 36, and 39.79 times greater than the value for the L-arginine phosphate monohydrate crystal 34. When we compare the third-order electric susceptibility, measured via the z-scan technique of chalcones derived from 4Br4MSP, 3Br4MSP, 4N4MSP, CTDMP and 3MPNP respectively 62, the value for NMZ is 328.89, 380.12, 319.17, 31.74 and 2.73 times higher than those values, respectively.

The NMZ exhibits a larger nonlinear optical effect and implies the occurrence of microscopic third-order

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1

301 2

3 4 302 5 6 303 7 8 304 9 10 11 305 12 13 14 306 15 16 17 307 18 19 308 20 21 309 22 23 24 310 25 26 311 27 28 312 29 30 31 313 32 33 314 34 35 315 36 37 316 38 39 40 317 41 42 318 43 44 319 45 46 47 320 48 49 321 50 51 322 52 53 54 323 55 56 324 57 58 59 60

Page 20 of 28

263

NLO behavior. The typical value of 𝜒(3) reported in the literature is of the order of 100 (𝑝𝑚 𝑉)

. It is

worth mentioning that the present approach constitutes an approximation to get NLO properties of organic crystals in which other factors could also affect the calculated NLO responses.

4) Conclusions The

synthetic

asymmetric

azine

(7E,

8E)-2-(3-methoxy-4-hydroxy-benzylidene)-1-(4-

Nitrobenzylidene)-hydrazineC15H13N3O4 (NMZ) crystalized in a centrosymmetric space group P21/n with four molecules in the asymmetric unit. Its molecular structure was found to be planar due intramolecular contacts and its crystal packing is stabilized by the classical hydrogen bond O3-H3⋯O2, non-classical hydrogen bonds C2-H2⋯O4, C1-H1⋯O1, C7-H7⋯O1 and C15-H15B⋯O3. The Hirshfeld surface identified an anion⋯π interaction N1⋯Cg2 forming parallel layers. Also, it was evidenced a nonconventional hydrogen interaction C5-H5⋯H14-C14. The donor-acceptor substituted and the π-conjugated systems NMZ molecule forms a charge transfer between them responsible for a dipole moment effect. The values of dipole moment converge at 5.67D for the isolated molecule and at 6.50D for the embedded molecule. The average values of the dipole moment and the linear polarizability of NMZ were 1.41 and 7.75 times the values of the urea

64,

respectively

41,64,65.

The results for linear polarization converged at

35.44 × 10-24 esu for the isolated molecule and at 35.67×10-24 esu for the embedded molecule. The value of the second hyperpolarizability converged at 262.75 × 10-36 esu for the isolated molecule and at 276.36 × 1036

esu for the embedded molecule. The macroscopic property of third order 𝜒(3)( ― ω;ω,𝜔, ― 𝜔) found for

the NMZ is 756 times greater than that reported in the literature

63

and 380.12 times greater than that of

chalcone derivative (2E)-1-(3-bromophenyl)-3-[4 (methylsulfanyl)phenyl]prop-2-en-1-one

66.

As

consequence, the NMZ exhibits great nonlinear optical effects, which implies the occurrence of microscopic third-order NLO behavior. As future perspective, attention will be given to the study of influences of different nonlinear crystals in the behavior of optical devices.

ACS Paragon Plus Environment

Page 21 of 28 1

325 2

3 4 326 5 6 7 327 8 9 328 10 11 12 329 13 14 15 330 16 17 331 18 19 332 20 21 22 333 23 24 25 334 26 27 28 335 29 30 336 31 32 337 33 34 338 35 36 339 37 38 340 39 40 341 41 342 42 43 44 343 45 344 46 47 345 48 49 346 50 51 347 52 53 348 54 349 55 56 350 57 58 351 59 60

The Journal of Physical Chemistry

Supporting Information Complete Bond lengths and angles, chromatogram, mass, infrared and 1H-NMR spectra, HMQC plot of the title compound as well as a scheme of the bulk representing the embedded molecule.

Acknowledgements The authors gratefully acknowledge the financial support of the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES), and Fundação de Amparo a Pesquisa de Goiás (FAPEG).

References (1)

SHEIKHSHOAIE, I.; FABIAN, W. Quantum Chemical Study on the Electronic Structure and Second-Order Nonlinear Optical Properties of Salen-Type Schiff Bases. Dye. Pigment. 2006, 70, 91–98. https://doi.org/10.1016/j.dyepig.2005.04.011.

(2)

Iwan, A.; Sek, D. Polymers with Triphenylamine Units: Photonic and Electroactive Materials. Prog. Polym. Sci. 2011, 36, 1277–1325. https://doi.org/10.1016/j.progpolymsci.2011.05.001.

(3)

Niu, H.-J.; Huang, Y.-D.; Bai, X.-D.; Li, X. Novel Poly-Schiff Bases Containing 4,4′-DiaminoTriphenylamine as Hole Transport Material for Organic Electronic Device. Mater. Lett. 2004, 58, 2979–2983. https://doi.org/10.1016/j.matlet.2004.05.007.

(4)

Kong, L.-Y.; Li, Z.-W.; Okamura, T.; Ma, G.-H.; Chu, Q.; Zhu, H.-F.; Tang, S.-H.; Sun, W.-Y.; Ueyama, N. Synthesis, Structure and Optical Limiting Property of CoII, MnII and CdII Complexes with Di-Schiff Base and Reduced Di-Schiff Base Ligands. Chem. Phys. Lett. 2005, 416, 176–181. https://doi.org/10.1016/j.cplett.2005.09.074.

(5)

Kaya, İ.; Yıldırım, M.; Avcı, A. Synthesis and Characterization of Fluorescent Polyphenol Species Derived from Methyl Substituted Aminopyridine Based Schiff Bases: The Effect of Substituent Position on Optical, Electrical, Electrochemical, and Fluorescence Properties. Synth. Met. 2010, 160, 911–920. https://doi.org/10.1016/j.synthmet.2010.01.044.

(6)

Briseno, A. L.; Mannsfeld, S. C. B.; Reese, C.; Hancock, J. M.; Xiong, Y.; Jenekhe, S. A.; Bao, Z.;

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1

Xia, Y. Perylenediimide Nanowires and Their Use in Fabricating Field-Effect Transistors and

352 2

3 353 4 5 354 6 7 355 8 356 9 10 11 357 12 358 13 14 359 15 16 360 17 18 361 19 20 362 21 22 363 23 24 364 25 365 26 27 366 28 29 367 30 31 368 32 33 369 34 35 370 36 37 371 38 39 372 40 41 373 42 43 374 44 45 375 46 376 47 48 49 377 50 378 51 52 53 379 54 380 55 56 381 57 58 382 59 60

Page 22 of 28

Complementary Inverters. Nano Lett. 2007, 7, 2847–2853. https://doi.org/10.1021/nl071495u. (7)

Tan, L.; Jiang, W.; Jiang, L.; Jiang, S.; Wang, Z.; Yan, S.; Hu, W. Single Crystalline Microribbons of Perylo[1,12-b,c,d]Selenophene for High Performance Transistors. Appl. Phys. Lett. 2009, 94, 153306. https://doi.org/10.1063/1.3120769.

(8)

Jazbinsek, M.; Mutter, L.; Gunter, P. Photonic Applications With the Organic Nonlinear Optical Crystal DAST. IEEE J. Sel. Top. Quantum Electron. 2008, 14, 1298–1311. https://doi.org/10.1109/JSTQE.2008.921407.

(9)

Polisseni, C.; Major, K. D.; Boissier, S.; Grandi, S.; Clark, A. S.; Hinds, E. A. Stable, SinglePhoton Emitter in a Thin Organic Crystal for Application to Quantum-Photonic Devices. Opt. Express 2016, 24, 5615. https://doi.org/10.1364/OE.24.005615.

(10)

Zhang, L.; Pavlica, E.; Zhong, X.; Liscio, F.; Li, S.; Bratina, G.; Orgiu, E.; Samorì, P. FastResponse Photonic Device Based on Organic-Crystal Heterojunctions Assembled into a VerticalYet-Open Asymmetric Architecture. Adv. Mater. 2017, 29, 1605760. https://doi.org/10.1002/adma.201605760.

(11)

Sun, W.; McKerns, M. M.; Lawson, C. M.; Gray, G. M.; Zhan, C.-L.; Wang, D.-Y. Solvent Effect on the Third-Order Nonlinearity and Optical Limiting Ability of a Stilbazolium-like Dye; 2000; p 280. https://doi.org/10.1117/12.408515.

(12)

T. W. Bell and A. T. Papoulie. Angew. Chem., Int. Ed.; 1992.

(13)

Safari, J.; Gandomi-Ravandi, S. Structure, Synthesis and Application of Azines: A Historical Perspective. RSC Adv. 2014, 4, 46224–46249. https://doi.org/10.1039/C4RA04870A.

(14)

Wardakhan, W. W.; Sherif, S. M.; Mohareb, R. M.; Abouzied, A. S. The Reaction of Cyanoacetylhdrazine with Furan-2-Aldehyde: Novel Synthesis of Thiophene, Azole, Azine and Coumarin Derivatives and Their Antitumor Evaluation. Int. J. Org. Chem. 2012, 02, 321–331. https://doi.org/10.4236/ijoc.2012.24044.

(15)

Safari, J.; Gandomi-Ravandi, S. Structure, Synthesis and Application of Azines: A Historical Perspective. RSC Adv. 2014, 4, 46224–46249. https://doi.org/10.1039/C4RA04870A.

(16)

Al-Rasheed, H. H.; Sholkamy, E. N.; Al Alshaikh, M.; Siddiqui, M. R. H.; Al-Obaidi, A. S.; ElFaham, A. Synthesis, Characterization, and Antimicrobial Studies of Novel Series of 2,4- Bis (Hydrazino)-6-Substituted-1,3,5-Triazine and Their Schiff Base Derivatives. J. Chem. 2018, 2018, 1–13. https://doi.org/10.1155/2018/8507567.

ACS Paragon Plus Environment

Page 23 of 28 1

383 2

(17)

3 384 4 5 385 6 7 386 8 387 9 10 388 11 12 389 13 14 390 15 16 391 17 18 392 19 393 20 21 22 394 23 395 24 25 396 26 27 397 28 29 398 30 31 399 32 33 400 34 35 401 36 402 37 38 403 39 40 404 41 42 405 43 44 406 45 407 46 47 48 408 49 409 50 51 410 52 53 411 54 55 412 56 57 413 58 59 414 60

The Journal of Physical Chemistry

Veena, K.; Ramaiah, M.; Vanita, G. K.; Avinash, T. S.; Vaidya, V. P. Synthesis of Symmetrical and Asymmetrical Azines Encompassing Naphtho[2,1-b]Furan by a Novel Approach. E-Journal Chem. 2011, 8, 354–360. https://doi.org/10.1155/2011/784932.

(18)

Viñuelas-Zahínos, E.; Luna-Giles, F.; Torres-García, P.; Bernalte-García, A. Synthesis, Characterization and Crystal Structure of a Novel Thiazoline/Thiazolidine Azine Ligand and Its Nickel, Copper and Zinc Complexes. Polyhedron 2009, 28, 4056–4064. https://doi.org/10.1016/j.poly.2009.09.032.

(19)

Van Deun, R.; Parac-Vogt, T. N.; Van Hecke, K.; Van Meervelt, L.; Binnemans, K.; Guillon, D.; Donnio, B. Liquid-Crystalline Azines Formed by the Rare-Earth Promoted Decomposition of Hydrazide “Habbe” Ligands: Structural and Thermal Properties. J. Mater. Chem. 2003, 13, 1639– 1645. https://doi.org/10.1039/B303941M.

(20)

Montagner, C.; Bacci, M.; Bracci, S.; Freeman, R.; Picollo, M. Library of UV–Vis–NIR Reflectance Spectra of Modern Organic Dyes from Historic Pattern-Card Coloured Papers. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2011, 79, 1669–1680. https://doi.org/10.1016/j.saa.2011.05.033.

(21)

Nalwa, H. S.; Kakuta, A.; Mukoh, A. Third‐order Nonlinear Optical Properties of Processable Polyazine Thin Films. J. Appl. Phys. 1993, 73, 4743–4745. https://doi.org/10.1063/1.353837.

(22)

Valverde, C.; Vaz, W. F.; Custodio, J. M. F.; Duarte, V. S.; Carvalho-Jr, P. S.; Figueredo, A. S.; de Aquino, G. L. B.; Baseia, B.; Napolitano, H. B. The Solid State Structure and Environmental Polarization Effect of a Novel Asymmetric Azine. New J. Chem. 2017, 41, 11361–11371. https://doi.org/10.1039/C7NJ00618G.

(23)

Castro, A. N.; Almeida, L. R.; Anjos, M. M.; Oliveira, G. R.; Napolitano, H. B.; Valverde, C.; Baseia, B. Theoretical Study on the Third-Order Nonlinear Optical Properties and Structural Characterization of 3-Acetyl-6-Bromocoumarin. Chem. Phys. Lett. 2016, 653, 122–130. https://doi.org/10.1016/j.cplett.2016.04.070.

(24)

Vaz, W. F.; Custodio, J. M. F.; Silveira, R. G.; Castro, A. N.; Campos, C. E. M.; Anjos, M. M.; Oliveira, G. R.; Valverde, C.; Baseia, B.; Napolitano, H. B. Synthesis, Characterization, and ThirdOrder Nonlinear Optical Properties of a New Neolignane Analogue. RSC Adv. 2016, 6, 79215– 79227. https://doi.org/10.1039/C6RA14961H.

(25)

Custodio, J.; Moreira, C.; Valverde, C.; de Aquino, G.; Baseia, B.; Napolitano, H. Hirshfeld Surfaces and Nonlinear Optics on Two Conformers of a Heterocyclic Chalcone. J. Braz. Chem. Soc. 2017, 08, 168–179. https://doi.org/10.21577/0103-5053.20170136.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1

415 2

(26)

3 416

4 5 417 6 7 418 8 419 9 10 420 11 12 421 13 14 422 15 16 423 17 18 424 19 20 425 21 22 426 23 24 427 25 26 428 27 28 429 29 30 430 31 32 431 33 432 34 35 36 433 37 434 38 39 40 435 41 436 42 43 437 44 45 438 46 47 439 48 49 440 50 441 51 52 442 53 54 443 55 56 444 57 58 445 59 60

Page 24 of 28

Sheldrick, G. M. A Short History of SHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2008, 64, 112–122. https://doi.org/10.1107/S0108767307043930.

(27)

Macrae, C. F.; Bruno, I. J.; Chisholm, J. a.; Edgington, P. R.; McCabe, P.; Pidcock, E.; RodriguezMonge, L.; Taylor, R.; van de Streek, J.; Wood, P. a. Mercury CSD 2.0 – New Features for the Visualization and Investigation of Crystal Structures. J. Appl. Crystallogr. 2008, 41, 466–470. https://doi.org/10.1107/S0021889807067908.

(28)

McKinnon, J. J.; Turner, M. J.; Jayatilaka, D.; Spackman, M. A.; Grimwood, D. J.; Wolff, S. K. Crystal Explorer (Version 3.1). University of Western Australia 2012.

(29)

Farrugia, L. J. WinGX and ORTEP for Windows : An Update. J. Appl. Crystallogr. 2012, 45, 849– 854. https://doi.org/10.1107/S0021889812029111.

(30)

Spackman, M. A.; Jayatilaka, D. Hirshfeld Surface Analysis. 2009, 19–32. https://doi.org/10.1039/b818330a.

(31)

Jelsch, C.; Ejsmont, K.; Huder, L. The Enrichment Ratio of Atomic Contacts in Crystals, an Indicator Derived from the Hirshfeld Surface Analysis. IUCrJ 2014, 1, 119–128. https://doi.org/10.1107/S2052252514003327.

(32)

McKinnon, J. J.; Spackman, M. a.; Mitchell, A. S. Novel Tools for Visualizing and Exploring Intermolecular Interactions in Molecular Crystals; 2004; Vol. 60. https://doi.org/10.1107/S0108768104020300.

(33)

Spackman, M. A.; McKinnon, J. J. Fingerprinting Intermolecular Interactions in Molecular Crystals. CrystEngComm 2002, 4, 378–392. https://doi.org/10.1039/B203191B.

(34)

Valverde, C.; Rodrigues, R. F. N.; Machado, D. F. S.; Baseia, B.; de Oliveira, H. C. B. Effect of the Crystalline Environment on the Third-Order Nonlinear Optical Properties of L-Arginine Phosphate Monohydrate: A Theoretical Study. J. Mol. Model. 2017, 23, 122. https://doi.org/10.1007/s00894-017-3274-3.

(35)

Castro, A. N.; Osório, F. A. P.; Ternavisk, R. R.; Napolitano, H. B.; Valverde, C.; Baseia, B. Theoretical Investigations of Nonlinear Optical Properties of Two Crystalline Acetamides Structures Including Polarization Effects of Their Environment. Chem. Phys. Lett. 2017, 681, 110– 123. https://doi.org/10.1016/j.cplett.2017.05.066.

(36)

Rodrigues, R. F. N.; Almeida, L. R.; Santos, F. G. dos; Carvalho Jr, P. S.; Souza, W. C. de; Moreira, K. S.; Aquino, G. L. B. de; Valverde, C.; Napolitano, H. B.; Baseia, B. Solid State Characterization and Theoretical Study of Non-Linear Optical Properties of a Fluoro-N-

ACS Paragon Plus Environment

Page 25 of 28 1

Acylhydrazide Derivative. PLoS One 2017, 12, e0175859.

446 2

3 447

4 5 448 6 7 449 8 450 9 10 451 11 12 452 13 14 453 15 16 454 17 18 455 19 20 456 21 22 457 23 458 24 25 459 26 27 460 28 29 461 30 31 462 32 33 463 34 35 464 36 37 465 38 39 466 40 467 41 42 468 43 44 469 45 46 470 47 48 471 49 50 472 51 52 473 53 474 54 55 56 475 57 476 58 59 60

The Journal of Physical Chemistry

https://doi.org/10.1371/journal.pone.0175859. (37)

Almeida, L. R.; Anjos, M. M.; Ribeiro, G. C.; Valverde, C.; Machado, D. F. S.; Oliveira, G. R.; Napolitano, H. B.; de Oliveira, H. C. B. Synthesis, Structural Characterization and Computational Study of a Novel Amino Chalcone: A Potential Nonlinear Optical Material. New J. Chem. 2017, 41, 1744–1754. https://doi.org/10.1039/C5NJ03214H.

(38)

Ribeiro, G. C.; Almeida, L. R.; Napolitano, H. B.; Valverde, C.; Baseia, B. Polarization Effects on the Third-Order Nonlinear Optical Properties of Two Polymorphs of Enamine Derivative. Theor. Chem. Acc. 2016, 135, 244. https://doi.org/10.1007/s00214-016-1999-1.

(39)

Valverde, C.; Vaz, W. F.; Custodio, J. M. F.; Duarte, V. S.; Carvalho-Jr, P. S.; Figueredo, A. S.; de Aquino, G. L. B.; Baseia, B.; Napolitano, H. B. The Solid State Structure and Environmental Polarization Effect of a Novel Asymmetric Azine. New J. Chem. 2017, 41, 11361–11371. https://doi.org/10.1039/C7NJ00618G.

(40)

Fonseca, T. L.; Sabino, J. R.; Castro, M. A.; Georg, H. C. A Theoretical Investigation of Electric Properties of L -Arginine Phosphate Monohydrate Including Environment Polarization Effects. J. Chem. Phys. 2010, 133, 144103. https://doi.org/10.1063/1.3501237.

(41)

Santos, O. L.; Fonseca, T. L.; Sabino, J. R.; Georg, H. C.; Castro, M. A. Polarization Effects on the Electric Properties of Urea and Thiourea Molecules in Solid Phase. J. Chem. Phys. 2015, 143, 234503. https://doi.org/10.1063/1.4937481.

(42)

Santos, O. L.; Sabino, J. R.; Georg, H. C.; Fonseca, T. L.; Castro, M. A. Electric Properties of the 3-Methyl-4-Nitropyridine-1-Oxyde (POM) Molecules in Solid Phase: A Theoretical Study Including Environment Polarization Effect. Chem. Phys. Lett. 2017, 669, 176–180. https://doi.org/10.1016/j.cplett.2016.12.042.

(43)

Seidler, T.; Champagne, B. Second-Order Nonlinear Optical Susceptibilities of Metal–Organic Frameworks Using a Combined Local Field Theory/Charge Embedding Electrostatic Scheme. J. Phys. Chem. C 2016, 120, 6741–6749. https://doi.org/10.1021/acs.jpcc.6b00217.

(44)

Seidler, T.; Stadnicka, K.; Champagne, B. Investigation of the Linear and Second-Order Nonlinear Optical Properties of Molecular Crystals within the Local Field Theory. J. Chem. Phys. 2013, 139, 114105. https://doi.org/10.1063/1.4819769.

(45)

Senthil, K.; Kalainathan, S.; Kumar, A. R.; Aravindan, P. G. Investigation of Synthesis, Crystal Structure and Third-Order NLO Properties of a New Stilbazolium Derivative Crystal: A Promising

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1

Material for Nonlinear Optical Devices. RSC Adv. 2014, 4, 56112–56127.

477 2

3 478 4 5 479 6 7 480 8 481 9 10 482 11 12 483 13 14 484 15 16 485 17 18 486 19 20 487 21 22 488 23 24 489 25 490 26 27 491 28 29 492 30 31 493 32 33 494 34 35 495 36 37 496 38 39 497 40 41 498 42 43 499 44 500 45 46 501 47 48 502 49 50 503 51 52 504 53 54 505 55 56 506 57 58 507 59 60

Page 26 of 28

https://doi.org/10.1039/C4RA09112D. (46)

Kongsted, J.; Osted, A.; Mikkelsen, K. V; Christiansen, O. Second Harmonic Generation Second Hyperpolarizability of Water Calculated Using the Combined Coupled Cluster Dielectric Continuum or Different Molecular Mechanics Methods. J. Chem. Phys. 2004, 120, 3787–3798. https://doi.org/10.1063/1.1642593.

(47)

KLEINMAN, D. A. Nonlinear Dielectric Polarization in Optical Media. Phys. Rev. 1962, 126, 1977–1979. https://doi.org/10.1103/PhysRev.126.1977.

(48)

Gubler, U.; Bosshard, C. Optical Third-Harmonic Generation of Fused Silica in Gas Atmosphere: Absolute Value of the Third-Order Nonlinear Optical Susceptibility Χ^(3). Phys. Rev. B 2000, 61, 10702–10710. https://doi.org/10.1103/PhysRevB.61.10702.

(49)

Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; et al. Gaussian 09, Revision A.02. Gaussian Inc Wallingford CT. Gaussian, Inc. 2009, p Wallingford CT. https://doi.org/10.1159/000348293.

(50)

Groom, C. R.; Allen, F. H. The Cambridge Structural Database in Retrospect and Prospect. Angew. Chemie - Int. Ed. 2014, 53, 662–671. https://doi.org/10.1002/anie.201306438.

(51)

Bruno, I. J.; Cole, J. C.; Kessler, M.; Luo, J.; Momerwell, W. D. S.; Purkis, L. H.; Smith, B. R.; Taylor, R.; Cooper, R. I.; Harris, S. E.; et al. Retrieval of Crystallographically-Derived Molecular Geometry Information. J. Chem. Inf. Comput. Sci. 2004, 44, 2133–2144. https://doi.org/10.1021/ci049780b.

(52)

Custodio, J. M. F.; Moreira, C. A.; Valverde, C.; De Aquino, G. L. B.; Baseiaa, B.; Napolitano, H. B. Hirshfeld Surfaces and Nonlinear Optics on Two Conformers of a Heterocyclic Chalcone. J. Braz. Chem. Soc. 2018, 29. https://doi.org/10.21577/0103-5053.20170136.

(53)

Echeverría, J.; Aullón, G.; Danovich, D.; Shaik, S.; Alvarez, S. Dihydrogen Contacts in Alkanes Are Subtle but Not Faint. Nat. Chem. 2011, 3, 323–330. https://doi.org/10.1038/nchem.1004.

(54)

Robertson, K. N.; Knop, O.; Cameron, T. S. C-H···H-C Interactions in Organoammonium Tetraphenylborates: Another Look at Dihydrogen Bonds. Can. J. Chem. 2003, 81, 727–743. https://doi.org/10.1139/v03-080.

(55)

de Almeida, L. R.; Carvalho, P. S.; Napolitano, H. B.; Oliveira, S. S.; Camargo, A. J.; Figueredo, A. S.; de Aquino, G. L. B.; Carvalho-Silva, V. H. Contribution of Directional Dihydrogen

ACS Paragon Plus Environment

Page 27 of 28 1

Interactions in the Supramolecular Assembly of Single Crystals: Quantum Chemical and Structural

508 2

3 509

4 5 510 6 7 511 8 512 9 10 11 513 12 514 13 14 515 15 16 516 17 18 517 19 20 518 21 22 519 23 24 520 25 26 521 27 522 28 29 523 30 31 524 32 33 525 34 35 526 36 37 527 38 39 528 40 41 529 42 43 530 44 531 45 46 47 532 48 533 49 50 534 51 52 535 53 54 536 55 56 537 57 58 538 59 60

The Journal of Physical Chemistry

Investigation of C17H17N3O2 Azine. Cryst. Growth Des. 2017, 17, 5145–5153. https://doi.org/10.1021/acs.cgd.7b00585. (56)

Ma, J. C.; Dougherty, D. A. The Cation−π Interaction. Chem. Rev. 1997, 97, 1303–1324. https://doi.org/10.1021/cr9603744.

(57)

Zhang, Z.; Wu, Y.; Zhang, G. N⋯π and O⋯π Interactions Rather than “Directed” Hydrogen Bonding in the Crystal Packing of Ethyl (Z)-2-Cyano-3-[(4-Alkoxyphenyl)Amino]Prop-2-Enoates. CrystEngComm 2011, 13, 4496. https://doi.org/10.1039/c1ce05199g.

(58)

Du, Q.-S.; Long, S.-Y.; Meng, J.-Z.; Huang, R.-B. Empirical Formulation and Parameterization of Cation-π Interactions for Protein Modeling. J. Comput. Chem. 2012, 33, 153–162. https://doi.org/10.1002/jcc.21951.

(59)

Maldonado, M.; Baltar, H. T. M. C. M.; Gomes, A. S. L.; Vaia, R.; Park, K.; Che, J.; Hsiao, M.; de Araújo, C. B.; Baev, A.; Prasad, P. N. Coupled-Plasmon Induced Optical Nonlinearities in Anisotropic Arrays of Gold Nanorod Clusters Supported in a Polymeric Film. J. Appl. Phys. 2017, 121, 143103. https://doi.org/10.1063/1.4980027.

(60)

Marques, S.; Castro, M. A.; Leão, S. A.; Fonseca, T. L. Second Hyperpolarizability of the Calcium-Doped Lithium Salt of Pyridazine Li–H 3 C 4 N 2 ⋯ Ca. Chem. Phys. Lett. 2016, 659, 76–79. https://doi.org/10.1016/j.cplett.2016.07.009.

(61)

Bishop, D. M.; De Kee, D. W. The Frequency Dependence of Nonlinear Optical Processes. J. Chem. Phys. 1996, 104, 9876–9887. https://doi.org/10.1063/1.471752.

(62)

Prabhu, S. R.; Jayarama, A.; Chandrasekharan, K.; Upadhyaya, V.; Ng, S. W. Synthesis, Growth, Structural Characterization, Hirshfeld Analysis and Nonlinear Optical Studies of a Methyl Substituted Chalcone. J. Mol. Struct. 2017, 1136, 244–252. https://doi.org/10.1016/j.molstruc.2017.01.069.

(63)

Günter, P. Nonlinear Optical Effects and Materials; G?nter, P., Ed.; Springer Series in Optical Sciences; Springer Berlin Heidelberg: Berlin, Heidelberg, 2000; Vol. 72. https://doi.org/10.1007/978-3-540-49713-4.

(64)

Adant, C.; Dupuis, M.; Bredas, J. L. Ab Initio Study of the Nonlinear Optical Properties of Urea: Electron Correlation and Dispersion Effects. Int. J. Quantum Chem. 1995, 56, 497–507. https://doi.org/10.1002/qua.560560853.

(65)

Pluta, T.; Sadlej, A. J. Electric Properties of Urea and Thiourea. J. Chem. Phys. 2001, 114, 136.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1

https://doi.org/10.1063/1.1328398.

539 2 3

540 4 5 541 6 7 542 8 9 543 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 28

(66)

D’silva, E. D.; Podagatlapalli, G. K.; Venugopal Rao, S.; Dharmaprakash, S. M. Study on ThirdOrder Nonlinear Optical Properties of 4-Methylsulfanyl Chalcone Derivatives Using Picosecond Pulses. Mater. Res. Bull. 2012, 47, 3552–3557. https://doi.org/10.1016/j.materresbull.2012.06.063.

ACS Paragon Plus Environment