Chalcone as Potential Nonlinear Optical Material: A Combined

ACS2GO © 2019. ← → → ←. loading. To add this web app to the home screen open the browser option menu and tap on Add to homescreen...
0 downloads 0 Views 6MB Size
Article Cite This: J. Phys. Chem. C XXXX, XXX, XXX−XXX

pubs.acs.org/JPCC

Chalcone as Potential Nonlinear Optical Material: A Combined Theoretical, Structural, and Spectroscopic Study Jean M. F. Custodio,*,† Giulio D. C. D’Oliveira,† Fernando Gotardo,‡ Leandro H. Z. Cocca,‡ Leonardo De Boni,‡ Caridad N. Perez,† Lauro J. Q. Maia,† Clodoaldo Valverde,*,§,∥ Francisco A. P. Osoŕ io,†,⊥ and Hamilton B. Napolitano*,§ †

Universidade Federal de Goiás, Goiânia, GO, Brazil Instituto de Fı ́sica de São Carlos, Universidade de São Paulo, São Carlos, 13566-590, SP, Brazil § Universidade Estadual de Goiás, Anápolis 75132-903, GO, Brazil ∥ Universidade Paulista, Goiânia, 74845-090, GO, Brazil ⊥ Pontifícia Universidade Católica de Goiás, Goiânia 74175-120, GO, Brazil

Downloaded via WASHINGTON UNIV on March 6, 2019 at 14:53:50 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



ABSTRACT: In this work, we propose the synthesis of a novel bromine chalcone (E)-3-(2-bromophenyl)-1-(2((phenylsulfonyl)amine))phenyl)prop-2-en-1-one (BRC) that has been crystallized by the slow evaporation technique. The second-order molecular optical scattering and twophoton absorption(2PA) spectrum of the BRC molecule dissolved in dimethyl sulfoxide (DMSO) were evaluated by using hyper-Rayleigh scattering and the femtosecond tunable Z-scan techniques. The first-order hyperpolarizability of BRC dissolved in DMSO was estimated by using a simplified twolevel model, in which one- and two-photon absorption parameters were used as input information to the model. The BRC crystal was characterized from single-crystal X-ray diffraction (XRD) and spectroscopy analyzes. Also, the thermogravimetric analyses and the fluorescence spectra were obtained. In addition, an ab initio calculation method, which includes the Møller− Plesset perturbation theory (MP2) and the density functional theory (DFT) at the CAM-B3LYP level, was used to estimate the crystal linear refractive index and the third-order electric susceptibility. Also, the average first hyperpolarizability of BRC molecules dissolved in DMSO was calculated and compared with the experimental results. The obtained values are good and qualify the BRC crystal as a potential candidate for application in nonlinear optical devices.

1. INTRODUCTION In the past decade, interest in the investigation of organic crystals exhibiting nonlinear optical properties (NLO) has been stimulated due to the ease of manipulation of these compounds. Their NLO properties can be controlled in a certain way by structural modifications that offer a great number of possibilities for the device design. The great potential applications of the organic crystals are in photonic devices with fast responses,1 in photovoltaic devices for highperformance photoelectric conversion,2 and also for manufacturing high-performance photodetectors.3 Chemically, chalcones are α,β-unsaturated carbonyl compounds with two phenyl rings, attached to carbonyl and β carbon, respectively.4 Some reviews summarized the synthetic routes of chalcones.5−8 There is considerable interest in chalcones due to their wide range of biological activities whether of natural origin or synthetic origin.9−13 Some activities highlighted are cytotoxic,14 anticancer,15,16 antiinflammatory,17 chemopreventive,18 and host biomolecule.19 Although its biological application has been widely studied, the use of chalcones as potential optical material is still a matter of interest in the scientific field. In this sense, this paper presents © XXXX American Chemical Society

the synthesis and characterization of a new bromine chalcone (E)-3-(2-bromophenyl)-1-(2-((phenylsulfonyl)amine)phenyl)prop-2-en-1-one (BRC), with molecular formula C21H16BrNO3SNa. The synthesized molecule was carefully characterized in terms of linear and nonlinear optical properties in dimethyl sulfoxide (DMSO) solution. For that, two well-known nonlinear experimental techniques were employed to determine precisely the second-order molecular scattering and twophoton absorption spectrum. Second-order molecular scattering was obtained at 1064 nm by using hyper-Rayleigh scattering (HRS) technique, and the two-photon absorption (2PA) spectrum was evaluated at the therapeutic window by using the femtosecond tunable Z-scan technique. Besides that, BRC was characterized from single-crystal X-ray diffraction (DRX) and spectroscopy analyses. In addition, an ab initio calculation method, which includes the Møller−Plesset perturbation theory (MP2) and the density functional theory (DFT) at the CAM-B3LYP exchange−correlation functional, Received: February 1, 2019 Published: February 21, 2019 A

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

cursor 1 (2.0 mmol, 0.5506 g) and 2-bromobenzaldehyde (4.0 mmol, 0.7401 g) were dissolved in 40 mL of basic ethanol (0.319 g of dissolved potassium hydroxide), and the solution was reacted (25 °C) for 25 h. The system was neutralized with hydrochloric acid, poured into water (80 mL), and then extracted with dichloromethane (20 mL). The organic phase evaporates slowly. Before being completely dry, the product was collected, rinsed with ethanol, and dried to afford a yellow crystalline material (6.85 g, 77.4%). Degree of purity of 96.7%, mp 151−153 °C. 1H NMR (CDCl3): δ 7.15 (ddd, J 1.19 Hz, 7.39 Hz, 7.99 Hz, 1H), δ 7.26 (d, J 15.55 Hz, 1H), δ 7.26− 7.30 (m, 1H), δ 7.35−7.37 (m, 1H), δ 7.38−7.46 (m, 3H), δ 7.50 (ddd, J 1.53 Hz, 7.40 Hz, 8.35 Hz, 1H), δ 7.65−7.67 (m, 2H), δ 7.77 (dd, J 1.05 Hz, 8.40 Hz, 1H), δ 7.82−7.85 (m, 3H), δ 7.99 (d, J 15.55 Hz, 1H), δ 11.10 (s, 1H). 13C NMR (CDCl3): δ 120.90, 123.21, 124.62, 124.89, 126.09, 127.26, 127.79, 127.90, 129.02, 130.80, 131.75, 132.90, 133.72, 134.53, 139.38, 139.94, 144.25, 192.66. IR 1641 (m), 1493 (m), 1330 (m), 932 (m), 751 (s ). HRMS calculated for C21H16BrNO3SNa 463.9932, found 463.9837. 2.2. Linear and Nonlinear Optical Properties of BRC Molecules Dissolved in DMSO. 2.2.1. Absorption Spectrum Measurement. The absorption spectrum of the sample dissolved in dimethyl sulfoxide (DMSO) was obtained in the UV−vis−NIR region by using a SHIMADZU UV-1800.These measurements were performed in a 10 mm optical path length fused silica cuvette with a concentration of about 10−5 mol/L. 2.2.2. Two-Photon Absorption Measurements. To determine two-photon absorption (2PA) cross section spectrum, the femtosecond tunable Z-scan technique was used.21 Z-scan measurements consist of translating the sample around the focus of the focalized laser beam and collecting changes in the optical transmittance of the sample as a function of this position (Z position). The consequence of establishing a focused beam is to create a spatial region with high intensity at the focal position, in which the nonlinear effects, such as 2PA, are induced. Therefore, around the focal position, the transmittance of the laser by the sample decreases proportionally to the 2PA events. However, when the sample is far from the focal position, the light intensity is low, and the nonlinear effects (2PA) are not created. Consequently, only the linear optical transmittance is observed. The ratio between the transmittance at the focal position by the transmittance far from the focus describes a normalized transmittance, in which the 2PA cross section can be evaluated. More details about this technique can be found elsewhere.21 The 2PA cross section spectrum was obtained by employing a tunable optical parametric amplifier (TOPAS), which allows one to choose wavelengths from 470 nm to 2000 nm. TOPAS is pumped by a Ti:sapphire CPA laser with a 150 fs pulse width at 775 nm, operating at a 1 kHz repetition rate. The wavelengths delivered by TOPAS present a 1 kHz repetition rate and 120 fs pulse width. Once the wavelength is set, the beam is focused by a convergent lens of about 15 cm, generating different intensity profiles around the focus, in which the sample is translated. The transmittance of the light by the sample as a function of each Z position is registered by a silicon photodetector and amplified by a commercial locking amplifier. The 2PA cross section is obtained by fitting the experimental normalized transmittance, T(Z), with the following relation:21

were used to calculate the electrical parameters of the BRC crystal via the supermolecule approach (SM); both calculation methods employed the 6-311++G(d,p) basis set. Such parameters include dipole moment, linear polarizability, and second hyperpolarizability. These parameters were used to estimate the crystal linear refractive index and the third-order electric susceptibility. Also, the average first hyperpolarizability of BRC molecules dissolved in DMSO was calculated at the DFT/CAM-B3LYP/cc-pVTZ level and the results were compared with the experimental and estimated values by the simplified two-level model, showing a good agreement between them.

2. EXPERIMENTAL: MATERIALS AND METHODS 2.1. General Procedures. Aiming to obtain the compounds, the precursor 2′N-(phenylsulfonyl)acetophenone (precursor 1) was synthesized by reaction between benzenesulfonyl chloride and 2-aminoacetophenone in dichloromethane (Scheme 1), based on conditions previously Scheme 1. General Conditions for the Synthesis of 2′N(Phenylsulfonyl)acetophenone

described in the literature.20 Compound BRC was synthesized by means of Claisen−Schmidt condensation between precursor 1 and a benzaldheyde via basic catalysis (Scheme 2). Scheme 2. General Conditions for the Synthesis of Compound BRC

2.1.1. Synthesis of 1-(2-(Phenylsulfonylamino)phenyl)ethanone (Precursor 1). Benzenesulfonyl chloride (70.65 g, 0.40 mol), 2-aminoacetophenone (67.58 g, 0.50 mol), and triethylamine (50.55 g, 0.50 mol) were dissolved in 500 mL of dichloromethane, and the solution was kept in the refrigerator for 37 h. The solution was filtered, and the crystals were rinsed with methanol. Yield: 55.1 g (50.1%) of a white crystalline solid. 1H NMR (CDCl3): δ 2.56 (s, 3H), δ 7.08 (ddd, J 7.98 Hz, 7.33 Hz, 1.18 Hz, 1H), 7.42−7.46 (m, 2H), 7.46 (dddd, J 8.44 Hz, 7.34 Hz, 1.56 Hz, 0.41 Hz, 1H), 7.51−7.54 (m, 1H), 7.70 (ddd, J 8.40 Hz, 1.15 Hz, 0.45 Hz, 1H), 7.80 (ddd, J 8.01 Hz, 1.59 Hz, 0.46 Hz, 1H), 7.84−7.87 (m, 2H), 11.50 (s, 1H). 13 C NMR (CDCl3): δ 28.17, 119.25, 122.41, 122.77, 127.24, 129.05, 131.91, 133.01, 134.97, 139.48, 139.92, 202.43. HRMS (high-resolution mass spectrometry) calculated for C14H14NO3S 276.0694, found 276.0725. 2 . 1 . 2 . S yn t h e s i s o f ( E ) - 3 - ( 2 - B r o m o p h e n y l ) - 1 (2phenylsulfonylamine)phenyl)prop-2-en-1-one (BRC). PreB

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C T (Z ) =

1 π q0(Z ,0)



solution program using direct methods and refined with the ShelXL28 refinement package using least squares minimization with no restrains. Aromatic/amide H atoms were refined with riding coordinates and fixed Uiso at 1.2 times. Platon29 and Parst30 softwares were used to validate the chemical parameters found for BRC. Tables, pictures, and images were generated by using Mercury,31 ORTEP, 32 and UCSF Chimera33 programs. Crystal data are available via the Cambridge Crystallographic database with CCDC code 1895208. 2.3.1. Thermal Analysis. Thermogravimetric analyses were carried out on a Shimadzu TGA-60 thermobalance. Approximately 3.5 mg of sample was placed on an alumina pan and heated from room temperature up to 550 °C, at 10 °C min−1, under a nitrogen flow (50 mL min−1). Differential scanning calorimetric measurements were performed on a Shimadzu DSC-60 instrument. Samples (3.5 ± 0.5 mg) were in aluminum pans and heated from room temperature up to 400 °C, at 10 °C min−1, under a nitrogen flow (50 mL min−1). All data were processed using the Shimadzu TA-60 thermal data analysis software. Microscopy was performed on a Leica DM2500P microscope connected to the Linkam T95-PE hotstage equipment. Data were visualized with the Linksys 32 software for hot stage control. The crystals of BRC were placed on a 13 mm glass coverslip, placed on a 22 mm diameter pure silver heating block inside the stage. The sample was heated at a ramp rate of 10 °C min−1 up to a final temperature of 355 °C but discontinued on melting of all material. 2.3.2. Photoluminescence Spectroscopy. Fluorescence spectra were measured by using a spectrofluorometer HORIBA-JOBIN YVON Fluorolog FL3-221 equipped with a Xe arc lamp, double monochromators, and a photomultiplier tube. These measurements were performed in a powdered sample placed on a black painted metallic holder.

2

∫−∞ ln[1 + q0(Z ,0)e−t ] dt

(1)

/Z02)−1. I0

in which q0 = α2PAI0L(1 + Z is the pulse intensity, Z0 is the Rayleigh length, L is the optical path of the sample, and the Z is the sample position. The nonlinear absorption coefficient, α2PA, is related to the 2PA cross-section, σ2PA, by σ2PA = (hν/N)α2PA, in which h is the Planck constant, ν is the light frequency, and N is the number of molecule by cm3. The Göppert−Mayer unit (GM),22 1 GM = 1 × 10−50 cm4 s photon−1, is commonly used to quantify the σTPA. 2.2.3. First-Order Hyperpolarizability Measurements. The hyper-Rayleigh scattering (HRS) technique23 was used to determine the first hyperpolarizability (β) of the chalcone dissolved in DMSO. In order to perform this experiment, a Nd:YAG laser was employed to induce the nonlinear optical scattering in the sample. This laser delivers pulses with 100 ps pulse width at 1064 nm (ν), which is focused on the sample by a set of convergent lenses. The scattered optical second harmonic at 532 nm (2ν) generated by the sample is collected by a photomultiplier tube positioned perpendicularly to the incident beam (1064 nm).To warranty that only the nonlinear scattered signal is achieving the photomultiplier, an optical filter centered at 532 nm with 10 nm of bandwidth is positioned at the photomultiplier window, allowing only the hyper-Rayleigh signal to pass. More details about the technique used in this work can be found elsewhere.24 The second harmonic signal I(2ω) is related to the laser intensity I(ω) by the following expression: 2

M

I(2ω) = G∑ Niβi 2I 2(ω) i=1

(2)

in which Ni is the molecular concentration and G is an instrumental factor that depends on the beam parameters and experimental details of the setup. This experimental factor can be calibrated by using another sample with a well-known β value. In this work, the p-nitroaniline (pNA) molecule dissolved in DMSO, β(1064 nm) = 25.3 × 10−30 cm5/esu,25 was used to determine G. This factor is constant during the experiment because the experimental conditions are always kept the same. To determine β for the chalcone sample, HRS measurements were performed at several molecular concentrations (N) for pNA and chalcone solutes. Consequently, it was possible to determine the quadratic coefficients of eq 2, G(Nsolventβsolvent2 + Nsoluteβsolute2), for each individual concentration of both samples. The outcome for each individual sample is a linear dependence between of the quadratic coefficient and N. From the linear dependences, the angular coefficients, given by αchalcone = Gβchalcone2 and αpNA = GβpNA2, can be correlated to each other by the same instrumental factor G by the following relation: βchalcone =

βpNA 2

αchalcone αpNA

3. THEORETICAL CALCULATIONS: LINEAR AND NONLINEAR OPTICAL PROPERTIES OF BRC The supermolecule (SM) approach was used to study the linear and nonlinear electrical properties of the BRC crystal. In the SM approach, a bulk consisting of a set of 11 × 11 × 11 unit cells was constructed, where each unit cell has four asymmetric units with 172 atoms, generating a bulk of 228 932 atoms; see a sketch in Figure 1. The atoms that surround the molecule in green in the center of the image are treated as point charges. Here we consider a single BRC molecule and through the CHELPG scheme the crystalline environment

(3)

2.3. Single-Crystal Analysis. Single crystals of BRC were crystallized from diffusion of ethyl ether vapor into a dichlomethane solution. A suitable crystal was selected and mounted on a XtaLAB Mini II diffractometer equipped with graphite-monochromated Mo Kα radiation (λ = 0.710 73 Å). The crystal was kept at 298.15 K during data collection. Using Olex2,26 the structure was solved with the ShelXS27 structure

Figure 1. Model of the bulk. C

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C polarization is simulated by considering the atoms that surrounding the BRC molecule as point charges. The SM approach was used in several theoretical works34,35,44−46,36−43 and the obtained results were found close to the experimental ones. Particularly in ref 47, the DFT/CAM-B3LYP/6-311+ +G(d,p) results for the nonlinear third-order susceptibility for a chalcone derivative have described satisfactorily the Z-scan experimental result of Prahbu et al.48 To build the BRC bulk, first the molecular electrostatic potential (MEP) of an isolated molecule is determined within the MP2/6-311++G(d,p) model and through the CHELPG scheme a set of partial charges that fits the MEP is found. Then the SM iterative process starts calculating the CHELPG charges for an isolated molecule of the unit cell. Then each corresponding atom of the bulk was replaced by the partial atomic charge previously obtained, and the static electric properties (dipole moment, linear polarizability, and second hyperpolarizabilities) and the new partial atomic charges of the asymmetric unit are calculated. The iterative process continues with the substitution of the partial atomic charges in each calculation step, until the convergence of the electric dipole moment is reached; see Figure 2.

N ⟨γ ⟩ , ϵoV

χ (3) =

(8)

where N is the number of molecules (Z) per unit cell volume (V). The HRS first hyperpolarizability (βHRS) for the BRC dissolved in a solvent medium (DMSO) can be calculated by the following expression: βHRS =

⟨βZZZ 2⟩ + ⟨βXZZ 2⟩

(9)

where the X-direction is assumed as the fundamental light beam propagation and polarization is in the Z-direction and βZZZ2 and βXZZ2 are macroscopic averages calculated from the first hyperpolarizability components (βijk) and described in refs 49 and 50. In this case, we adopted the laboratory system of reference by the X, Y, and Z coordinates, and the molecular system of reference by the x, y, and z coordinates. All the quantum molecular calculations were obtained from the Gaussian 0951 output file and converted by the electronic units (esu).

4. RESULTS AND DISCUSSION 4.1. Linear Absorption, First-Order Hyperpolarizability, and Two-Photon Absorption Spectrum of BRC

Figure 2. Evolution of values of the MP2/6-311+G(d) dipole moment of the BRC with the respective iteration numbers. Step 0 indicates the isolated molecule and the other steps indicate the BRC crystal.

Figure 3. Linear absorption of BRC dissolved in DMSO solvent.

Linear and nonlinear electric parameters of the BRC crystal embedded molecules were calculated with the MP2 perturbation theory (electric dipole moment and linear polarizability) and DFT/CAM-B3LYP (second hyperpolarizability), both with the 6-311++G(d,p) set basis. The expressions used in the calculations of the electric dipole moment (μ), the average linear polarizability (α), and the average second hyperpolarizability (γ) were μ = (μx 2 + μy 2 + μz 2 )1/2 α=

γ=

Dissolved in DMSO. Figure 3 displays the absorption spectrum of chalcone dissolved in DMSO solvent at room temperature. The HOMO−LUMO transition is located at ca. 315 nm with a molar absorptivity of approximately 12 400 mol−1 L cm−1. For wavelengths longer than 450 nm up to 1200 nm, the sample appears to be totally transparent. The fluorescence signal was not observed for this compound in solution, indicating that the relaxation of the first singlet excited state to the ground state is nonradiative to the molecule in solution. From the molar absorptivity, ε, it was possible to calculate the transition dipole moment from the ground singlet state to the first singlet excited state, μ⃗01, by using the following relation:52

(4)

αxx + αyy + αzz 3

1 [γ + γyyyy + γzzzz + 2(γxxyy + γxxzz + γyyzz)] 5 xxxx

(5) (6)

The average linear polarizability and the average second hyperpolarizability can be related with the linear refractive index (n) and with the third-order electric susceptibility (χ(3)) of the BRC crystal, through the expressions n2 − 1 4πN = ⟨α⟩ 3 n2 + 2

μ01 =

3·103 ln(10)hc n (2π )3 NAν01 L2

∫ ε( ν ) d ν

(10)

in which c is the speed of light, h is Planck’s constant, NA is Avogadro’s number, and ν01 is the transition frequency between the ground and first excited state. n = 1.42 is the

(7) D

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 6. Experimental HRS signal measured at 1064 nm (∼1.17 eV) square and simplified two-level dispersion model (line).

Figure 4. Two-photon absorption spectrum for BRC dissolved in DMSO solvent (open red circles). The red line is the adjustment obtained with the sum-over-states model. The inset shows typical Zscan signatures as a function of the sample position with respect to the focal point (Z = 0); the solid lines represent the fits obtained with eq 1.

photon absorption but highly allowed by 2PA. At wavelengths close to 500 nm (250 nm), the 2PA cross sections increase to values greater than 17 GM due to either the presence of other TPA allowed excited states energetically close or the resonance enhancement (due to the OPA process). In order to estimate some important photophysical parameters of BRC, such as transition dipole moments of the excited states (|μ⃗ ij|) and the permanent dipole moment difference between the ground and first excited states (|Δμ⃗01| = |μ⃗11 − μ⃗00|), the sum-over-states (SOS) model was employed together with the 2PA experimental results. In Figure 4, the solid red line represents the result obtained with the SOS approach, which is given by54,55

Table 1. Values of Transitions Dipole Moments and Permanent Dipole Moment Difference in Debye Unit μ01

Δμ01

μ12

μ13

7.6

4.3

4.4

4.4

ν2 4 (2π )4 4 L 2 5π (ch) (ν01 − ν)2 + Γ012 ÅÄÅ 2 2 |μ12 |2 |μ01|2 Γ02 ÅÅ |μ01| |Δμ01| Γ01 × ÅÅÅ + ÅÅ (ν − 2ν)2 + Γ 2 (ν02 − 2ν)2 + Γ02 2 01 ÅÇ 01 ÑÉ |μ13 |2 |μ01|2 Γ03 ÑÑÑ ÑÑ + Ñ (ν03 − 2ν)2 + Γ032 ÑÑÑÖ (11) in which ν is the frequency of the laser. The spectroscopic parameters ν0n and Γ0n correspond to the transition frequency and damping constant, respectively, of the 0 → n transition (n = 1, 2, 3). In most of the cases, the electronic transitions for organic molecules at the UV−vis region present a damping constant of about 0.3 eV. As μ01 was calculated from the linear absorption, Δμ01, μ12, and μ13 were determined from eq 11 independently, once the line shapes that describe the electronic transitions are separated. The values calculated are shown in Table 1. It is important to mention that the values reported in Table 1 and also the magnitude of the 2PA cross section, taking into consideration the experimental errors, agree very well with the ones reported for small conjugated chalconebased structures.24 The first-order hyperpolarizability of BRC was obtained by comparing with pNA, both dissolved in DMSO solvent. The quadratic dependence of the scattered signal (532 nm) as a function of the pump intensity (1064 nm) can be seen in the inset of Figure 5. Each quadratic curve represents the signal for a distinct solute concentration. In Figure 5, the main graphic shows the linear dependence of the ratio between HRS signal and square of the pump intensity (normalization) as a function SOS σ2PA (ν ) =

Figure 5. Linear dependence of HRS signal as a function of the molecules concentrations for pNA (squares) and chalcone (circles). Lines represent the linear fits. The inset shows the first hyperpolarizability scattering signals as a function of the pump intensity for five distinct molar concentrations. Lines are the best second-order polynomial fits.

DMSO refractive index at room temperature. L = 3n2/(2n2 + 1) is the Onsager local field factor, which is used to take into account the medium effect.53 The transition dipole moment calculated for the lower transition is 7.6 D (D). The two-photon absorption spectrum of BRC dissolved in DMSO solvent is displayed in Figure 4 (open circles). The inset of this figure shows several Z-scan measurements (symbols) for different wavelengths. A decrease in the transmittance around the focal position (Z = 0) can be seen, indicating the nonlinear effect is taking place. The solid lines represent the adjustment obtained by using eq 1. In the main graphic, it is possible to see two 2PA bands, one located at ca. 630 nm (315 nm) that matches the HOMO− LUMO transition and an extra one at ca. 550 nm (275 nm).The latter describes a state that is not allowed by oneE

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 7. Heating−cooling−reheating DSC and TGA data for BRC. The solid−liquid transition via hot-stage microscopy analysis is also shown.

Figure 8. Thermal ellipsoids drawn at the 50% probability level showing the crystal structure of BRC. The hydrogen atoms are represented by spheres with arbitrary radii.

between static and dynamic β’s considers the resonance enhancement effect due to the optical frequency dispersion. Using the undamped two-level model,56−58 which neglects the optical transition line broadening, it is possible to extrapolate the dynamic hyperpolarizability when the frequency (ν) of the incident light is far from the frequency of the lowest electronic transition (ν01) by the following expression:56

of the molar concentration. The open squares are the signals obtained for pNA for five different molar concentration, and the open circles are the ones measured for the BRC molecule. By using eq 3 and βpNA(1064 nm), it was possible to determine the first-order hyperpolarizability for BRC. The value obtained was of about (11 ± 3) × 10−30 cm5/esu at 1064 nm, 2 times lower than the pNA value. Chalcones derivatives24 have shown low values of β at 1064 nm, almost the same order of pNA. One explanation for this low value at 1064 nm is because the HOMO−LUMO transition is located in the UV region, at approximately 340 nm. Consequently, the resonance effect on the hyper-Rayleigh signal is not so significantly at 532 nm. Static and dynamical first-order hyperpolarizability can be estimated by using a simplified two-level model.56 With this approach, the static first-order hyperpolarizability (β0) can be expressed as β0(0;0,0) =

⃗ |2 |Δμ01 ⃗ | 3 |μ01 2 hν01

β( −2ν ;ν ,ν) =

ν014 (ν012 − 4ν 2)(ν012 − ν 2)

β0

(13)

As the result of the dispersion term in eq 13, the dynamical hyperpolarizability at 1064 nm is about 15.6 × 10−30 cm5/esu, 1.6 times greater than the one obtained with the HRS technique. Taking into account the experimental errors in the HRS measurement and in the determination of the dipole moments, it is possible to say that the results obtained by using two different approaches are in reasonable agreement. To illustrate the dispersion model on the hyperpolarizability, eq 13 was plotted and compared with the β-value obtained experimentally, as shown in Figure 6. 4.2. Solid State Characterization. 4.2.1. Thermal Analysis. With a view to fully characterize the crystalline structure of BRC, DSC/TGA and hot-stage microscopy analysis have been performed on this phase and the results

(12)

By using μ⃗01 and Δμ⃗01 obtained, respectively, from the linear absorption and 2PA spectrum, and considering hν01 the energy of the first electronic transition at 3.96 eV, β0(0) was estimated to be approximately 9 × 10−30 cm5/esu. The relationship F

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

with unit cell metrics a = 8.2534(9) Å, b = 15.2608(13) Å, c = 16.5301(17) Å, α = 105.115(8)°, β = 103.538(9)°, and γ = 96.788(8)°. Each motif is composed of two independent molecules here named I and II. Thermal ellipsoids of BRC are presented in Figure 8, followed by the main crystallographic data in Table 2. There are no major differences between molecular structures of both conformers I and II, as seen in Figure 9. It is possible to see in this figure that a slight clockwise twist of the aromatic ring B converts I into II. As result of this difference, the bromine atom has +syn-clinal and −syn-periplanar conformations regarding the carbonyl group in I and II, respectively. For I, the dihedral angle C8−C9−C10−C15 measures ω1 = −179.07°, while this same dihedral angle is ω2 = −179.07° in II. Both I and II show planarity deviations in chalcone moieties, as confirmed by the angle formed between aromatic rings A and B in I (∠1 = 31.29°) and II(∠2 = 41.18°). Figure 10a shows the front view of the BRC crystal packing, which is stabilized by C−H···Br and C−H···O interactions. I and II are assembled in a chain through C39−H39···Br1 interaction (Figure 10b). In addition, it is observed another chain composed only of molecules I assembled by C8−H8··· O21 and C4−H4···O1 interactions. These interactions pack in C(8) and C(4) motifs, respectively, and run along [100] (Figure 10c). The last intermolecular interactions, C26−H26··· O5 and C25−H25···O4, pack II in a chain running along [100] with C(11) and C(8) motifs, respectively (Figure 10d). Further details about geometric parameters of these intermolecular interactions are given in Table 3. 4.2.3. Photoluminescence Properties. Finally, in Figure 11 is shown the excitation and emission photoluminescence spectra of the BRC crystal. The excitation spectrum was collected by monitoring the emission centered at 430 nm, and the emission spectrum was detected under excitation at 365 nm. This excitation band is well correlated to the broad absorption band centered around 315 nm (Figure 11) (extending from 270 to 420 nm) due to the HOMO− LUMO transition. As noted in Figure 11, the emission covers the full visible electromagnetic range with the highest emission intensity in the blue region, related to the charge transfer between aromatic rings (B or C identified in Figure 8) and 0 = S = 0/NH groups. The energy difference between the excitation and emission peaks is due to losses in the crystal structure activating vibrational modes of organic molecules and structural groups. 4.2.4. BRC Crystal Computational Calculation Results. In Table 4 the MP2/6-311++G(d,p) results for the dipole moment (μi (i = x, y, z) and μ) and the linear polarizability (αij(i, j = x, y, z) and average value α) for the BRC embedded molecule (crystal) are listed. The calculated value of the total dipole moment was 12.95 D, and the highest value of the dipole moment component was μx = −11.19 D. As can be seen, the average linear polarizability presents the value, α = 42.59 × 10−24 esu, and the diagonal components are those that present the highest values. Also, in Table 4 the DFT/CAM-B3LYP/6311++G(d,p) results for the second hyperpolarizability components (γijkl(i,j,k,l=x,y,z)) and average value (γ), are shown. The component with highest value is γyyyy = 83.97 × 10−36 esu and the average second hyperpolarizability value is γ = 66.53 × 10−36 esu. The DFT/CAM-B3LYP/6-311++G(d,p) dynamic results for the average linear polarizability, α(−ω,ω), and the average second hyperpolarizabilities, γ(−ν;ν,0,0) and γ(−2ν;ν,ν,0), as

Table 2. Crystal Data and Structure Refinement for BRC empirical formula formula weight temperature (K) crystal system space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) volume (Å3) Z ρcalc (g/cm3) μ (mm−1) F(000) crystal size (mm3) radiation 2Θ range for data collection (deg) index ranges no. of reflections collected no. of independent reflections data/restraints/parameters goodness-of-fit on F2 final R indexes [I ≥ 2σ(I)] final R indexes [all data] largest diff peak per hole (e/Å3)

C21H16BrNO3S 442.32 293(2) triclinic P1̅ 8.2534(9) 15.2608(13) 16.5301(17) 105.115(8) 103.538(9) 96.788(8) 1918.2(3) 4 1.532 2.273 896.0 0.351 × 0.274 × 0.161 Mo Kα (λ = 0.710 73) 4.41−49.998 −9 ≤ h ≤ 9, −18 ≤ k ≤ 18, −19 ≤ l ≤ 19 27 520 6725 [Rint = 0.0423, Rσ = 0.0531] 6725/0/487 1.034 R1 = 0.0447, wR2 = 0.0955 R1 = 0.0899, wR2 = 0.1098 +0.41/−0.43

Figure 9. Overlapping of molecules I (blue) and II (gray).

are presented in Figure 7. The DSC measurement shows two events (blue trace, Figure 7), the first endothermic and a second exothermic, centered at 153.97 and 289.20 °C, respectively. No weight loss is detected in the TGA (black trace, Figure 7) curve until near 270 °C, the beginning of exothermic event. The endothermic event is associated with a solid−liquid phase transition, also observed in the hot stage. This transformation is irreversible, since no corresponding event is occurring during the cooling−reheating cycles of DSC, probably due to the exothermic event that can be associated with a total degradation of BRC. Further, the event at 153.97 °C in the DSC curve corresponds to melting of the sample and was confirmed by the hot-stage analysis. No habit or color change is observed between 25 and 140 °C, the solid−liquid phase transformation is detected in hot-stage photography near 150 °C, and complete melting of the sample is achieved at 157.3 °C. 4.2.2. Structural Properties. BRC is a sulfonamide chalcone with a bromine atom para-bonded to the aromatic ring A (see Figure 8 for numbering scheme). It was crystallized in the triclinic crystal system and P1̅ centrosymmetric space group, G

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 10. Front view of the BRC crystal packing (a) stabilized by C−H···Br (b) and C−H···O [(c) and (d)] interactions.

Table 3. Geometric Parameters of Intermolecular Interactions Observed in the Crystal Packing of BRC D

H

A

d(D−H) (Å)

d(H−A) (Å)

d(D−A) (Å)

D−H−A (deg)

C8 C26 C4 C25 C39

H8 H26 H4 H25 H39

O2a O5a O1a O4a Br1b

0.93 0.93 0.93 0.93 0.93

2.66 2.69 2.67 2.50 2.98

3.401(4) 3.570(4) 3.480(5) 3.356(4) 3.909(4)

137.0 157.9 146.0 152.8 173.5

Table 4. MP2/6-311++G(d,p) Results for the Dipole Moment (D) and Linear Polarizability (10−24 esu) and DFT/CAM-B3LYP Results for BRC Average Second Hyperpolarizability (10−36 esu) μx μy μz μ

1 + X, +Y, +Z. b−1 + X, −1 + Y, +Z.

a

−11.19 6.18 2.02 12.95

αxx αxy αyy αxz αyz αzz α

44.23 7.16 49.06 0.59 7.26 34.50 42.59

γxxxx γyyyy γzzzz γxxyy γyyzz γxxzz γ

56.14 83.97 30.16 47.00 20.04 14.14 66.53

present a monotonic increase with the increasing frequency, and from ν = 0.00 au to ν = 0.10 au, the values of these functions are enhanced by 15.4% and 210.4%, respectively. In Figure 12 also the dispersion of the direct-current second harmonic generation (dc-SHG) function, γ(−2ν;ν,ν,0) is shown. As can be observed, a monotonic increase of this function occurs in the region 0 < ν < 0.07 au; for ν > 0.07 au, the function presents discontinuities at the resonant frequencies. From eq 4 the static (dynamic at λ = 1064 nm) linear refractive index estimated for BRC crystal was n = 1.65 (1.68). In order to make an appropriate estimation of the χ(3), we have used an approximation to the frequency-dependent second hyperpolarizability, (γ(−ν;ν,ν,−ν)), associated with the nonlinear optical process59 of the intensity dependent refractive index (IDRI) from dc-Kerr result, γ(−ν;ν,0,0). Following a previous work,60 for small frequencies,61 we have

Figure 11. Excitation and emission photoluminescence spectra of BRC crystals.

a function of the electric field frequencies (ν (au)) for the BRC embedded molecules are shown in the Figure 12. As can be seen the dispersion relations of α(−ν,ν) and γ(−ν;ν,0,0) H

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 12. Dynamic evolution of the calculated values for average linear polarizability α(−ν;ν) (10−24 esu) (a), average second hyperpolarizability γ(−ν;ν,0,0) (10−36 esu) (b), and the average second hyperpolarizability γ(−2ν;ν,ν,0) (10−36 esu) (c) for BRC embedded molecules.

average first hyperpolarizability of BRC molecules dissolved in DMSO, in which both approaches demonstrated a good agreement for β. The 2PA experimental spectrum, measured by using the Z-scan technique, revealed a considerable effect on the therapeutic window region. Also, a supramolecular approach was employed in combination with an iterative scheme of electrostatic polarization to simulate the nonlinear optical properties of BRC crystal. The calculated value of the third-order macroscopic susceptibility obtained for the BRC crystal was 10 times higher than the values measured for three chalcone derivatives using the Z-scan technique, reported experimentally in the literature.62 The frequency dispersion effects on these parameters also was studied and the dc-Kerr second hyperpolarizability (γ(−ν;ν,0,0)) shown an increasing of 210.4% from the static value (ν = 0) to ν = 0.10 au. The BRC crystal macroscopic quantities as the linear refractive index (n) and the third-order nonlinear susceptibility (χ(3)) were estimated from the Clausius−Mossoti equation and from the frequency-dependent second hyperpolarizability associated with the nonlinear optical process of the intensity dependent refractive index (IDRI). The obtained values (at λ = 1064 nm) were n = 1.68 and χ(3)(−ν;ν,ν,−ν) = 2.57 × 10−21 m2/V2. Simulated χ(3)value shows to be greater than the ones reported for other chalcone derivatives, which should qualify BRC crystal as a potential candidate for application in nonlinear optical devices.

Table 5. Third-Order Optical Nonlinearities of BRC Compared with Some Organic Nonlinear Crystals sample

n0

χ(3)(−ν;ν,ν,−ν) (10−20 m2/V2)

BRC (2E)-1-(4-bromophenyl)-3-[4(-methylsulfanyl) phenyl]prop-2-en-1-one (4Br4MSP)62 (2E)-1-(3-bromophenyl)-3-[4-(methylsulfanyl) phenyl]prop-2-en-1-one (3Br4MSP)62 (2E)-3[4-(methylsulfanyl)phenyl]-1-(4nitrophenyl)prop-2-en-1-one (4N4MSP)62

1.68 1.363

0.257 0.023

1.365

0.0199

1.36

0.0237

Table 6. Static and Dynamic (λ = 1064 nm) r®esults for the BRC HRS First Hyperpolarizability (10−30 cm5/esu) two-level model DFT/CAM-B3LYP/HRS HRS experimental

βBRC(0,0,0)

βBRC(−2ν;ν,ν)

9 12.41

15.6 16.06 9.8 ± 0.9

⟨γ( −ν ;ν ,ν ,−ν)⟩ ≅ 2⟨γ( −ν ;ν ,0,0)⟩ − ⟨γ(0;0,0,0)⟩

Therefore, the calculated value for γ(−ν;ν,ν,−ν) is 88.08 × 10−36 esu at ν = 0.0428 au (λ = 1064 nm) and the BRC crystal third-order susceptibility is χ(3)(−ν;ν,ν,−ν) = 2.57 × 10−21 m2/V2. The BRC χ(3)-value is greater, about more than 10 times the values obtained experimentally for three 4methylsulfanyl chalcone derivatives obtained experimentally by D’Silva et al.;62 see Table 5. 4.2.5. BRC Molecules Dissolved in DMSO Computational Calculation Results. Table 6 show the results for the βBRC static and dynamic (ν = 0.0428 au) obtained from eqs 12 and 13 corresponding to the simplified phenomenological two-level model and from the DFT theoretical results using eq 9. As shown in Table 6, both results, the simplified two-level model and the theoretical (DFT/CAM-B3LYP/cc-pVTZ), are close to the one experimentally measured, taking into consideration the experimental errors. Therefore, we can conclude that the DFT model using CAM-B3LYP/cc-pVTZ is convenient to treat the optical properties of the BRC.



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. *E-mail: [email protected]. ORCID

Jean M. F. Custodio: 0000-0002-2295-2112 Leonardo De Boni: 0000-0002-1875-1852 Caridad N. Perez: 0000-0003-0586-0774 Hamilton B. Napolitano: 0000-0002-6047-9995 Notes

The authors declare no competing financial interest.



5. CONCLUSION In this work, we characterized a novel bromine chalcone (E)-3(2-bromophenyl)-1-(2-((phenylsulfonyl)amine)phenyl)prop2-en-1-one (BRC) from single-crystal X-ray diffraction and spectroscopy analyzes. The DSC/TGA and hot-stage microscopy analysis were performed, and the results show the crystal presents a good stability in relation to temperature variation. We also investigated, theoretically and experimentally, the

ACKNOWLEDGMENTS The authors gratefully acknowledge the financial support of the ́ Conselho Nacional de Desenvolvimento Cientifico e Tecnológico (CNPq), Coordenaçaõ de Aperfeiçoamento de Pessoal ́ Superior (CAPES), Fundaçaõ de Amparo à Pesquisa de Nivel do Estado de São Paulo (FAPESP; Grant 2016/20886-1), and Fundação de Amparo à Pesquisa do Estado de Goiás I

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

(17) Vogel, S.; Ohmayer, S.; Brunner, G.; Heilmann, J. Natural and Non-Natural Prenylated Chalcones: Synthesis, Cytotoxicity and AntiOxidative Activity. Bioorg. Med. Chem. 2008, 16, 4286−4293. (18) Wang, G.; Xue, Y.; An, L.; Zheng, Y.; Dou, Y.; Zhang, L.; Liu, Y. Theoretical Study on the Structural and Antioxidant Properties of Some Recently Synthesised 2,4,5-Trimethoxy Chalcones. Food Chem. 2015, 171, 89−97. (19) Alvim, H. G. O.; Fagg, E. L.; de Oliveira, A. L.; de Oliveira, H. C. B.; Freitas, S. M.; Xavier, M.-A. E.; Soares, T. A.; Gomes, A. F.; Gozzo, F. C.; Silva, W. A.; et al. Probing Deep into the Interaction of a Fluorescent Chalcone Derivative and Bovine Serum Albumin (BSA): An Experimental and Computational Study. Org. Biomol. Chem. 2013, 11, 4764. (20) Kim, J. H.; Ryu, H. W.; Shim, J. H.; Park, K. H.; Withers, S. G. Development of New and Selective Trypanosoma Cruzi TransSialidase Inhibitors from Sulfonamide Chalcones and Their Derivatives. ChemBioChem 2009, 10, 2475−2479. (21) Sheik-Bahae, M.; Said, A. A.; Wei, T.-H.; Hagan, D. J.; Stryland, E. W. Van. Sensitive Measurement of Optical Nonlinearities Using a Single Beam. IEEE J. Quantum Electron. 1990, 26, 760−769. (22) Göppert-Mayer, M. Elementary Processes with Two Quantum Transitions. Ann. Phys. 2009, 18, 466−479. (23) Clays, K.; Persoons, A. Hyper-Rayleigh Scattering in Solution. Phys. Rev. Lett. 1991, 66, 2980−2983. (24) Abegão, L. M. G.; Fonseca, R. D.; Santos, F. A.; Souza, G. B.; Barreiros, A. L. B. S.; Barreiros, M. L.; Alencar, M. A. R. C.; Mendonça, C. R.; Silva, D. L.; De Boni, L.; et al. Second- and ThirdOrder Nonlinear Optical Properties of Unsubstituted and MonoSubstituted Chalcones. Chem. Phys. Lett. 2016, 648, 91−96. (25) Franzen, P. L.; Misoguti, L.; Zilio, S. C. Hyper-Rayleigh Scattering with Picosecond Pulse Trains. Appl. Opt. 2008, 47, 1443. (26) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H. OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42, 339−341. (27) Sheldrick, G. M. A. Short History of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112−122. (28) Sheldrick, G. M. Crystal Structure Refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (29) Spek, A. L. Single-Crystal Structure Validation with the Program PLATON. J. Appl. Crystallogr. 2003, 36, 7−13. (30) Nardelli, M. PARST 95 an Update to PARST: A System of Fortran Routines for Calculating Molecular Structure Parameters from the Results of Crystal Structure Analyses. J. Appl. Crystallogr. 1995, 28, 659. (31) Macrae, C. F.; Edgington, P. R.; McCabe, P.; Pidcock, E.; Shields, G. P.; Taylor, R.; Towler, M.; van de Streek, J. Mercury: Visualization and Analysis of Crystal Structures. J. Appl. Crystallogr. 2006, 39, 453−457. (32) Farrugia, L. J. WinGX and ORTEP for Windows: An Update. J. Appl. Crystallogr. 2012, 45, 849−854. (33) Pettersen, E. F.; Goddard, T. D.; Huang, C. C.; Couch, G. S.; Greenblatt, D. M.; Meng, E. C.; Ferrin, T. E. UCSF Chimera - A Visualization System for Exploratory Research and Analysis. J. Comput. Chem. 2004, 25, 1605−1612. (34) 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 Third-Order Nonlinear Optical Properties of a New Neolignane Analogue. RSC Adv. 2016, 6, 79215−79227. (35) Santos, O. L.; Sabino, J. R.; Georg, H. C.; Fonseca, T. L.; Castro, M. A. Electric Properties of the 3-Methyl-4-Nitropyridine-1Oxyde (POM) Molecules in Solid Phase: A Theoretical Study Including Environment Polarization Effect. Chem. Phys. Lett. 2017, 669, 176−180. (36) 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.

(FAPEG). The authors also thank Leonardo Rodrigues de Almeida for accessing the facility of thermal analysis at LaMuCrEs Multiuser Structural Crystallography Laboratory. Research was done with the support of the High Performance Computing Center at the Universidade Estadual de Goiás (UEG).



REFERENCES

(1) Zhang, L.; Pavlica, E.; Zhong, X.; Liscio, F.; Li, S.; Bratina, G.; Orgiu, E.; Samorì, P. Fast-Response Photonic Device Based on Organic-Crystal Heterojunctions Assembled into a Vertical-Yet-Open Asymmetric Architecture. Adv. Mater. 2017, 29, 1605760. (2) Park, K. S.; Baek, J.; Park, Y.; Lee, L.; Hyon, J.; Koo Lee, Y.-E.; Shrestha, N. K.; Kang, Y.; Sung, M. M. Heterogeneous Monolithic Integration of Single-Crystal Organic Materials. Adv. Mater. 2017, 29, 1603285. (3) Chen, H.; Liu, H.; Zhang, Z.; Hu, K.; Fang, X. Nanostructured Photodetectors: From Ultraviolet to Terahertz. Adv. Mater. 2016, 28, 403−433. (4) de Oliveira, G. R.; de Oliveira, H. C. B.; Silva, W. A.; da Silva, V. H. C.; Sabino, J. R.; Martins, F. T. Structure and Theoretical Approaches to a Chalcone Derivative. Struct. Chem. 2012, 23, 1667− 1676. (5) Rozmer, Z.; Perjési, P. Naturally Occurring Chalcones and Their Biological Activities. Phytochem. Rev. 2016, 15, 87−120. (6) Bukhari, S. N.; Jasamai, M.; Jantan, I. Synthesis and Biological Evaluation of Chalcone Derivatives (Mini Review). Mini-Rev. Med. Chem. 2012, 12.1394 (7) Bukhari, S. N. A.; Jasamai, M.; Jantan, I.; Ahmad, W. Review of Methods and Various Catalyst Used for Chalcones Synthesis. MiniRev. Org. Chem. 2013, 10, 73−83. (8) Ritter, M.; Martins, R. M.; Rosa, S. A.; Malavolta, J. L.; Lund, R. G.; Floresd, A. F. C.; Pereira, C. M. P. Green Synthesis of Chalcones and Microbiological Evaluation. J. Braz. Chem. Soc. 2015, 26, 1201− 1210. (9) Wang, L.; Chen, G.; Lu, X.; Wang, S.; Han, S.; Li, Y.; Ping, G.; Jiang, X.; Li, H.; Yang, J.; et al. Novel Chalcone Derivatives as Hypoxia-Inducible Factor (HIF)-1 Inhibitor: Synthesis, Anti-Invasive and Anti-Angiogenic Properties. Eur. J. Med. Chem. 2015, 89, 88−97. (10) Kamal, A.; Ramakrishna, G.; Raju, P.; Viswanath, A.; Janaki Ramaiah, M.; Balakishan, G.; Pal-Bhadra, M. Synthesis and AntiCancer Activity of Chalcone Linked Imidazolones. Bioorg. Med. Chem. Lett. 2010, 20, 4865−4869. (11) Markovic, V.; Debeljak, N.; Stanojkovi, T.; Kolundija, B.; Sladi, D.; Vuji, M.; Janovi, B.; Tani, N.; Perovi, M.; Tei, V.; et al. Anthraquinone-Chalcone Hybrids: Synthesis, Preliminary Antiproliferative Evaluation and DNA-Interaction Studies. Eur. J. Med. Chem. 2015, 89, 401−410. (12) Venkateswararao, E.; Sharma, V. K.; Yun, J.; Kim, Y.; Jung, S. Bioorganic & Medicinal Chemistry Anti-Proliferative Effect of Chalcone Derivatives through Inactivation of NF- j B in Human Cancer Cells. Bioorg. Med. Chem. 2014, 22, 3386−3392. (13) Pingaew, R.; Saekee, A.; Mandi, P.; Nantasenamat, C.; Prachayasittikul, S.; Ruchirawat, S.; Prachayasittikul, V. Synthesis, Biological Evaluation and Molecular Docking of Novel ChalconeCoumarin Hybrids as Anticancer and Antimalarial Agents. Eur. J. Med. Chem. 2014, 85, 65−76. (14) Dimmock, J. R.; Elias, D. W.; Beazely, M. A.; Kandepu, N. M. Bioactivities of Chalcones. Curr. Med. Chem. 1999, 6, 1125−1149. (15) Go, M. L.; Wu, X.; Liu, X. L. Chalcones: An Update on Cytotoxic and Chemoprotective Properties. Curr. Med. Chem. 2005, 12, 483−499. (16) Silva, W. a.; Andrade, C. K. Z.; Napolitano, H. B.; Vencato, I.; Lariucci, C.; De Castro, M. M. R. C.; Camargo, A. J. Biological and Structure-Activity Evaluation of Chalcone Derivatives against Bacteria and Fungi. J. Braz. Chem. Soc. 2013, 24, 133−144. J

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C (37) 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, 29, 168−179. (38) 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. (39) Rodrigues, R. F. N.; Almeida, L. R.; Santos, F. G. dos; Carvalho, P. S., Jr.; 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-Acylhydrazide Derivative. PLoS One 2017, 12, No. e0175859. (40) 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. (41) Baseia, B.; Osório, F.; Lima, L.; Valverde, C. Effects of Changing Substituents on the Non-Linear Optical Properties of Two Coumarin Derivatives. Crystals 2017, 7, 158. (42) 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. (43) 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. (44) 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. (45) 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. (46) Valverde, C.; Vaz, W. F.; Custodio, J. M. F.; Duarte, V. S.; Carvalho, P. S., Jr.; 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. (47) Valverde, C.; Osorio, A. P. F.; Fonseca, T. L.; Baseia, B. DFT Study of Third-Order Nonlinear Susceptibility of a Chalcone Crystal. Chem. Phys. Lett. 2018, 706, 170−174. (48) 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. (49) Bersohn, R.; Yoh-Han, P. A. O.; Frisch, H. L. Double-Quantum Light Scattering by Molecules. J. Chem. Phys. 1966, 45, 3184−3198. (50) Verbiest, T.; Clays, K.; Rodriguez, V. Second-Order Nonlinear Optical Characterization Techniques An Introduction; CRC Press, 2009. (51) 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 D.01; Gaussian, Inc.: Wallingford CT, 2009. (52) Fonseca, R. D.; Vivas, M. G.; Silva, D. L.; Eucat, G.; Bretonnière, Y.; Andraud, C.; De Boni, L.; Mendonça, C. R. FirstOrder Hyperpolarizability of Triphenylamine Derivatives Containing Cyanopyridine: Molecular Branching Effect. J. Phys. Chem. C 2018, 122, 1770−1778. (53) Onsager, L. Electric Moments of Molecules in Liquids. J. Am. Chem. Soc. 1936, 58, 1486−1493. (54) Kamada, K.; Ohta, K.; Iwase, Y.; Kondo, K. Two-Photon Absorption Properties of Symmetric Substituted Diacetylene: Drastic

Enhancement of the Cross Section near the One-Photon Absorption Peak. Chem. Phys. Lett. 2003, 372, 386−393. (55) Neves, U. M.; De Boni, L.; Ye, Z.; Bu, X. R.; Mendonça, C. R. Two-Photon Absorption Spectra of Salen Dye Complexes with Azo Dyes. Chem. Phys. Lett. 2007, 441, 221−225. (56) Oudar, J. L. Optical Nonlinearities of Conjugated Molecules. Stilbene Derivatives and Highly Polar Aromatic Compounds. J. Chem. Phys. 1977, 67, 446−457. (57) Orr, B. J.; Ward, J. F. Perturbation Theory of the Non-Linear Optical Polarization of an Isolated System. Mol. Phys. 1971, 20, 513− 526. (58) Campo, J.; Wenseleers, W.; Goovaerts, E.; Szablewski, M.; Cross, G. H. Accurate Determination and Modeling of the Dispersion of the First Hyperpolarizability of an Efficient Zwitterionic Nonlinear Optical Chromophore by Tunable Wavelength Hyper-Rayleigh Scattering. J. Phys. Chem. C 2008, 112, 287−296. (59) de Araújo, C. B.; Gomes, A. S. L.; Boudebs, G. Techniques for Nonlinear Optical Characterization of Materials: A Review. Rep. Prog. Phys. 2016, 79, 036401. (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. (61) Bishop, D. M.; De Kee, D. W. The Frequency Dependence of Nonlinear Optical Processes. J. Chem. Phys. 1996, 104, 9876−9887. (62) D’Silva, E. D.; Podagatlapalli, G. K.; Venugopal Rao, S.; Dharmaprakash, S. M. Study on Third-Order Nonlinear Optical Properties of 4-Methylsulfanyl Chalcone Derivatives Using Picosecond Pulses. Mater. Res. Bull. 2012, 47, 3552−3557.

K

DOI: 10.1021/acs.jpcc.9b01063 J. Phys. Chem. C XXXX, XXX, XXX−XXX