Impact of Homocyclic and Heterocyclic Rings of Chalcones on Charge

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Chemical Physics Impact 8 (2024) 100565

Contents lists available at ScienceDirect

Chemical Physics Impact


journal homepage: www.sciencedirect.com/journal/chemical-physics-impact

Full Length Article

Impact of Homocyclic and Heterocyclic Rings of Chalcones on Charge


Transfer Behaviour: A Nonlinear Optical Study
Nur Aisyah Mohamad Daud a, Qin Ai Wong a, *, Bi Sheng Ooi a, Ching Kheng Quah a, *,
Farah Diana Ramzi a, Yip-Foo Win b, Parutagouda Shankaragouda Patil c
a
X-ray Crystallography Unit, School of Physics, Universiti Sains Malaysia, Penang 11800, Malaysia
b
Department of Chemical Science, Faculty of Science, Universiti Tunku Abdul Rahman, Perak Campus, Jalan Universiti, Bandar Barat, 31900 Kampar, Perak, Malaysia
c
Department of Physics, B.L.D.E. Association’s S.B. Arts and K.C.P. Science College, Vijayapura 586103, Karnataka, India

A R T I C L E I N F O A B S T R A C T

Keywords: This paper presents the effects of incorporating thienyl and phenyl rings as the terminal ring in the cinnamoyl
Nonlinear Optics counterparts of chalcones with donor-acceptor-donor (D-A-D) push-pull setting, on their photophysical re­
Density Functional Theory sponses. Two fluoro-methoxy substituted chalcone derivatives with distinct cinnamoyl systems, 4-methylthio­
Chalcones
phene-2-carbaldehyde (2F4MA48) and 2,4-dimethylbenzaldehyde (2F4MA96), were synthesized.
Intramolecular Charge Transfer
Z-scan
Characterization of these chalcone derivatives was performed using nuclear magnetic resonance (NMR), Fourier-
transform infrared (FTIR) and single-crystal X-ray diffraction (SCXRD). Comparative studies on the linear and
nonlinear optical responses were carried out through both experimental analyses and theoretical density func­
tional theory (DFT) calculations. The electronic absorption spectra computed using DFT were compared with the
UV-Vis spectra to elucidate the intramolecular charge transfer mechanism via the hole-electron distribution and
interfragmentary charge transfer analyses. Molecular properties were examined through electrostatic potential
(ESP) maps, π-electron localization function (ELF-π), and frontier molecular orbitals (FMOs). The incorporation
of thienyl ring is found to give a stronger push-pull effect and greater electron delocalization. Experimental
nonlinear optical responses were investigated using Z-scan technique with a continuous-wave incident laser at a
wavelength of 532 nm. Both compounds showed two-photon-induced reverse saturation absorption (RSA) and
self-defocusing responses. It was observed that the thienyl-based chalcone, exhibited greater optical nonlinearity
in terms of nonlinear absorption (NLA), nonlinear refraction (NLR), and optical limiting. The experimental third-
order nonlinear optical susceptibility (χ (3) ) was determined to be 8.04 × 10− 8 esu for 2F4MA48 and 1.08 × 10− 8
esu for 2F4MA96. The onset limiting threshold (FS) was found to be 2.63 kJ cm− 2 for 2F4MA48 and 5.23 kJ
cm− 2 for 2F4MA96. The steep optical limiting traces of 2F4MA48 show superior performance compared to
2F4MA96.

1. Introduction devices hinges on the understanding and utilization of NLO, it is


crucial to design materials that exhibit strong optical nonlinearity. For
Nonlinear optics (NLO) is a phenomenon that arises from the that reason, structural-property relationships of novel NLO materials are
nonlinear relationship between the polarization of the materials and the being extensively explored in both organic and inorganic materials
applied electric field, leading to changes in frequencies and phases of the [5–11]. Inorganic materials, known for high thermal stability and good
incident light [1]. NLO finds applications spanning diverse scientific and mechanical strength, can be brittle due to their high density [12–14].
technological domains including optical limiting, optical signal pro­ This brittleness limits the tunability of inorganic NLO materials to meet
cessing, optical communication systems, and nonlinear imaging. Driven specific requirements. In contrast, low-density organic materials offer
by the need for faster computational speeds and broader bandwidth, greater flexibility, enabling fine-tuning of their NLO responses [15,16].
achievable by optical signals, there has been a substantial increase in Furthermore, organic materials are known for their relatively simple
demand for NLO materials [2–4]. Since the development of all-optical synthesis process, ease of structural modification, and cost-effectiveness

* Corresponding author.
E-mail addresses: qinai940224@gmail.com (Q.A. Wong), ckquah@usm.my (C.K. Quah).

https://doi.org/10.1016/j.chphi.2024.100565
Received 8 December 2023; Received in revised form 15 February 2024; Accepted 10 March 2024
Available online 15 March 2024
2667-0224/© 2024 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY-NC license (http://creativecommons.org/licenses/by-
nc/4.0/).
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

for mass production [17–20]. In addition to these practical advantages, water, filtered, and let to dry. Single crystals of 2F4MA48 and 2F4MA96
organic compounds also exhibit ultra-fast response, great nonlinear were grown through slow evaporation technique. The synthesis path­
refraction (NLR), excellent nonlinear absorption (NLA), and high laser ways are shown in Fig. 1.
damage threshold [21,22]. (E)-1-(2-fluoro-4-methoxyphenyl)-3-(4-methylthiophen-2-yl)
Chalcone is a naturally occurring compound in plants, characterized prop-2-en-1-one (2F4MA48). Solvent for growing single crystal:
by general structure consisting of two aromatic rings linked by an acetone; yield 78%; 1H NMR (400 MHz, CDCl3, δ, ppm): 7.87–7.80 (m,
aliphatic three-carbon chain of α,β-unsaturated carbonyl system [23, 2H, 5CH and 9CH), 7.23–7.19 (d, 1H, J = 16.0 Hz, 8CH), 7.14 (s, 1H,
13
24]. The presence of alternating single and double bonds facilitates the C), 6.97 (s, 1H, 11CH), 6.77–6.75 (d, 1H, J = 8.0 Hz, 4CH), 6.64–6.62
free movement of electrons across the molecules through overlapping (d, 1H, J = 8.0 Hz, 2CH), 3.85 (s, 3H, 14CH), 2.24 (s, 3H, 15CH); 13C NMR
π-orbitals. This electron delocalization improves the NLO response of the (100 MHz, CDCl3, δ, ppm): 186.70, 186.66 (C7), 164.61, 164.49 (C3),
materials, making chalcone derivatives promising NLO materials. The 164.36, 161.83, (C1), 140.26 (C10), 139.07 (C12), 136.65 (C9), 134.02
distinctive push-pull conjugated structure of chalcone derivatives adds (C13), 132.70, 132.66 (C5), 124.59 (C11), 124.08, 123.99 (C8), 119.79,
additional aspect that has garnered significant attention among re­ 119.66 (C6), 110.85, 110.83 (C4), 102.01, 101.74 (C2), 55.94 (C14),
searchers for their multifaceted properties and wide-ranging applica­ 15.64 (C15); FT-IR (ATR (solid), ν, cm− 1): 3100 (sp2 C–H), 3012,2945,
tions across fields such as biology, pharmaceuticals, agriculture, and 2842 (sp3 C–H), 2969, 2924, 2861 (sp3 C–H), 1647 (C– –O), 1578 (En,
nonlinear optics [25,26]. The introduction of specific electron-donating C–
–C), 1611, 1418 (Ar1, C– –C), 1541, 1438 (Ar2, C– –C), 977 (C–H),
(D) or electron-withdrawing (A) groups in D-π-D, D-π-A, D-A-D, or D-A-A 1227, 1042 (C–O), 1098 (C–F), 868, 819 (Ph, C–H), 849, 738 (Th,
arrangements induces molecular polarization, thereby enhancing their C–H), 839, 656 (C–S).
NLO capabilities [27–30]. This flexibility allows for precise fine-tuning (E)-3-(2,4-dimethylphenyl)-1-(2-fluoro-4-methoxyphenyl)prop-
of the push-pull effects in chalcone derivatives to meet specific appli­ 2-en-1-one (2F4MA96). Solvent for growing single crystal: acetone;
cation requirements. While the majority of studies on chalcone de­ yield 84%; 1H NMR (400 MHz, CDCl3, δ, ppm): 8.08–8.04 (d, 1H, J =
rivatives are centred around structures based on the benzene unit [31], 16.0 Hz, 9CH), 7.90–7.86 (t, 1H, J = 8.0 Hz, 5CH), 7.60–7.58 (d, 1H, J =
such as anthracene [32,33], pyrene [34,35], phenanthrene [36], or 8.0 Hz, 15CH), 7.37–7.33 (d, 1H, J = 16.0 Hz, 8CH), 7.04–7.03 (m, 2H,
12
naphthalene [37], there has been only occasional exploration of het­ CH and 14CH), 6.79–6.77 (d, 1H, J = 8.0 Hz, 4CH), 6.65–6.63 (d, 1H, J
erocyclic derivatives [38,39]. However, there have been reports in = 8.0 Hz, 2CH), 3.86 (s, 3H, 16CH), 2.44 (s, 3H, 17CH), 2.33 (s, 3H,
18
which replacing phenyl rings with thienyl rings has led to favourable CH); 13C NMR (100 MHz, CDCl3, δ, ppm): 187.39, 187.36 (C7), 164.59,
optoelectronic properties [40–43]. The electron-rich sulphur atom in the 164.47 (C3), 164.40, 161.87 (C1), 141.61 (C9), 140.68 (C11), 138.50
thiophene unit increases polarizability, thus enhancing the NLO (C13), 132.78, 132.74 (C5), 131.77 (C12), 131.16 (C10), 127.24 (C15),
response of the compounds [44,45]. Nevertheless, at the time of writing, 126.69 (C14), 125.70, 125.62 (C8), 120.00, 119.88 (C6), 110.85, 110.83
the Cambridge Structural Database (CSD) has recorded 5055 (C4), 102.00, 101.73 (C2), 55.93 (C16), 21.46 (C18), 19.90 (C17); FT-IR
benzene-based chalcones, compared to only 514 thienyl-based chalcone, (ATR (solid), ν, cm− 1): 3085 (sp2 C− H), 2986, 2921, 2849 (sp3 C–H),
according to the CCDC ConQuest search results (Version 5.44, June 2968, 2953 (sp3 C–H), 1652 (C– –O), 1587 (En, C– –C), 1610, 1498 (Ar1,
2023). C–
–C), 1610, 1498 (Ar2, C– –C), 975 (C–H), 1234, 1041 (C–O), 1094
In this view, a thienyl based chalcone, (E)-1-(2-fluoro-4-methox­ (C–F), 873, 857, 822, 802 (Ph, C–H).
yphenyl)-3-(4-methylthiophen-2-yl)prop-2-en-1-one (2F4MA48) and a
benzene based chalcone, (E)-3-(2,4-dimethylphenyl)-1-(2-fluoro-4- 2.2. Characterization
methoxyphenyl)prop-2-en-1-one (2F4MA96) were synthesized. Both
experimental measurement and theoretical computations were con­ 2.2.1. Spectroscopy
1
ducted to examine their linear and nonlinear optical properties. The H and 13C NMR were conducted on a JOEL JNM-ECX 400 FT-NMR
results were compared for correlation and to understand the intra­ spectrometer at 400 MHz and 100 MHz, respectively, in deuterated
molecular charge transfer mechanism of thienyl and phenyl based chloroform (CDCl3) with tetramethylsilane (TMS) as internal reference.
chalcones that contributes to these optical properties. To our knowledge, Proton and carbon resonances were reported as ‘s’ for singlet, ‘d’ for
this has not been studied in detailed by other research groups. Linear doublet, and ‘m’ for multiplet. FTIR spectra were measured using a
and nonlinear optical experimental measurements were carried out Perkin Elmer Frontier FTIR Spectrophotometer with attenuated total
using UV-Vis spectroscopy and Z-scan technique, respectively. Whereas reflectance (ATR) sampling technique in the region of 4000–450 cm− 1
the theoretical calculations were performed through density functional (4 scans). UV-Vis spectra were recorded using a Varian Cary 5000 UV-
theory (DFT) modelling. To further explore the extent of the structure- Vis-NIR spectrophotometer in the range of 200–900 nm at 0.0001 M
property relationship, the charge transfer mechanism and molecular concentration in n,n-dimethylformamide (DMF) solution. The mea­
properties were investigated using DFT through a comprehensive anal­ surements were conducted with the samples contained inside 10 mm
ysis of electrostatic potential (ESP) maps, π-electron localization func­ quartz cells.
tion (ELF-π), frontier molecular orbitals (FMOs) and hole-electron
distribution for both chalcone derivatives. 2.2.2. Single-Crystal X-ray Diffraction
The molecular structures of 2F4MA48 and 2F4MA96 were deter­
2. Experimental mined with single crystal X-ray diffraction (SCXRD) method at room
temperature. The selected crystals had dimensions of 0.193 mm × 0.225
2.1. Synthesis and Crystallization mm × 0.428 mm for 2F4MA48 and 0.159 mm × 0.268 mm × 0.644 mm
for 2F4MA96. X-ray measurement was carried out using Bruker APEX II
Compounds 2F4MA48 and 2F4MA96 were synthesized through a DUO CCD area detector diffractometer with graphite-monochromated
crossed-Aldol reaction using the Claisen-Scmidt method. 2’-fluoro-4’- Mo Kα radiation (0.71073 Å). The raw data were reduced using SAINT
methoxyacetophenone 1 (0.24 g, 1.4 mmol) with 4-methylthiophene-2- program [46] and the empirical absorption were corrected with SADABS
carbaldehyde 2 (0.18 g, 1.4 mmol) (for 2F4MA48) or 2,4-dimethylben­ program [46]. The structures were solved by intrinsic phasing method
zaldehyde 3 (0.19 g, 1.4 mmol) (for 2F4MA96) were dissolved in 8 mL and refined by full-matrix least-squares techniques with SHELXTL soft­
methanol in a 50 mL round-bottom flask. The reaction was catalysed by ware package [47,48]. Non-hydrogen atoms were refined with aniso­
an aqueous solution of sodium hydroxide (10%) and the resulting tropic displacement coefficients while hydrogen atoms were positioned
mixture was further stirred for six hours at room temperature inside a geometrically and were assigned isotropic displacement coefficients
fume cupboard. The precipitate obtained was rinsed with cold distilled with riding model, Uiso (H) = 1.2 Ueq (C) or 1.5 Ueq (C). Molecular

2
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Fig. 1. Synthesis pathways of compounds 2F4MA48 and 2F4MA96.

graphics and publications materials were prepared using Mercury [49], focal length of 0.25 m was used, resulting in a peak intensity (at focus),
CrystalExplorer [50] and PLATON [48]. Crystal data, data collection and I0 of 2.36 × 108 Wm− 2 and beam waist radius, ω0 of 23.0 μm. The
structure refinement details are summarized in Table 1. The crystallo­ Rayleigh range, zR value was calculated at 3.18 mm, satisfying the thin
graphic information file (CIF) of 2F4MA48 and 2F4MA96 were depos­ sample condition for Z-scan measurements as the sample thickness was
ited to the Cambridge Crystallographic Data Centre (CCDC) and may be less than zR.
retrieved with the deposition numbers CCDC: 2049735 and CCDC: The measurements were conducted by scanning the quartz cells
2049744, respectively. containing samples, which were placed on a translation stage, as it
moved along the Z-axis (beam propagation direction) at a speed of 1.2
2.2.3. Z-scan and Optical Limiting Studies mm per second. Transmittance was recorded as a function of position
Third-order nonlinear optical responses were investigated through a with respect to the focal length. Nonlinear absorption (NLA) co­
single beam Z-scan approach [51,52], employing open-aperture and efficients, α2 were determined using open-aperture setup (without an
close-aperture techniques using a 532 nm continuous-wave diode-­ aperture), and nonlinear refraction (NLR) coefficients, n2 were evalu­
pumped solid-state (DPSS) operating at 200 mW. The samples (0.01 M) ated based on the ratio of close-aperture (a finite aperture in the far-
were dissolved in DMF and contained within 1 mm thick quartz cells. To field) transmittance to open-aperture transmittance. Optical limiting
focus the laser beam onto the samples, a 7 mm plano-convex lens with responses were extracted using open-aperture Z-scan data and using a
fluence equation as: F(z) = (2Ein )/[πω(z)2 ], where Ein is the input laser
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
Table 1
energy and ω(z) = ω0 1 + (z/zR )2 is the beam radius at a given z-po­
Summary of crystal data and refinement parameters of 2F4MA48 and
2F4MA96. sition along the beam.
Compound 2F4MA48 2F4MA96

CCDC deposition number; CCDC: 2049735; CCDC: 2049744; 2.3. Computational Details
CCDC entry FOHYOU FOHYUA
Molecular formula C15H13FO2S C18H17FO2
Molecular weight 276.31 284.32
2.3.1. Hirshfeld Surface Analysis
Crystal system Monoclinic Triclinic Hirshfeld surface analysis was conducted using CrystalExplorer21
Space group P21/n P1 program [50,53]. The calculations were performed using TONTO [54], a
a (Å) 6.62 4.84 built-in quantum chemistry program, at the calibrated DFT model of
b (Å) 12.09 7.92 B3LYP/6-311G(d,p) level of theory [55,56]. Hirshfeld surfaces mapping
c (Å) 16.28 19.32
α(◦ ) 90.00 98.09
properties over dnorm, shape index, and curvedness were generated, at
β (◦ ) 92.74 91.49 the standard high resolution. The expanded two-dimensional (2D)
γ(◦ ) 90.00 96.31 fingerprint plots of the individual intermolecular interactions were
V (Å3) 1300.60 728.20 plotted as de against di, where de and di refer as the distance from
Z 4 2
Hirshfeld surface to the external nuclei and internal nuclei, respectively.
Dcalc (g cm− 3) 1.411 1.297
Crystal dimensions (mm) 0.193 × 0.225 × 0.428 0.159 × 0.268 ×
0.644 2.3.2. Gaussian 09 Molecular Analysis
Crystal color, shape Brown, block Yellow, plate Theoretical computations were carried out using density functional
µ (mm− 1) 0.255 0.092 theory (DFT) with the Gaussian package; GaussView [57] and
Radiation, λ (Å) Mo Kα, 0.71073 Mo Kα, 0.71073
Tmin, Tmax 0.899, 0.952 0.7408, 0.9509
Gaussian09 [58] programs. The coordinates of the atom arrangement
Measured reflections 19570 17631 from SCXRD CIF file served as the stable starting model for geometry
hkl range − 8≤h≤8 -5 ≤ h ≤ 5 optimization. The optimized geometry was utilized for subsequent DFT
− 15 ≤ k ≤ 15 − 9≤k≤9 calculations. In all calculations, CAM-B3LYP functional [59–61] was
− 21 ≤ l ≤ 20 − 23 ≤ l ≤ 23
combined with 6-311++G(d,p) basis set [62]. DFT simulated 1H and 13C
θ limit (◦ ) 2.099 – 27.707 2.132 – 25.999
Unique reflections 3037 2848 NMR spectra were calculated using gauge-independent atomic orbital
Observed reflections [I > 2σ(I)] 1667 1655 (GIAO) method in chloroform solvent, with TMS as the internal refer­
Parameters 174 193 ence. Vibrational frequencies assignments were determined through
Goodness of fit on F2 1.016 1.018 potential energy distribution (PED) analysis of VEDA4 software [63] and
R [F2 > 2σ(F2)], wR(F2) 0.0544, 0.1317 0.0565, 0.1366
vibrational animation of GaussView [57]. Time-dependent DFT

3
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

(TD-DFT) was employed for computing electronic spectra [64,65] in 3.2. FTIR Analysis
DMF solvent environment.
Nonlinear optical properties were investigated using CAM-B3LYP The molecular structures of 2F4MA48 and 2F4MA96 are further
and an additional B3LYP [58,66] functionals at the same level of the­ validated by using FTIR experimentally and DFT computationally. The
ory (6-311++G(d,p)) via derivative method and were performed in DMF spectra are presented in Figs. S5 and S6. The computed frequencies are
solvent environment for comparison purposes. The parameters deter­ scaled by 0.958 and 0.983 for frequencies above and below 1800 cm− 1,
mined included dipole moment (μ), polarizability (α), first hyper­ respectively [69]. The experimental and scaled DFT-computed FTIR
polarizability (β), and second hyperpolarizability (γ). Solvent-phase DFT bands and their respective band assignments are in good agreement
calculations were conducted via the integral equation formalism variant (Table S2).
of the Polarizable Continuum Model (IEFPCM). The choice of solvents; The C–H stretching modes are observed as medium intensity bands
chloroform and DMF were to match the experimental solvent conditions. in the standard region of 3100–2840 cm− 1. These bands correspond to
Analysis of the DFT-output results were performed using Multiwfn pro­ the C–H stretching of the aromatic rings and alkyl groups for wave­
gram [67], while Visual Molecular Dynamic (VMD) [68] was used for number longer and shorter than 3050 cm− 1, respectively. Another
visualizing and rendering the graphical results. vibrational mode that is assigned to C–H bonds of the aromatic rings is
the out-of-plane bending between 900–690 cm− 1, which informs the
3. Results and Discussion pattern of substitution. The medium intensity band near 873–822 cm− 1
and the strong intensity band near 857–802 cm− 1 in experimental
3.1. NMR Analysis spectra of both compounds is attributed to C–H out-of-plane bending of
the 1,2,4-substituted benzene ring. For 2F4MA48, thienyl ring gives
The molecular structures and the number of hydrogen and carbon characteristics absorption at 738 and 849 cm− 1. Meanwhile, the
atoms in 2F4MA48 and 2F4MA96 molecules are confirmed by 1H NMR stretching of the carbonyl group (C– –O) appeared as an intense and
and 13C NMR analytical method. The 1H NMR and 13C NMR spectra of unique band between 1652–1647 cm− 1 in experimental spectra and
both compounds are shown in Figs. S1–S4, whereas the experimental between 1678–1675 cm− 1 in theoretical spectra. The intense stretching
and theoretical chemical shifts are listed in Table S1. The experimental is due to the large dipole moment between the large partial positive
chemical shifts are slightly deviated from the theoretical values, but the charge of the carbonyl carbon and the large partial negative charge of
results are highly consistent, and the number of protons and carbons the oxygen atom.
present are in good agreement. The most intense absorptions are observed between 1611–1578
The proton NMR chemical shifts are towards downfield between cm− 1 of the experimental spectra in both compounds, while the rela­
8.1–6.6 ppm for aromatic and trans hydrogens and towards upfield tively weaker vibrational bands are detected between 1541–1418 cm− 1.
(shielded) between 2.5–2.2 ppm for methylbenzene. The presence of These bands signify the stretching of C– –C bonds on the enone bridge
methoxy protons is confirmed by the occurrence of singlet centring at δ and the aromatic ring. The trans C– –C bond of the enone bridge is
≈ 3.85 ppm with an integration value of 3. In 2F4MA96, the presence of indicated by the medium strength experimental band at 977 cm− 1 and
the two vinylic hydrogens (Hα and Hβ) are signified by the two doublets 975 cm− 1 for 2F4MA48 and 2F4MA96, respectively. The stretching of
in the range of 7.37–7.33 ppm and 8.08–8.04 ppm for H8 and H9, the aryl fluorine and the aryloxy group are demonstrated as strong bands
respectively. A similar pattern should be observed in 2F4MA48. How­ between 1098–1094 cm− 1 and 1234–1041 cm− 1, respectively. Addi­
ever, the doublet signal of H9 overlapped with the triplet signal of H5 tional vibrational mode is observed for 2F4MA48 at 839 and 650 cm− 1,
causing it to resolve into a multiplet between 7.88–7.80 ppm. For both and is assigned to the C–S band stretching in the thienyl ring.
compounds, the presence of fluorine atom with nuclear spin of I = ½ led
to the coupling with both adjacent hydrogen and fluorine atoms, 3.3. Single Crystal X-ray Diffraction and Hirshfeld Surface Analysis
resulting in doublet signal for H2 in 6.65–6.62 ppm and a triplet signal
for H5 between 7.90–7.83 ppm. The presence of methyl protons on the 2F4MA48 is crystallised in the centrosymmetric monoclinic crystal
thienyl ring in 2F4MA48 is determined as a singlet with integration system of space group P21/n and 2F4MA96 is crystallised in the
values of 3 at 2.25 ppm whereas the presence of alkyl protons on the
centrosymmetric triclinic crystal system of space group P1. The phenyl
dimethyl phenyl ring in 2F4MA96 is denoted by the singlets at 2.33 ppm
ring (C1–C6) of 2F4MA48 and 2F4MA96 have a fluorine atom attached
and 2.44 ppm.
to the ortho-position and a methoxy group attached to the meta-position
In 13C NMR spectra, the most downfield (deshielded) peaks centring
(Fig. 2). With respect to the π-conjugated carbon-carbon double bond
at δ ≈ 188–186 ppm confirmed the presence of the carbonyl carbon (C7).
C8––C9, both molecules exist in trans configuration as suggested by the
A significant deshielding effect is observed at 165–161 ppm for C1 and
C3 that are directly bound to the fluorine atom and methoxy group torsion angle (C7–C8– –C9–C10) of nearly ±180◦ at 179.8◦ and -179.3◦
respectively, due to the nature of fluorine atom and methoxy groups that for 2F4MA48 and 2F4MA96, respectively. The phenyl ring (C1–C6) for
are more electronegative than carbon atoms, decreasing the surrounding both compounds is almost co-planar with the corresponding aromatic
electron density. The same effect is observed for aromatic carbon that is ring that is thienyl ring (S1/C10–C13), with a dihedral angle of 3.63◦ for
attached to the methyl group which results in signals around 141–138 2F4MA48, and phenyl ring (C10–C15), with a dihedral angle of 3.92◦
ppm, which is assigned to C12 for 2F4MA48 and C11 and C13 for for 2F4MA96.
2F4MA96. The vinylic carbons (C8 and C9) are indicated by the char­ In the crystal structure of 2F4MA48 (Fig. 3a), adjacent molecules
acteristic peaks of α-carbon and β-carbon between 126–124 ppm and form centrosymmetric dimers via a pair of intermolecular C2—H2A⋅⋅⋅F1
142–137 ppm, respectively. The aromatic carbons that are bridged with hydrogen bonds (Table 2) with R22 (8) graph-sets motif. These dimers are
enone moiety (C6 and C10) exhibits a medium and a weak intensity joined through C13—H13A⋅⋅⋅O1 hydrogen bond to form two-
signals around 120 ppm and 140–131 ppm, respectively, whereas the dimensional corrugated sheets (Fig. 3b) and are further linked into
remaining aromatic carbons are observed between 133–101 ppm. The edge-to-edge fashions through the bifurcated C9—H9A⋅⋅⋅O2 and
high field region is mainly assigned to the methyl carbons (22–15 ppm) C11—H11A⋅⋅⋅O2 hydrogen bonds with R12 (6) graph-set motif, forming a
and methoxy carbons (~56 ppm) due to their weak resonance structure. three-dimensional (3D) packing. Viewing the 3D molecular packing
Carbon atoms in the proximity of fluorine atom shows splitting reso­ along c-axis shows the zig-zag packing arrangement (Fig. 3c) that are
nance due to the strong and long-range fluorine-carbon couplings. further consolidated by C—H⋅⋅⋅π interaction between H15C and Cg2 at
the separation distance of 2.76 Å (Table 3). The parallel-displaced π⋅⋅⋅π
interactions between the molecules are very weak, with separation

4
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Fig. 2. Asymmetric units of (a) 2F4MA48 and (b) 2F4MA96 shown in 50% probability displacement ellipsoids with labels for non-hydrogen atoms.

distances between 4.28 to 5.75 Å. 2F4MA48 compared to 2F4MA96. The characteristic ‘wings’ feature in
The structural arrangement of 2F4MA96 involves centrosymmetric the plot of H⋅⋅⋅C/C⋅⋅⋅H is evidenced by the C–H⋅⋅⋅π interactions, sup­
dimers (Fig. 4a) formed by a pair of C16—H16A⋅⋅⋅O2 hydrogen bonds ported by the shape index and curvedness plots (Fig. S8). On the shape
(Table 2) with R22 (6) graph-set motif. These dimers are linked into two- index surfaces, the complementary pair of the red concave and blue
column infinite chains through C2—H2A⋅⋅⋅O1 hydrogen bond into side- convex region (circled in red) over the aromatic ring and hydrogen
by-side fashions, forming C22 (6) graph-set motif that are propagating in atoms imply the interplanar stacking arrangement by C–H⋅⋅⋅π interac­
b-axis direction (Fig. 4b). The two-column chains lie as two-dimensional tion. Whereas, on the curvedness surface, it is demonstrated by the light
sheets parallel to (1 0 4) and are stacked together (Fig. 4c) by the offset blue region (circled in red) over the aromatic ring and hydrogen atoms.
C18—H18C⋅⋅⋅Cg4 interactions (2.78 Å) (Table 3). Similarly to The Hirshfeld surface of 2F4MA48 is additionally contributed by S⋅⋅⋅H/
2F4MA48, the offset-displaced π⋅⋅⋅π interactions between the centroids H⋅⋅⋅S interaction from the thienyl ring at 8.6%.
of 2F4MA96 are very weak, with separation distances ranging from 4.84
to 5.84 Å. Additionally, both structures possess an intramolecular
3.4. Molecular Electronic Structure Analysis
C8—H8A⋅⋅⋅F1 hydrogen bond interaction with a S(6) graph-set motif
and comparable separation distance.
3.4.1. Molecular Electrostatic Potential
The Hirshfeld surfaces are calculated to visualize the intermolecular
The intermolecular interactions of 2F4MA48 and 2F4MA96 were
contacts within the crystal structures of 2F4MA48 and 2F4MA96. The
studied by examining their molecular electrostatic potential (ESP) maps
Hirshfeld surface mapped over the normalised contact distance dnorm are
of the vdW surfaces [70]. The ESP maps provide a visual representation
described in blue, white, and red colour scheme which represent distant
of the charge distribution within the molecule with red, white, and blue
contact, contact at approximate length with van der Waals (vdW) radii,
regions indicating negative, neutral, and positive electrostatic potential,
and short contact, respectively. In Fig. 5, the respective acceptor and
respectively. The generated ESP maps, along with the values of the
donor sites showing the strong intermolecular C–H⋅⋅⋅O and C–H⋅⋅⋅F
selected electrostatic potential for the surface minima and maxima are
hydrogen bonds, labelled as ① and ②, and ③, respectively, are revealed
presented in Fig. 6. From the ESP maps, the electrophilic attack sites are
on the Hirshfeld surface as red spots. Meanwhile, the bifurcated
around the vicinity of electron donating groups; oxygen, and fluorine
hydrogen bonds in 2F4MA48, the C9–H9A⋅⋅⋅O2 and C11–H11A⋅⋅⋅O2
atoms, displaying values ranging from -54.93 to -21.27 kcal mol− 1.
interactions appear as medium red spots (Fig. 5a: ②). The C–H⋅⋅⋅O
These regions of negative electrostatic potential arise from the presence
hydrogen bonds (Fig. 5b: ①,②) in 2F4MA96 are indicated by the bright
of lone pairs, with the oxygen atoms possessing two lone pairs and the
red spots of similar size due to the comparable separation distance of ~
fluorine atom exhibiting three lone pairs. Consequently, these areas
2.6 Å (Table 2). An additional red spot of C⋅⋅⋅H intercontact is revealed
exhibit a higher electron density, rendering them more susceptible to
for 2F4MA96 between C16 and H5A (Fig. 5b: ⑤). Meanwhile, the pale
electrophilic interactions. The most prominent red region surrounding
red spots labelled ④ represent the C–H⋅⋅⋅π short contacts that are
the carbonyl oxygen atom indicates the global minima with values
involved in the molecular stacking in both compounds.
-53.69 kcal mol− 1 for 2F4MA48 and -54.93 kcal mol− 1 for 2F4MA96,
The subsequent 2D fingerprint plots are developed to summarize the
suggesting a greater electron-withdrawing ability of the carbonyl group
contribution of individual intermolecular interaction on the Hirshfeld
relative to the methoxy group and fluorine atom. Meanwhile, the low
surface through de against di plot (Fig. S7). The fingerprint plots utilized
electron density hydrogen atoms are mainly associated with nucleo­
a blue-green-red colour scheme to visualize the relative scale of small,
philic attack sites (blue regions) with values ranging from 17.57 to 33.29
intermediate, and greatest contributions, respectively. The character­
kcal mol− 1. The positive electrostatic potential of both compounds
istic pair of sharp spikes representing the strong hydrogen bond is
shows a comparable charge distribution with the higher ESP values
observed for H⋅⋅⋅O/O⋅⋅⋅H interactions for both compounds. The H⋅⋅⋅O/
predominately surrounding the methoxy hydrogens. The analysis of
O⋅⋅⋅H pair of sharp spikes in 2F4MA96 is slightly shorter (de +
molecular electrostatic potential indicates that both compounds are in a
di ≃ 2.5Å) than those of 2F4MA48 (de + di ≃ 2.3Å), consistent with its
donor-acceptor-donor (D-A-D) push-pull system. The incorporation of
longer separation distance. In 2F4MA48, a second pair of sharp spikes at
an electron-rich thienyl ring has induced a significantly stronger
de + di ≃ 2.4Å, superimposed with the H⋅⋅⋅F/F⋅⋅⋅H interactions is
push-pull effect (33.29 kcal mol− 1 compared to 18.70 kcal mol− 1 at
spotted, signifying the C–H⋅⋅⋅F hydrogen bond. In 2F4MA96, the H⋅⋅⋅F/
relative sites) which could be the reason for the improved NLO prop­
F⋅⋅⋅H interactions appear as a pair of blunt spikes at de + di ≃ 3.0Å,
erties in 2F4MA48. These points of surface extrema are evidence of the
suggesting that the interaction does not portray hydrogen bonds nor
C–H⋅⋅⋅O and C–H⋅⋅⋅F chemical interactions within the compounds,
short intercontact. A short intermolecular H⋅⋅⋅H intercontact (de +
consistent with the Hirshfeld surface analysis in Section 3.3.
di ≃ 2.1Å) is observed in 2F4MA96, covering the most surface area at
Additionally, the contribution of each fragment: methylthienyl/
51.1%. Meanwhile, the H⋅⋅⋅H intercontact in 2F4MA48 shows a longer
dimethylphenyl, enone bridge, and fluoro-methoxyphenyl on the mo­
separation distance (de + di ≃ 2.4Å) and a smaller contribution of
lecular ESP was calculated via the quantitative molecular surface anal­
38.6%. The highly different contribution is attributed to the introduc­
ysis module [71]. The results are graphically presented in Fig. S9. The
tion of thienyl ring and relatively less amount of methyl group in
polar surface areas (|ESP|>10 kcal mol− 1) determined from the plots are

5
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Fig. 3. Structural arrangement of 2F4MA48 showing (a) R22 (8) centrosymmetric dimer, (b) two-dimensional corrugated sheet viewed along a-axis, and (c) three-
dimensional packing through bifurcated hydrogen bonds showing zig-zag structural arrangements along with C—H⋅⋅⋅π interactions, viewed along c-axis. Green
and blue dashed lines are intra- and intermolecular C—H⋅⋅⋅F hydrogen bonds. Magenta dashed lines are C—H⋅⋅⋅O hydrogen bonds and orange dashed lines are
C—H⋅⋅⋅π interactions.

191.27Å (61.26%) for 2F4MA48 and 199.80Å (59.11%) for 2F4MA96. relative to the phenyl ring [72,73].
Noting that the benzoyl system in both compounds is similar, the higher
percentage of polar surface area in 2F4MA48 can be attributed to its 3.4.2. Electron Delocalization
cinnamoyl system which is composed of sulphur atom. This difference The degree of π-electron delocalization in the molecules is assessed
may arise from the greater polarity of the electron-rich thienyl ring using real space function and topology analyses based on electron

6
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Table 2
Hydrogen-bonds geometry of 2F4MA48 and 2F4MA96 (Å, ◦ ).
Compound D—H⋅⋅⋅A D—H H⋅⋅⋅A D⋅⋅⋅A D—H⋅⋅⋅A Symmetry code

2F4MA48 C2—H2A⋅⋅⋅F1 0.93 2.58 3.479 163 -x, 1-y, 1-z


C13—H13A⋅⋅⋅O1 0.93 2.45 3.114 129 1/2+x, 3/2-y, 1/2+z
C9—H9A⋅⋅⋅O2 0.93 2.68 3.473 143 1/2-x, 1/2+y, 1/2-z
C11—H11A⋅⋅⋅O2 0.93 2.64 3.414 141 1/2-x, -1/2+y, 1/2-z
C8—H8A⋅⋅⋅F1 0.93 2.15 2.798 125
2F4MA96 C2—H2A⋅⋅⋅O1 0.93 2.62 3.525 165 x, -1+y, z
C16—H16A⋅⋅⋅O2 0.96 2.60 3.382 139 3-x, -y, -z
C8—H8A⋅⋅⋅F1 0.93 2.16 2.810 127

Table 3
C—H⋅⋅⋅π interactions geometry (Å, ◦ ).
Compound C—H⋅⋅⋅π C—H H—π C⋅⋅⋅π C—H⋅⋅⋅π Symmetry code

2F4MA48 C15—H15C⋅⋅⋅Cg2 0.96 2.76 3.664 157 2-x, 2-y, 1-z


2F4MA96 C18—H18C⋅⋅⋅Cg4 0.96 2.78 3.681 156 -1+x, y, z

*Cg2 and Cg4 are the centroid for S1/C10–C13 and C10–C15 rings, respectively.

localization function (ELF) [74]. The ELF-π isosurfaces, along with the 2F4MA48 is more chemically reactive than 2F4MA96, potentially
bifurcation points defined in the topology analysis framework, are leading to a higher optical hyperpolarizability.
depicted in Fig. 7. A greater degree of electron delocalization within
ELF-π domains is indicated by the large value of bifurcation points and is
visually presented as continuous isosurfaces. The lower aromaticity of 3.6. Linear Absorption Studies
the thienyl ring is demonstrated by the split ELF-π domains over the
sulphur atom, in contrast to the quasi-doughnut shape observed for 3.6.1. UV-Vis Spectroscopy
phenyl rings. This suggests a greater electron localization on the phenyl The linear absorption spectra are experimentally measured using
ring of 2F4MA96 than on the thienyl ring of 2F4MA48. Consequently, ultraviolet-visible (UV-Vis) spectroscopy in DMF solution and are
this inhibits the outward flow of electrons from the phenyl ring theoretically computed by using TDDFT [65]. Both the experimental and
compared to the thienyl ring. Accordingly, the splits between ELF-π theoretical absorption spectra are plotted in Fig. 9, and they show a close
domains over the enone bridge for both compounds imply that electrons agreement. The absorption spectra of both compounds exhibit two ab­
are not easily delocalized from one end to the other. Yet, the high ability sorption maxima, as expected for chalcone [82,83], with the prominent
of charge transfer flow through the cinnamoyl system compared to the band (band I) located between 329 nm and 351 nm, and the second
benzoyl system is characterized by the relatively high bifurcation values slightly weaker band (band II) located between 264 nm and 308 nm.
of 0.42–0.56 at bonds C7–C8 and C9–C10 compared to 0.33 at band Both bands are attributed to π→π* electronic transitions [84]
C6–C7. Additionally, the larger bifurcation values over the enone throughout different moieties of chalcone; where band I is due to the
bridge in 2F4MA48 suggests that the delocalization of electrons from absorption of the cinnamoyl group, while band II is due to the absorp­
both terminal ring to the carbonyl group is relatively easier than in tion of the benzoyl group. This is supported by the hole-electron dis­
2F4MA96. Thus, the higher degree of π-electron delocalization in tribution analysis (Section 3.6.3). 2F4MA48 shows a tiny bathochromic
2F4MA48 may result in a greater optical nonlinearity. shift in band I absorbance compared to 2F4MA96, implying that both
compounds require similar energy for electronic transition. Meanwhile,
in terms of the absorbance intensity of band I, 2F4MA96 is slightly
3.5. Frontier Molecular Orbitals (FMOs) greater than 2F4MA48 due to the higher aromaticity of the phenyl ring
relative to the thienyl ring, which increases the intensity of π→π*
The chemical stability of 2F4MA48 and 2F4MA96 are evaluated transitions.
based on the frontier molecular orbitals (FMOs) from TD-DFT calcula­ Tauc’s plot is used to estimate the optical energy gap, Eg , while the
tion [75]. The FMOs are decomposed into three fragments, and their extrapolation of the experimental spectra is carried out based on relation
compositions are calculated (Table S3) using Hirshfeld weighting (1):
function of Multiwfn program [76]. The three-dimensional (3D) plots of ( )12
the highest occupied molecular orbital (HOMO) and the lowest unoc­ α0 hν = C hν − Eg (1)
cupied molecular orbital (LUMO) are presented in Fig. 8, along with the
energy level map of the three highest occupied and three lowest unoc­ where C is a constant, h is Planck’s constant, ν is the frequency of the
cupied molecular orbitals. From the fragment composition plots; the incident photons and α0 is the optical absorption coefficient; α0 =
delocalization of HOMO primarily occurs on the methyl­ 2.303 × (A /d), A is the absorbance and d is the thickness of the sample.
1
thienyl/dimethylphenyl ring at ~68%. Meanwhile, LUMO mainly lies The plot of (α0 hν)2 against photon energy (hν) for both compounds is
on the enone bridge at ~57% while also equally contributed by the shown in Fig. 9 (insert graphs). The Eg are determined to be 3.21 eV and
terminal rings at ~21% each. The calculated HOMO-LUMO gap is 5.88 3.42 eV for 2F4MA48 and 2F4MA96, respectively, and this trend is
eV for 2F4MA48 and 5.92 eV for 2F4MA96. The slightly smaller value consistent with the HOMO-LUMO energy gap. The cut-off wavelength is
for 2F4MA48 can be attributed to the presence of a heteroatom in the determined by extrapolating the straight-line portion of the higher en­
terminal ring of the molecule. The more electronegative sulphur atom, ergies experimental band I absorbance peak. The cut-off points are 401
compared to carbon, leads to an asymmetrical distribution of electron nm for 2F4MA48 and 377 nm for 2F4MA96.
density due to sulphur’s electron-withdrawing behaviour [77,78].
Moreover, the lone pair from sulphur contributes to additional electron 3.6.2. Interfragmentary Charge Transfer Analysis
delocalization [79,80] by interacting with the π-electron in the mole­ The contribution of each fragment towards the linear absorption
cule, lowering the energy levels for LUMO [81]. This suggests that spectra is analysed via charge-transfer spectrum of interfragmentary

7
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Fig. 4. Structural arrangement of 2F4MA96 showing (a) R22 (6) centrosymmetric dimer, (b) a pair of one-dimensional two-column chains, propagating along b-axis
direction, and (c) three separated layers joined by C—H⋅⋅⋅π interaction. Magenta and green dashed lines are intermolecular C—H⋅⋅⋅O and intramolecular C—H⋅⋅⋅F
hydrogen bonds, respectively. Orange dashed lines are C—H⋅⋅⋅π interactions.

charge transfer (IFCT) analysis [85]. The electronic transition is shown the benzoyl system. The S0 → S3 excitation is mainly contributed by
as discrete spikes below the spectra relative to the oscillator strength HOMO-1 → LUMO accounting for 92% contribution, while S0 → S4
(Fig. 10). The band I absorption peak in both compounds is attributed to excitation is mainly due to HOMO-2 → LUMO at 83% contribution. The
the same S0 → S1 excitation, primarily from HOMO → LUMO electronic main electronic transitions contributing to the linear absorption spectra
transition. The IFCT plot reveals that this transition is predominantly are listed in Table 4 along with their corresponding excitation energies,
characterized by charge transfer and charge redistribution associated excitation wavelength, oscillator strength and orbital contributions.
with the respective cinnamoyl systems; transfer from methylthienyl ring In addition, the charge-transfer length (Δr) during electronic exci­
(Me-Th) in 2F4MA48 and dimethylphenyl ring (Me2-Ph) in 2F4MA96 to tations is calculated to determine the suitability of the functionals cho­
the central carbonyl group, and followed by charge redistribution on sen for DFT calculations. The assessment is based on whether the
enone bridge and corresponding aromatic ring (Me-Th or Me2-Ph). electronic excitation occurs in short-range (Δr ≤ 1.5 Å) or long-range
Conversely, band II originates from the doubly degenerate S0 → S3 (Δr ≥ 2.0 Å) [86]. From Table 4, the Δr values for all primary elec­
and S0 → S4 excitations in 2F4MA48, and solely S0 → S4 excitation in tronic transitions are found to be greater than 2.0 Å, implying that the
2F4MA96. For 2F4MA48, the doubly degenerate excitations involve charge-transfer excitations are long-range. As a result, it is necessary to
contributions from the entire molecule, representing a mix of charge- choose the long-range corrected functionals with a high composition of
transfer spectrum between the cinnamoyl and benzoyl systems, with a Hartree-Fock exchange component (≥ 33%) [86]. This shows the rele­
slightly greater contribution from the former. In contrast, for 2F4MA96, vance of employing CAM-B3LYP functional, which has 65% of
it is due to the charge redistribution and charge transfer pertaining to Hartree-Fock exchange composition, in this study.

8
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Fig. 5. Intermolecular interaction of (a) 2F4MA48 and (b) 2F4MA96 visualized by Hirshfeld surfaces plotted over dnorm.

Fig. 6. Molecular electrostatic potential (ESP) map of the vdW surface of (a) 2F4MA48 and (b) 2F4MA96. Points of surface minima and maxima are shown as dots
and labelled with the corresponding ESP values (kcal/mol) in red and blue, respectively.

Fig. 7. ELF-π isosurfaces of (a) 2F4MA48 and (b) 2F4MA96 with the bifurcation points indicated by the orange spheres and labelled for selected points (isovalue
of 0.7).

3.6.3. Hole-electron Distribution Analysis measure of the spatial extension of hole and electron distribution in x, y,
The path of the intramolecular charge transfer (ICT) of the electronic z, and CT directions, the t index describes the separation degree of hole
excitations is investigated using hole-electron analysis [84]. It is visu­ and electron in CT direction, ECoul signifies the hole-electron Coulomb
alized as ‘electron’ (yellow isosurface) and ‘hole’ (grey isosurface) to attractive energy, whereas HDI and EDI refer to hole and electron
signify the increment and decrement of electron density, respectively delocalization indices, respectively. For S0 → S1 excitation, the D index
(Fig. 11). The distributions are quantitatively characterized [87] to of both compounds is slightly larger than the average C–C bond length
distinguish between the local excitation (LE) and the charge-transfer (≃1.5 Å), supported further by an Sr index of ~ 0.7, indicating that more
(CT) electron excitations (Table 5). The D index refers to the total than 70% of hole and electron are overlapping. Moreover, t index below
magnitude of the CT length, the Sr index describes the overlapping 0 indicates that the separation between electron and hole distributions
extent of hole and electron (upper limit is 1.0), the H index is the overall are indistinct. By taking all these analyses into account, S0 → S1

9
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Fig. 8. Frontier molecular orbitals and fragment composition plots of (a) 2F4MA48 and (b) 2F4MA96 (isovalue of 0.05).

Fig. 9. UV-Vis absorption spectrum of (a) 2F4MA48 and (b) 2F4MA96 and Tauc’s plot (insert graphs) showing the optical energy gaps. The solid blue lines represent
the experimental UV-Vis spectra while the dashed red lines represent DFT-simulated spectra.

Fig. 10. Electronic absorption spectra (black) of (a) 2F4MA48 and (b) 2F4MA96. The charge-transfer spectrum is shown in coloured lines. The peak positions of the
absorption bands (λmax ) are labelled for each band. The oscillator strength of the selected electronic transitions (major contribution for absorption bands) is shown as
distinct spikes and labelled with Roman numerals.

10
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Table 4 gap of both compounds (see Section 3.6.1) is larger than the energy of a
The excitation energy, Eex , absorption wavelength, λex , oscillator strength, f0 , single photon (2.33 eV) but lower than the combined energy of two
and charge transfer length, Δr of electronic transitions from ground state, S0 to photons (4.66 eV) at wavelength of 532 nm, suggesting the involvement
excited state, Sn with the percentage contribution of FMOs electronic transition of 2PA process. To characterize the efficiency of the 2PA transition, the
in 2F4MA48 and 2F4MA96. value of 2PA cross-section, σ 2 is calculated using Eq. A.4. The greater σ2
S0 →Sn Eex (eV) λex (nm) f0 Δr (Å) Major contribution ( > 10%) value observed for 2F4MA48 is consistent with its stronger NLA
2F4MA48 response.
S0 →S1 3.61 343.49 0.75 2.91 H→L (93.6%) In the 2PA process, two photons can be absorbed either simulta­
S0 →S3 4.37 284.00 0.22 3.53 H − 1→L (92.2%) neously (referred to as genuine 2PA) or sequentially, with one photon
S0 →S4 4.70 263.65 0.16 3.46 H − 2→L (83.3%)
excites the electron to a near-resonant state (known as a virtual inter­
2F4MA96
S0 →S1 3.64 340.63 0.93 2.83 H→L (94%) mediate state) and another photon facilitating the electron transition to
S0 →S4 4.67 265.63 0.20 3.48 H − 2→L (83.1%) a higher energy level. In the latter 2PA case, an ESA process is involved,
leading to the depletion of the ground state [90–92]. To determine if the
ESA process is involved, the excited-state and ground-state absorption
excitation is determined to be LE type. Similarly, S0 → S4 excitation is cross-sections (σ ex and σg ) are calculated using Eqs. A.5 and A.6,
also regarded as LE type, evidenced by the overlapping between the respectively. σg of 2F4MA48 and 2F4MA96 are 4.19 × 10− 21 and 1.33 ×
distribution of hole and electron. In contrast, S0 → S3 excitation of 10− 21 cm2 whereas σex of 2F4MA48 and 2F4MA96 are 1.25 × 10− 21 and
2F4MA48, involving the thienyl ring which is absent in 2F4MA96, is 0.61 × 10− 21 cm2, respectively. For both compounds, σ ex values are
attributed to CT excitation based on the high degree of separation found to be smaller than σ g values, indicating that ESA does not
indicated by a positive t index. significantly contribute to the NLA property. Therefore, the NLA prop­
erty is determined to be a result of genuine 2PA.
3.7. Nonlinear optical studies In closed-aperture (CA) Z-scan, the peak-to-valley configuration
signifies the negative NLR property, a self-defocusing effect. However, in
3.7.1. Nonlinear Absorption (NLA) and Nonlinear Refraction (NLR) laser-material interactions, continuous wave laser can lead to thermal-
Z-scan technique is employed to investigate the third-order nonlinear effect-built-up due to the slow response time which results in negative
optical response under 532 nm continuous-wave laser domain. The nonlinearity. H. Y. Tan et al. [93] demonstrated the thermal lens effect
open-aperture and closed-aperture Z-scan data are fitted to the Sheik caused by temperature and refractive index gradient arising from local
Bahae model (see Appendix) [51,52] following relations A.1 and A.2, heat accumulation. Meanwhile, Maurya et al. [94] presented the influ­
respectively. The fitted data are presented in Fig. 12. Subsequently, ence of thermal effects on the nonlinear refractive index. Their findings
based on Sheik Bahae’s theory, the nonlinear absorption (NLA) coeffi­ suggest that thermal effect primarily affects NLR property and not NLA
cient, α2 , and the nonlinear refraction (NLR) coefficient, n2 are extrac­ property. S. M. Mian et al. [95] demonstrated the impact of localized
ted. This Z-scan experiment is characterized as an off-resonant process, heating on NLR property and proposed that a positive deviation of the
as both compounds exhibit transparency across visible and infrared peak-valley separation from Kerr value of 1.7zR was due to thermal
ranges, with cut-off points at 401 nm and 377 nm for 2F4MA48 and contributions. In our study, we found that the separation between the
2F4MA96, respectively (see Section 3.6.1). The NLO parameters of both peak and valley of the CA graph is indeed greater than 1.7zR, indicating
compounds, including NLA and NLR coefficients, real and imaginary
parts as well as the absolute third-order NLO susceptibility, molecular
second hyperpolarizability, two-photon absorption cross-section, Table 5
ground and excited state absorption cross-section, and thermal-optical Quantitative indices defined in the hole-electron analysis framework of com­
pounds 2F4MA48 and 2F4MA96.
coefficient are listed in Table 6. In open-aperture (OA) Z-scan, the
presence of the valley at z = 0 indicates a positive NLA response, a S0 →Sn D(Å) Sr (a.u.) H(Å) t(Å) ECoul (eV) HDI EDI
reverse saturable absorption (RSA) process. The deeper valley observed 2F4MA48
for 2F4MA48 compared to 2F4MA96 suggests a stronger NLA response, S0 →S1 1.90 0.70 2.99 -0.51 4.93 7.92 7.83
as indicated by the larger α2 value. In NLO materials, NLA response can S0 →S3 2.53 0.56 2.73 0.44 4.72 9.21 7.77
2.27 0.68 2.98 -0.28 4.88 8.86 7.81
result from various photon absorption mechanisms, such as two-photon
S0 →S4
2F4MA96
absorption (2PA), excited state absorption (ESA), multiple-photon ab­ S0 →S1 1.83 0.71 3.05 -0.66 4.99 7.64 7.93
sorption (MPA), free carrier absorption, nonlinear scattering, or a S0 →S4 2.33 0.66 2.93 -0.16 4.88 9.02 7.89
combination of these processes [88,89]. However, the optical energy

Fig. 11. Hole-electron distribution of the main excited transitions contributing to bands I and II linear absorption peaks of 2F4MA48 and 2F4MA96. Yellow and grey
isosurfaces represent the increment (electron) and decrement (hole) of electron density, respectively (isovalue of 0.002).

11
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Fig. 12. (a) Open-aperture and (b) closed-aperture Z-scan curves of 2F4MA48 (blue) and 2F4MA96 (red), under 532 nm continuous wave laser domains. The dotted
lines depict the theoretical fits of Sheik Bahae’s formalism while the scattered plots represent the experimental data.

the presence of thermal lensing effect in the NLR response. This thermal of the compound. This finding aligns with the discussion in the ELF-π
lens effect arises from variations in the medium’s refractive index with analysis (Section 3.4.2) which has demonstrated that the degree of
( )
temperature n = n0 + dT dn
T . To compare the significance of thermal electron delocalization between terminal aromatic rings in 2F4MA48 is
effects in both compounds, the thermal-optical coefficient, dT dn
is calcu­ greater than that in 2F4MA96.
lated using relation A.7. The values are -7.63 × 10− 3 K− 1 for 2F4MA48
3.7.2. Optical limiting and Figures-of-Merits
and -2.97 × 10− 3 K− 1 for 2F4MA96. The larger magnitude of dT dn
for
Since both compounds exhibit RSA responses, wherein the amount of
2F4MA48 may explain its stronger self-defocusing effect relative to
transmitted light decreases as the intensity of incident laser light in­
2F4MA96 and it is consistent with its more prominent CA scan trace
creases, these compounds could potentially be used as optical limiters to
depth and greater n2 .
protect sensitive optical components or the naked eye from high-
From α2 and n2 values extracted based on Sheik Bahae theory, the
intensity irradiance [97,98]. Therefore, to evaluate their potential, the
real, ℜ(χ (3) ) and imaginary, ℑ(χ (3) ) parts of third-order susceptibility, optical limiting response of 2F4MA48 and 2F4MA96 is measured using
χ (3) ( − ω; ω, ω, − ω) are calculated by using Eq.s A.8 and A.9, respec­ the open-aperture Z-scan technique with input intensities of a 532 nm
tively. Afterwards, the absolute value of χ (3) are determined by applying CW incident laser in DMF solution. The plot of normalized transmittance
Eq. (2): against input fluence is presented in Fig. 13. From the plot, it is clearly
⃒ (3) ⃒ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ portrayed that 2F4MA48 demonstrates superior optical limiting
⃒χ ( − ω; ω, ω, − ω)⃒ = [ℜ(χ (3) )]2 + [ℑ(χ (3) )]2 (2) behaviour compared to 2F4MA96. The steep slope observed in the
2F4MA48 trace blocks 12% of transmittance at the maximum input
The χ (3) values are found to be 8.04 × 10− 8 esu and 1.08 × 10− 8 esu
fluence recorded (23.9 kJ cm− 2). In contrast, 2F4MA96 limits only up to
for 2F4MA48 and 2F4MA96, respectively. Following χ (3) , the molecular 2% of transmittance at the same input fluence. The specific input fluence
second hyperpolarizability, γ( − ω; ω, ω, − ω) is calculated using Eq. at which transmittance begins to attenuate due to nonlinear process is
(3): referred to as the onset limiting threshold (FS) [99]. The values obtained
χ (3) ( − ω; ω, ω, − ω) are 2.63 kJ cm− 2 for 2F4MA48 and 5.23 kJ cm− 2 for 2F4MA96. The
γ( − ω; ω, ω, − ω) = (3) smaller FS of 2F4MA48 is consistent with its higher 2PA cross-section
Nc L 4
analysis [90,92], suggesting that it has the ability to provide a faster
where Nc = NA C × 10− 3 is the molecular number density in unit volume, response in limiting light by attenuating it at lower intensities. Consis­
NA is Avogadro’s number, C is concentration of the sample solution in tently, the optical limiting threshold (FOL) which is the input fluence at
mol L− 1, and L is Lorent local field. The γ values are found to be 3.98 × which the transmittance is 50% of linear transmittance, is determined to
10− 27 and 1.08 × 10− 27 esu for 2F4MA48 and 2F4MA96, respectively. be 5.32 kJ cm− 2 and 5.55 kJ cm− 2 for 2F4MA48 and 2F4MA96,
Notably, the larger χ (3) and γ of 2F4MA48 indicate the greater optical respectively. The lower FOL value of 2F4MA48 suggests its capability to
nonlinearity compared to 2F4MA96, accordant to the presumption that limit transmitted light at a lower input intensity. These findings align
a smaller optical energy gap leads to better NLO responses. The with the greater optical nonlinearity of 2F4MA48 compared to
enhanced optical nonlinearity in 2F4MA48 can be attributed to the 2F4MA96.
presence of a sulphur atom in one of the terminal aromatic rings, which The optical limiting thresholds of the studied compounds were
improves its electron-withdrawing capabilities and, consequently, in­ compared with values reported for both thienyl and phenyl based
creases the polarity of the compound [96]. Additionally, the stronger chalcone derivatives under the same experimental conditions of a 532
push-pull effect and lower aromaticity of thienyl ring in 2F4MA48 nm continuous wave laser domain (Table 7). It is observed that the
compared to phenyl ring in 2F4MA96, are another contributing factors. majority of reported FOL values for thienyl based chalcones are notice­
As a result, 2F4MA48 exhibits greater intramolecular charge transfer to ably smaller than those of phenyl based chalcones, implying their better
central carbonyl group from both terminal rings: benzoyl ring with the performance. The lowest value is 1.67 kJ cm− 2, as presented by V. S.
strong donor − OCH3 group and thienyl ring with weak donor CH3 Naik et. al. for 1-(5-bromo-2-thienyl)-3-(4-bromophenyl)prop-2-en-1-
group, inducing a larger polarization and promoting the NLO response one (B5B2SC). In their study, V. S. Naik et al. have also recorded FOL

12
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

K− 1 )
5
(10−

-7.66
-2.98
dT
dn
cm2)
21
σex(10−

1.25
0.61
cm[2])
21
σg(10−

4.19
1.33
GM)

Fig. 13. Optical limiting response of 2F4MA48 (blue) and 2F4MA96 (red)
8
σ2 (10−

under 532 nm continuous wave laser domain. The continuous lines are the
9.06
1.63

theoretical fits, and the scattered plots are the experimental data.
esu)

Table 7
27

The optical limiting threshold (FOL) of 2F4MA48 and 2F4MA96 with several
γ(10−

3.98
0.53

reported chalcone derivatives under CW regime at 532 nm.


Compound Crystal Name C FOL (kJ Ref.
(M) cm− 2)
esu)

2F4MA48 1-(2-fluoro-4-methoxyphenyl)-3-(4- 0.01 5.32 Present


8

methylthiophen-2-yl)prop-2-en-1-
χ(3) (10−

8.04
1.08

one
Nonlinear optical parameters of 2F4MA48 and 2F4MA96 in DMF solution under 532 nm continuous wave domain.

2F4MA96 3-(2,4-dimeth ylphenyl)-1-(2-fluoro- 0.01 5.55 Present


4-methoxyphenyl)prop-2-en-1-one
F3BC 1-(3-bromophenyl)-3-(4- 0.01 7.41 [100]
esu)

fluorophenyl) prop-2-en-1-one
DA-ANC 1-(anthracen-9-yl)-3-(4- 0.01 6.47 [101]
8
ℑ (χ (3) )(10−

(dimethylamino)phenyl)prop-2-en-
1-one
3.23
0.58

3CAMC 1-(4-aminophenyl)-3-(3- 0.01 6.13 [102]


chlorophenyl) prop-2-en-1-one
3CAMC 1-(4-aminophenyl)-3-(3- 0.05 2.84 [102]
chlorophenyl) prop-2-en-1-one
esu)

B5B2SC 1-(5-bromo-2-thienyl)-3-(4- 0.01 1.67 [103]


8

bromophenyl) prop-2-en-1-one
ℜ (χ (3) )(10−

3C5B2SC 1-(5-bromo-2-thienyl)-3-(3- 0.01 7.38 [103]


chlorophenyl) prop-2-en-1-one
-7.36
-0.91

2M5B2SC 1-(5-bromo-2-thienyl)-3-(2- 0.01 3.17 [103]


methoxyphenyl) prop-2-en-1-one
TTMP 1-(thiophen-2-yl)-3-(3, 4, 5-trime­ 0.01 5.38 [29]
thoxyphenyl) prop-2-en-1-one
cmW− 1 )

I 3-(4-bromo-2-fluorophenyl)-1-(3,4- 0.02 5.45 [104]


dimethoxyphenyl)prop-2-en-1-one
II 3-(4-bromophenyl)-1-(3,4- 0.02 9.89 [104]
10

dimethoxyphenyl)prop-2-en-1-one
n2 (10−

-14.1
-1.74

i 2-chloro-4-fluoro-2’-hydroxy-4’,5’- 0.01 20.20 [18]


methylchalcone
ii 2-chloro-4-fluoro-2’,5’- 0.01 17.80 [18]
cmW− 1 )

dichlorothienyl chalcone
iii 2-chloro-4,6’-difluoro-4’- 0.01 11.10 [18]
methoxychalcone
5
α2 (10−

14.60
2.62

values for two other 5-bromo-2-thienyl based chalcones; 3.17 cm− 2


(2M5B2SC) and 7.38 kJ cm− 2 (3C5B2SC). Meanwhile, D. Haleshappa
et al. reported a value of 5.38 kJ cm− 2 for 1-(thiophene-2-yl)-3-(3, 4, 5-
α0 (m− 1 )

trimethoxyphenyl)prop-2-en-1-one (TTMP), which is slightly larger


2.52
0.80

than the thienyl based chalcone in this study (2F4MA48). For phenyl
based chalcone, previously reported values in literature include 7.41 kJ
Compound

cm− 2 for 1-(3-bromophenyl)-3-(4-fluorophenyl)prop-2-en-1-one


2F4MA48
2F4MA96

(F3BC), 6.47 kJ cm− 2 for 1-(anthracen-9-yl)-3-(4-(dimethylamino)


Table 6

phenyl)prop-2-en-1-one (DA-ANC), 6.13 kJ cm− 2 (0.01 M) and 2.84

13
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

(0.05 M) for 1-(4-aminophenyl)-3-(3-chlorophenyl)prop-2-en-1-one polarizability 〈α〉, anisotropic polarizability (Δα), projection of first
(3CAMC), 5.45 kJ cm− 2 (0.02 M) for 3-(4-bromo-2-fluorophenyl)-1- hyperpolarizability on dipole moment (βvec ), and the average of second
(3,4-dimethoxyphenyl)prop-2-en-1-one (I), 9.89 kJ cm− 2 (0.02 M) for 3- hyperpolarizability 〈γ〉 (for related formulas, see Appendix B).
(4-bromophenyl)-1-(3,4-dimethoxyphenyl)prop-2-en-1-one (II), 20.2 kJ The frequency-dependent (hyper)polarizabilities are visualized as
cm− 2 for 2-chloro-4-fluoro-2’-hydroxy-4’,5’-methylchalcone (i), and 3D trajectories in Figs. 14–16, and their values are given in Table S4-
11.10 kJ cm− 2 for 2-chloro-4,6’-difluoro-4’-methoxychalcone (iii). S10. The calculated parameters are found to exhibit consistent trend of
Additionally, A. Ekbote et al. demonstrated the effect of concentration decreasing (hyper)polarizabilities with increasing incident wavelengths,
on the optical limiting response of 3CAMC at 0.01 M and 0.05 M. and converging at static limit [19]. This relationship between (hyper)
Considering the concentration of the samples, 2F4MA96 is determined polarizabilities and external frequency is attributed to the optical reso­
to be the best optical limiter among the phenyl based chalcones in the nance effect [110]. Values calculated using B3LYP are indeed larger
comparison. Nevertheless, thienyl based chalcones exhibit a clear su­ than those calculated using CAM-B3LYP for all parameters [107]. The
periority against phenyl based chalcones. total dipole moments of 2F4MA48 and 2F4MA96 fall within the range
Figures-of-merits (FOMs) are used to assess their potential for all- of 4.57 to 5.34 Debye, with the major components directed from μy ,
optical switching applications, utilizing the nonlinear absorption coef­ along the negative y-axis (Fig. S10). In static field (frequency-inde­
ficient, α2 and nonlinear refraction coefficient, n2 (see Section 3.7.1). pendent), isotropic polarizability, 〈α0 〉 exhibits a larger value than
For a material to be considered suitable for such applications, it should anisotropic polarizability, Δα0 . However, the trend reversed in the
fulfil the two limits following relations (4) and (5): frequency-dependent field. T. Sugino et al. [111] have demonstrated
that the opposite trend is due to different contributing factors in which
n2 I0
one − photon FOM, W = ≫1 (4) 〈α0 〉 increases with the introduction of substituents while Δα is more
α0 λ
affected by the nature of elements and molecular geometry. This ex­
α2 λ plains the larger 〈α0 〉 in 2F4MA96, which contains two methyl groups,
two − photon FOM, T = ≪1 (5) and the larger Δα in 2F4MA48 due to its higher planarity and
n2
electron-rich thienyl ring. It is also worth noting that the experimentally
The studied compounds are determined to have satisfied the two
determined third-order susceptibility, χ (3) , is directly related to the
limits, and therefore are considered as potential materials for all-optical
anisotropic polarizability, which arises from the reorientation of the
switching devices operating under CW laser input.
molecule with respect to the field direction [112,113].
The static and dynamic first hyperpolarizabilities of electro-optic
3.7.3. Static and Dynamic (Hyper)polarizabilities
Pockel’s effect (EOPE), β(− ω; ω, 0) and second harmonic generation
To further investigate the relationship between molecular structure
(SHG) of first hyperpolarizability, β(− 2ω; ω, ω) are presented in Fig. 15.
and optical nonlinearity, DFT method is employed. When calculating
βEOPE exhibits a concave decrease with a steep slope, indicating a strong
(hyper)polarizabilities of chalcones, it is commonly reported in the
correlation between β(− ω; ω, 0) response and external frequency.
literature that the B3LYP functional is used. However, the long-range
Meanwhile, apart from the abrupt peaks at 637 and 655 nm, βSHG is
corrected B3LYP, as proposed by Tawada et al. [61], referred as
suggested to have a weaker relationship with external frequency, as
CAM-B3LYP, is known for its relatively better performance in estimating
hyperpolarizabilities due to its higher percentage of the exchange en­ opposed to βEOPE . The abrupt peaks of high NLO response, occurring at
ergy term [105,106]. Nonetheless, both functionals are associated with wavelengths approximately double of that of the band I absorption
different issues. B3LYP is known to overestimate (hyper)polarizabilities (Table 4), align with the two-state model postulation, associated with
[107], while CAM-B3LYP is associated with issues related to low-lying 2PA process. This postulation posits that a higher NLO response can be
charge transfer (CT) energies [108,109]. Therefore, both functionals achieved when the incident wavelength is twice of the maximum ab­
are employed to study the static and dynamic (hyper)polarizabilities of sorption wavelength [114].
2F4MA48 and 2F4MA96. The dynamic NLO parameters are computed The static and dynamic second hyperpolarizabilities of electro-optic
at 532, 637, 655, 800, 1064, 1274 and 999999 nm incident wave­ Kerr effect (DC-Kerr), γ( − ω; ω, 0, 0), and electric field-induced second
lengths. The DFT-estimated parameters are dipole moment (μ), isotropic harmonic generation (EFSHG), γ( − 2ω; ω, ω, 0), are depicted in

Fig. 14. (a) Isotropic and (b) anisotropic static and dynamic polarizabilities of 2F4MA48 and 2F4MA96.

14
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Fig. 15. The projection of first hyperpolarizability on dipole moment (βvec ) of 2F4MA48 and 2F4MA96 for (a) electron-optic Pockels effects (EOPE) β(− ω; ω, 0) and
(b) second harmonic generation (SHG) β(− 2ω; ω, ω) with the static hyperpolarizability β(0; 0, 0). Some of βvec (− 2ω; ω, ω) values are negative due to the deviation of
the total hyperpolarizability away from the dipole moment and are omitted in the graph for clarity.

Fig. 16. For electro-optic effects, γKerr response displays values ranging (2F4MA96) using Tauc’s plot method. Consistent with the smaller Eg,
from 17.21 × 10− 35 to 150.47 × 10− 35 esu, which are found to be 2F4MA48 demonstrated greater optical nonlinearity in terms of
weaker than βEOPE with values ranging from 26.60− 30 to 129.77 × 10− 30 nonlinear absorption (NLA), nonlinear refraction (NLR) and optical
esu, as predicted [115]. Similar to βSHG , another process related to limiting responses. In both compounds, NLA responses are primarily
second harmonic generation, the abrupt peaks of γEFSHG are also asso­ contributed by genuine 2PA process. The greater optical nonlinearity
ciated with 2PA process. For comparison with experimentally obtained response of 2F4MA48 is further supported by density functional theory
parameters, the intensity dependent refractive index (IDRI), (DFT) studies. The theoretical analysis, based on ESP, ELF-π, and FMOs
γ( − ω; ω, ω, − ω) and third harmonic generation (THG), γ( − 3ω; ω, ω, analyses, revealed that the incorporation of a thienyl ring in 2F4MA48
ω) are calculated, utilizing static second hyperpolarizability, γ0 and resulted in greater polarity, an enhanced push-pull effect, lower
aromaticity, a higher degree of electron delocalization, and a smaller
electro-optic Kerr effect, γKerr . The relations are given by Eqs. (6) and (7)
HOMO-LUMO energy gap compared to 2F4MA96, which incorporates a
[116,117]:
phenyl ring. The third-order NLO susceptibility (χ (3) ) is determined to be
〈γ( − ω; ω, ω, − ω)〉 ≅ 2〈γ( − ω; ω, 0, 0)〉 − 〈γ(0; 0, 0, 0)〉 (6) 8.04 × 10− 8 esu for 2F4MA48 and 1.08 × 10− 8 esu for 2F4MA96. The
second hyperpolarizability (γ) is calculated as 3.98 × 10− 27 esu and 0.53
〈γ( − 3ω; ω, ω, − ω)〉 ≅ 6〈γ( − ω; ω, 0, 0)〉 − 5〈γ(0; 0, 0, 0)〉 (7) × 10− 27 esu for 2F4MA48 and 2F4MA96, respectively. The onset
The calculated dynamic γ values at 532 nm are 26.50 × 10
IDRI
esu − 34 limiting threshold (FS) for 2F4MA48 and 2F4MA96 is 2.63 and 5.32 kJ
for 2F4MA48 and 25.27 × 10− 34 esu for 2F4MA96 using B3LYP func­ cm− 2, respectively while the optical limiting threshold is 5.23 and 5.55
tional. In CAM-B3LYP, the values are 8.79 × 10− 34 esu (2F4MA48) and kJ cm− 2, respectively.
8.56 × 10− 34 esu (2F4MA96). These computed values are considerably
smaller than the experimental values, which are 3.98 × 10− 27 esu for CRediT authorship contribution statement
2F4MA48 and 0.53 × 10− 27 esu for 2F4MA96. Despite the difference in
magnitude, the trend remains consistent in which 2F4MA48 shows a Nur Aisyah Mohamad Daud: Writing – original draft, Visualization,
better NLO response than 2F4MA96. However, it is worth noting that Software, Methodology, Data curation, Conceptualization. Qin Ai
the larger γ obtained from Z-scan experiment may be influenced by Wong: Writing – review & editing, Validation, Methodology, Concep­
environmental factors contributing to NLO response, such as thermal tualization. Bi Sheng Ooi: Methodology, Data curation. Ching Kheng
effects or sample purity [118]. Nonetheless, it is further supported Quah: Writing – review & editing, Validation, Supervision, Software,
through DFT-estimated (hyper)polarizabilities, the superiority of Resources, Project administration, Funding acquisition, Conceptualiza­
2F4MA48 in term of optical nonlinearity as opposed to 2F4MA96. tion. Farah Diana Ramzi: Methodology, Data curation. Yip-Foo Win:
Resources. Parutagouda Shankaragouda Patil: Resources.
4. Conclusion
Declaration of competing interest
Two fluoro-methoxy substituted chalcone derivatives with different
cinnamoyl systems; methylthienyl (2F4MA48) and dimethylphenyl The authors declare the following financial interests/personal re­
(2F4MA96) were synthesized to study the impact of homocyclic and lationships which may be considered as potential competing interests:
heterocyclic rings on photophysical responses. These compounds are in Quah Ching Kheng reports financial support was provided by Min­
donor-acceptor-donor (D-A-D) configuration, as indicated by the elec­ istry of Higher Education Malaysia for Fundamental Research Grant
trostatic potential (ESP) maps. The band I absorption peak of 2F4MA48 Scheme (FRGS) with Project Code: Ref: FRGS/1/2023/STG07/USM/02/
occurred at 351 nm, slightly redshifted compared to 2F4MA96 at 329 4.
nm. Interfragmentary charge transfer (IFCT) analysis revealed that band
I is associated with the cinnamoyl system in both compounds. The op­
tical energy gap (Eg) is estimated to be 3.21 eV (2F4MA48) and 3.42 eV

15
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

Fig. 16. Average of cubic hyperpolarizability 〈γ〉 of 2F4MA48 and 2F4MA96 for (a) electro-optic Kerr effect (DC-Kerr) γ( − ω; ω, 0, 0), (b) electric field-induced
second harmonic generation (EFSHG) γ( − 2ω; ω, ω, 0), (c) third-harmonic generation (THG) γ( − 3ω; ω, ω, ω) and (d) intensity-dependent refractive index
(IDRI) γ( − ω; ω, ω, − ω), with the static cubic hyperpolarizability γ(0; 0, 0, 0). Some of γ( − 2ω; ω, ω, 0), values are negative due the direction away from dipole
moment and are omitted in the graph for clarity.

Data Availability Acknowledgements

Data will be made available on request. The authors thank Ministry of Higher Education Malaysia for
Fundamental Research Grant Scheme (FRGS) with Project Code: Ref:
FRGS/1/2023/STG07/USM/02/4 to conduct this work. The authors
also thank Universiti Sains Malaysia (USM) for the research facilities.

Supplementary materials

Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.chphi.2024.100565.

Appendix

A.1 Sheik Bahae formalism of Z-scan experiment


To determine the third-order nonlinear optical parameters, the open-aperture (OA) and closed-aperture (CA) Z-scan data are fitted to Sheik Bahae
formalism [119], following Eq.s A.1 and A.2, respectively:

16
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

α2 I0 Leff
TOA = 1 − [( )2 ] (A.1)
3
z
zR
+ 1 (2)2

( )
4Δϕ0 zzR
TCA = 1 + [( ) 2 ][( )
2
] (A.2)
z z
zR
+ 9 zR
+ 1

2P 2 1 α0 l )
where TOA (TCA ) is the normalized transmittance of OA (CA) Z-scan data, I0 = πω 2 is the focus intensity in W/cm , Leff = α (1 − e
0

is the effective
0
πω20
length of the sample, zR = λ is the Rayleigh range, z is the distance between the sample and the focus, α0 is the linear absorption coefficient of the
sample in cm− 1, l is the sample length, ω0 is the beam waist radius at the focus, λ is the wavelength of incident laser and Δϕ0 is the one-axis phase shift
at the focus.
Through Eq. A.1, nonlinear absorption (NLA) coefficient, α2 (in cm2W− 1) is directly extracted. Meanwhile, through Eq. A.2, Δϕ0 is determined and
subsequently, nonlinear refraction (NLR) coefficient, n2 is calculated using Eq. A.3:
( )( )
λ Δϕ0 ( )
n2 = in cm2 W − 1 (A.3)
2π I0 Leff

With α2 value, the absorption cross section of two-photon absorption (2PA) process is determined by using Eq. A.4:
( )

σ2 = α2 (in GM) (A.4)
NC

where hν is the energy of the excited photon in Joules, NC = NAC is the molecular number density in unit volume, NA is Avogadro’s number and C is the
concentration of the sample solution. (1 GM is 10− 50 cm4s photon− 1)
To determine the excited-state absorption cross-section, σex (in cm2), OA Z-scan data is fitted to Eq. A.5:
⎡ ⎤ ⎡ ⎤
/
⎢ x0 ⎥ ⎢ x0 ⎥
⎢ ⎥ ⎢ ⎥
TOA = ln⎢1 + ( )2 ⎥ ⎢ ( )2 ⎥ (A.5)
⎣ ⎦ ⎣ ⎦
1 + zzR 1 + zzR

0 F0 Leff
where x0 = σex α2h ν , F0 is the on-axis fluence at focus.
Then, for ground-state absorption cross-section, σg , the calculation is conducted following Eq. A.6:
α ( )
σ g = 0 in cm2 (A.6)
NC

Using n2 value, the thermal-optical coefficient, dT


dn
(in K− 1 ) is calculated with Eq. A.7 to compare the significance of thermal lensing in the sample:
( )( ) ( )( )
dn I α0 τ dn α0 ω20
n2 = = (A.7)
dT ρc dT 4κ

where I is the incident intensity, ρ is the sample density, c is the specific heat capacity, τ is the thermal relaxation time, and κ = 0.184 Wm− 1 K− 1 is the
thermal conductivity of DMF (solvent used for Z-scan experiment) at room temperature.
The α2 and n2 values are further used to determine the real and imaginary parts of third-order susceptibility (ℜ(χ (3) ) and ℑ(χ (3) )) using Eqs. A.8 and
A.9, respectively:
(( )( ) 4ε0 cn20 ( )
ℜ χ (3) ( − ω; ω, ω, − ω) in m2 V − 2 = 10− 4 n2 in cm2 W − 1 (A.8)
3
( )( ) 2ε0 c2 n20 ( )
ℑ χ (3) ( − ω; ω, ω, − ω) in m2 V − 2 = 10− 2 α2 in cm W − 1 (A.9)

where ε0 is the permittivity of free space, c is the speed of light, n0 = 1.4305 is the linear refractive index of DMF at room temperature, and ω = 2λπc is
the angular frequency of light field.
Since the calculated real and imaginary parts of χ (3) are in SI unit (m2 V− 2), unit conversion is conducted from SI unit to gaussian unit (esu)
following Eq. A.10:
( ) 4π
χ (3) in m2 V − 2
=( )2 χ
(3)
(in esu) (A.10)
3 × 104

B. Static and dynamic (hyper)polarizabilities [19]


Taylor expansion of energy if a system in terms of external electric field, F is given as relation B.1.

17
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

∑ 1 ∑∑ 1 ∑∑∑ 1 ∑∑∑∑
E(F) = E(0) − μi F i − αij Fi Fj − β Fi Fj Fk − γ Fi Fj Fk Fl − …, i, j, k, l = {x, y, z} (B.1)
i
2! i j 3! i j k ijk 4! i j k l ijkl

Where μi , αij , βijk , and γijkl are tensor components of dipole moment, polarizability, first and second hyperpolarizabilities.
The magnitude of electric dipole moment is given by relation B.2
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
|μ| = μ2x + μ2y + μ2z (B.2)

The isotropic and anisotropic polarizability are following relations B.3 and B.4, respectively.
1( )
〈α〉 = αxx + αyy + αzz (B.3)
3
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )̅
1 ( )2 ( )2
Δα = √̅̅̅ αxx − αyy + αyy − αzz + (αzz − αxx )2 + 6 α2xy + α2yz + α2xz (B.4)
2
The magnitude of first hyperpolarizability, β and the projection on dipole moment vector are given by relations B.5 and B.6, respectively:
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
|β| = β2x + β2y + β2z (B.5)

∑μ βi
βvec = i
(B.6)
i
|μ|

Where βi (i = x, y, z) is given by equation B.7,


1 ∑( )
βi = β + βjij + βjji (B.7)
3 j=x,y,z ijj

The averaged isotropic second hyperpolarizability, 〈γ〉 is following relation B.8:


1[ ( )]
〈γ〉 = γ + γyyyy + γ zzzz + 2 γ xxyy + γxxzz + γ yyzz (B.8)
5 xxxx

References [12] G. Peramaiyan, P. Pandi, N. Vijayan, G. Bhagavannarayana, R.M. Kumar, Studies


on growth, structural, optical and mechanical properties of xylenol orange dye
admixtured l-arginine phosphate single crystal, Optik (Stuttg) 124 (2013)
[1] J. Kunjumon, M. George, A. K, S.K. Gopi, G. Vinitha, D. Sajan, R. Philip, Tunable
4058–4063.
photoluminescence emission and third order optical nonlinearity in Sm2/
[13] R.V. Rajan, M. George, D.R. Leenaraj, R. Ittyachan, D. Sajan, G. Vinitha, Growth,
3Cu3Ti4O12 ceramics for optical limiting applications, Surfaces and Interfaces 39
Z-scan and density functional theoretical study for investigating the nonlinear
(2023) 102965.
optical properties of guanidinium l-glutamate for optical limiting applications,
[2] M. Rana, N. Singla, A. Chatterjee, A. Shukla, P. Chowdhury, Investigation of
J Mol Struct 1222 (2020) 128937.
nonlinear optical (NLO) properties by charge transfer contributions of amine
[14] J. Mathai, A.K Jose, M.P. Anjana, P.A. Aleena, J. Kunjumon, R. Ittyachan, S. S.
functionalized tetraphenylethylene, Opt Mater (Amst) 62 (2016) 80–89.
Nair, G. Vinitha, D. Sajan, Substantial effect of Cr doping on the third-order
[3] R. Perkins, New Optical Switch Could Lead to Ultrafast All-Optical Signal
nonlinear optical properties of ZnO nanostructures, Opt Mater (Amst) 142 (2023)
Processing.
114128.
[4] M. Y. Kariduraganavar, R. V Doddamani, B. Waddar and S. R. Parne, in Nonlinear
[15] T. Agag, A. Akelah, Handbook of Benzoxazine Resins, Elsevier, 2011,
Optics, eds. İ. Bakırtaş and N. Antar, IntechOpen, Rijeka, 2021, p. Ch. 9.
pp. 495–516.
[5] L. Buglioni, F. Raymenants, A. Slattery, S.D.A. Zondag, T. Noël, Technological
[16] M. George, D. Sajan, P. Sankar, R. Philip, Vibrational spectra, linear and
Innovations in Photochemistry for Organic Synthesis: Flow Chemistry, High-
nonlinear optical investigations on 3chloro 4-fluoro aniline and 2-iodo aniline for
Throughput Experimentation, Scale-up, and Photoelectrochemistry, Chem Rev
optical limiting applications, J Mol Struct 1238 (2021) 130412.
122 (2022) 2752–2906.
[17] Q. Xie, Z. Shao, Y. Zhao, L. Yang, Q. Wu, W. Xu, K. Li, Y. Song, H. Hou, Novel
[6] K. Xu, T. Wang, R. Chen, J. Ma, X. Mu, D. Zhong, L. Cao, B. Teng, Design Novel
photo-controllable third-order nonlinear optical (NLO) switches based on
Chalcone Crystals for Enhancement of Nonlinear Optical Activity and Higher
azobenzene derivatives, Dyes and Pigments 170 (2019) 107599.
Molecular Structural Stability, Cryst Growth Des 23 (2023) 592–601.
[18] Q.A. Wong, C.K. Quah, X.A. Wong, S.R. Maidur, H.C. Kwong, Y.-F. Win, P.S. Patil,
[7] R. Sebastian, S. Sankararaman, Development and nonlinear optical
N.B. Gummagol, Structure-Property Relationship of Three 2-Chloro-4-fluoro
characterization of phthalocyanine-incorporated stable natural dye with
Chalcone Derivatives: A Comprehensive Study on Linear and Non-linear Optical
wideband absorption for solar cell applications, Journal of the Optical Society of
Properties, Structural Characterizations and Density Functional Theory, J Mol
America B 37 (2020) 110–116.
Struct 1267 (2022) 133584.
[8] I.D. Borges, J.A.V Danielli, V.E.G. Silva, L.O. Sallum, J.E. Queiroz, L.D. Dias,
[19] Q.A. Wong, C.K. Quah, X.A. Wong, Y.-F. Win, H.C. Kwong, P.S. Patil, N.
I. Iermak, G.L.B. Aquino, A.J. Camargo, C. Valverde, F.A.P. Osório, B. Baseia, H.
B. Gummagol, R.S. V, A comprehensive study on the photophysical and non-
B. Napolitano, Synthesis and structural studies on (E)-3-(2,6-difluorophenyl)-1-
linear optical properties of thienyl-chalcone derivatives, Physical Chemistry
(4-fluorophenyl)prop-2-en-1-one: a promising nonlinear optical material, RSC
Chemical Physics 24 (2022) 21927–21953.
Adv 10 (2020) 22542–22555.
[20] M. George, R. Ida Malarselvi, A. Thiruvalluvar, P. Dominic, D. Sajan, R. Philip,
[9] S.R. Maidur, P.S. Patil, N.K. Katturi, V.R. Soma, Q.A. Wong, C.K. Quah, Ultrafast
Linear and nonlinear optical investigations of (E) − 4 methyl-2-(N-
Nonlinear Optical and Structure–Property Relationship Studies of Pyridine-Based
Phenylcarboximidoyl) phenol supported by vibrational spectral analysis for
Anthracene Chalcones Using Z-Scan, Degenerate Four-Wave Mixing, and
photonic applications, Materials Science and Engineering: B 283 (2022) 115860.
Computational Approaches, J Phys Chem B 125 (2021) 3883–3898.
[21] D.S. Chemla, J. Zyss, Nonlinear Optical Properties of Organic Molecules and
[10] P.S. Menon, J. Kunjumon, M. Bansal, S.S. Nair, C. Beryl, G. Vinitha, T. Maity, P.
Crystals, Academic Press, 1987.
M. Abraham, D. Sajan, R. Philip, Role of surface defects in the third order
[22] P.S. Menon, M.P. Anjana, A.K. Jose, J. Kunjumon, A.P. A, S. Chandran,
nonlinear optical properties of pristine NiO and Cr doped NiO nanostructures,
M. George, G. Vinitha, D. Sajan, The role of defects on linear and nonlinear
Ceram Int 49 (2023) 5815–5827.
optical properties of pristine and nickel doped zinc oxide nanoparticles, Surfaces
[11] P.S. Menon, J. Kunjumon, A.K. Jose, A.P. A, M. Bansal, G. Vinitha, T. Maity, P.
and Interfaces 34 (2022) 102393.
M. Abraham, D. Sajan, S.D. George, Structural, Third Order Nonlinear and
[23] J. Rojas, J.N. Domı́nguez, J.E. Charris, G. Lobo, M. Payá, M.L. Ferrándiz,
magnetic properties of pristine and Ni-doped CuO nanoparticles: Diluted
Synthesis and inhibitory activity of dimethylamino-chalcone derivatives on the
magnetic semiconductors, Colloids Surf A Physicochem Eng Asp 650 (2022)
induction of nitric oxide synthase, Eur J Med Chem 37 (2002) 699–705.
129582.
[24] H.R. Puranik, H.J. Ravindra, AIP Conf. Proc 29 (2019) 020048.

18
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

[25] J. Jia, Y. Li, J. Gao, A series of novel ferrocenyl derivatives: Schiff bases-like push- analysis, visualization and quantitative analysis of molecular crystals, J Appl
pull systems with large third-order optical responses, Dyes and Pigments 137 Crystallogr 54 (2021) 1006–1011.
(2017) 342–351. [51] M. Sheik-Bahae, A.A. Said, T.-H. Wei, D.J. Hagan, E.W. Van Stryland, Sensitive
[26] D.A. Zainuri, M. Abdullah, M.F. Zaini, H. Bakhtiar, S. Arshad, I. Abdul Razak, measurement of optical nonlinearities using a single beam, IEEE J Quantum
Fused ring effect on optical nonlinearity and structure property relationship of Electron 26 (1990) 760–769.
anthracenyl chalcone based push-pull chromophores, PLoS One 16 (2021) [52] M. Sheik-bahae, A.A. Said, E.W. Van Stryland, High-sensitivity, single-beam n_2
e0257808. measurements, Opt Lett 14 (1989) 955.
[27] S. Muhammad, A.G. Al-Sehemi, A. Irfan, A.R. Chaudhry, H. Gharni, S. AlFaify, [53] M.A. Spackman, J.J. McKinnon, Fingerprinting intermolecular interactions in
M. Shkir, A.M. Asiri, The impact of position and number of methoxy group(s) to molecular crystals, CrystEngComm 4 (2002) 378–392.
tune the nonlinear optical properties of chalcone derivatives: a dual substitution [54] D. Jayatilaka and D. J. Grimwood, in Computational Science — ICCS 2003, eds. P.
strategy, J Mol Model 22 (2016) 73. M. A. Sloot, D. Abramson, A.V Bogdanov, Y. E. Gorbachev, J. J. Dongarra and A.
[28] M. Albota, D. Beljonne, J.-L. Brédas, J.E. Ehrlich, J.-Y. Fu, A.A. Heikal, S.E. Hess, Y. Zomaya, Springer Berlin Heidelberg, Berlin, Heidelberg, 2003, pp. 142–151.
T. Kogej, M.D. Levin, S.R. Marder, D. McCord-Maughon, J.W. Perry, H. Röckel, [55] J.J. McKinnon, M.A. Spackman, A.S. Mitchell, Novel tools for visualizing and
M. Rumi, G. Subramaniam, W.W. Webb, X.-L. Wu, C. Xu, Design of Organic exploring intermolecular interactions in molecular crystals, Acta Crystallogr B 60
Molecules with Large Two-Photon Absorption Cross Sections, Science (1979) 281 (2004) 627–668.
(1998) 1653–1656. [56] J.J. McKinnon, D. Jayatilaka, M.A. Spackman, Towards quantitative analysis of
[29] D. Haleshappa, A. Jayarama, R. Bairy, S. Acharya, P.S. Patil, Second and third intermolecular interactions with Hirshfeld surfaces, Chemical Communications
order nonlinear optical studies of a novel thiophene substituted chalcone (2007) 3814.
derivative, Physica B Condens Matter 555 (2019) 125–132. [57] R. Dennington, T. Keith and J. Millam, GaussView 5.0, Gaussian, Inc.,
[30] M. Sai Kiran, B. Anand, S. Siva Sankara Sai, G. Nageswara Rao, Second- and third- Wallingford.
order nonlinear optical properties of bis-chalcone derivatives, J Photochem [58] A.D. Becke, Density-functional thermochemistry. III. The role of exact exchange,
Photobiol A Chem 290 (2014) 38–42. J Chem Phys 98 (1993) 5648–5652.
[31] Y. Ju, X. Wu, W. Zhou, J. Jia, J. Yang, Y. Sun, Y. Song, Two novel thiophene [59] Z.-L. Cai, M.J. Crossley, J.R. Reimers, R. Kobayashi, R.D. Amos, Density
derivatives based on chalcone skeleton:a systematic study of ultrafast nonlinear Functional Theory for Charge Transfer: The Nature of the N-Bands of Porphyrins
absorption and excited-state dynamics, Opt Mater (Amst) 114 (2021) 110969. and Chlorophylls Revealed through CAM-B3LYP, CASPT2, and SAC-CI
[32] E. Mathew, V.V. Salian, B. Narayana, I.H. Joe, Experimental and theoretical Calculations, J Phys Chem B 110 (2006) 15624–15632.
approach on third-order optical nonlinearity of a highly efficient anthracene- [60] A.J. Cohen, P. Mori-Sánchez, W. Yang, Challenges for Density Functional Theory,
based chalcone derivative for optical power limiting, J Mol Struct 1250 (2022) Chem Rev 112 (2012) 289–320.
131704. [61] Y. Tawada, T. Tsuneda, S. Yanagisawa, T. Yanai, K. Hirao, A long-range-corrected
[33] S. Liu, Y. Zhang, K. Sun, B. Graff, P. Xiao, F. Dumur, J. Lalevée, Design of time-dependent density functional theory, J Chem Phys 120 (2004) 8425–8433.
photoinitiating systems based on the chalcone-anthracene scaffold for LED [62] M.P. Andersson, P. Uvdal, New Scale Factors for Harmonic Vibrational
cationic photopolymerization and application in 3D printing, Eur Polym J 147 Frequencies Using the B3LYP Density Functional Method with the Triple-ζ Basis
(2021) 110300. Set 6-311+G(d,p), J Phys Chem A 109 (2005) 2937–2941.
[34] V. Kumar, P. Kumar, P. Kaur, K. Singh, A bis-pyrene chalcone based fluorescent [63] M.H. Jamróz, Vibrational Energy Distribution Analysis (VEDA): Scopes and
material for ratiometric sensing of hydrazine: An acid/base molecular switch and limitations, Spectrochim Acta A Mol Biomol Spectrosc 114 (2013) 220–230.
solid-state emitter, Anal Chim Acta 1178 (2021) 338807. [64] R.E. Stratmann, G.E. Scuseria, M.J. Frisch, An efficient implementation of time-
[35] S.K. Alsaee, M.A. Abu Bakar, D.A. Zainuri, A.H. Anizaim, M.F. Zaini, M.M. Rosli, dependent density-functional theory for the calculation of excitation energies of
M. Abdullah, S. Arshad, I. Abdul Razak, Nonlinear optical properties of pyrene- large molecules, J Chem Phys 109 (1998) 8218–8224.
based chalcone: (E)-1-(4′-bromo-[1,1′-biphenyl]-4-yl)-3-(pyren-1-yl)prop-2-en-1- [65] M.E. Casida, C. Jamorski, K.C. Casida, D.R. Salahub, Molecular excitation
one, a structure-activity study, Opt Mater (Amst) 128 (2022) 112314. energies to high-lying bound states from time-dependent density-
[36] J. Li, J. Xu, X. Yang, L. Ren, Y. Wang, D. Ma, P. Fan, H. Wang, L. Liu, B. Dong, functional response theory: Characterization and correction of the time-
Q. Chen, T. Wu, Effects of phenanthrene on the essential oil composition and leaf dependent local density approximation ionization threshold, J Chem Phys 108
metabolome in peppermint plants (Mentha piperita L.), Ind Crops Prod 187 (1998) 4439–4449.
(2022) 115383. [66] C. Lee, W. Yang, R.G. Parr, Development of the Colle-Salvetti correlation-energy
[37] S. Tandel, N.C. Patel, S. Kanvah, P.N. Patel, An efficient protocol for the synthesis formula into a functional of the electron density, Phys Rev B 37 (1988) 785–789.
of novel hetero-aryl chalcone: A versatile synthon for several heterocyclic [67] T. Lu, F. Chen, Multiwfn: A multifunctional wavefunction analyzer, J Comput
scaffolds and sensors, J Mol Struct 1269 (2022) 133808. Chem 33 (2012) 580–592.
[38] Mohd. Shkir, S. Muhammad, S. AlFaify, A. Irfan, M.A. Khan, A.G. Al-Sehemi, I. [68] W. Humphrey, A. Dalke, K. Schulten, VMD: Visual molecular dynamics, J Mol
S. Yahia, B. Singh, I. Bdikin, A comparative study of key properties of glycine Graph 14 (1996) 33–38.
glycinium picrate (GGP) and glycinium picrate (GP): A combined experimental [69] S. Zeroual, J. Meinnel, A. Lapinski, A.Boudjada S.Parker, A. Boucekkine,
and quantum chemical approach, Journal of Saudi Chemical Society 22 (2018) Vibrational spectroscopy and DFT calculations of 1,3-dibromo-2,4,6-trimethyl­
352–362. benzene: Anharmonicity, coupling and methyl group tunneling, Vib Spectrosc 67
[39] P.J. Tejkiran, M.S. Brahma Teja, P. Sai Siva Kumar, P. Sankar, R. Philip, (2013) 27–43.
S. Naveen, N.K. Lokanath, G. Nageswara Rao, D-A-π-D Synthetic approach for [70] T. Lu, F. Chen, Quantitative analysis of molecular surface based on improved
thienyl chalcones – NLO – a structure activity study, J Photochem Photobiol A Marching Tetrahedra algorithm, J Mol Graph Model 38 (2012) 314–323.
Chem 324 (2016) 33–39. [71] T. Lu, S. Manzetti, Wavefunction and reactivity study of benzo[a]pyrene diol
[40] K.D. Katariya, K.J. Nakum, M. Hagar, New thiophene chalcones with ester and epoxide and its enantiomeric forms, Struct Chem 25 (2014) 1521–1533.
Schiff base mesogenic Cores: Synthesis, mesomorphic behaviour and DFT [72] Y.M. Dikova, D.S. Yufit, J.A.G. Williams, Platinum(IV) Complexes with
investigation, J Mol Liq 359 (2022) 119296. Tridentate, NNC-Coordinating Ligands: Synthesis, Structures, and Luminescence,
[41] S.A. Ponomarenko, Y.N. Luponosov, J. Min, A.N. Solodukhin, N.M. Surin, M. Inorg Chem 62 (2023) 1306–1322.
A. Shcherbina, S.N. Chvalun, T. Ameri, C. Brabec, Design of donor–acceptor star- [73] S.L. Olmsted, P. Tongcharoensirikul, E. McCaskill, K. Gandiaga, D. Labaree, R.
shaped oligomers for efficient solution-processible organic photovoltaics, Faraday B. Hochberg, R.N. Hanson, Synthesis and evaluation of 17α-E-20-(heteroaryl)
Discuss 174 (2014) 313–339. norpregn-1,3,5(10),20 tetraene-3,17β-diols [17α-(heteroaryl)vinyl estradiols] as
[42] H. Hayasaka, T. Miyashita, M. Nakayama, K. Kuwada, K. Akagi, Dynamic ligands for the estrogen receptor-α ligand binding domain (ERα-LBD), Bioorg Med
Photoswitching of Helical Inversion in Liquid Crystals Containing Chem Lett 22 (2012) 977–979.
Photoresponsive Axially Chiral Dopants, J Am Chem Soc 134 (2012) 3758–3765. [74] T. Lu, Q. Chen, A simple method of identifying π orbitals for non-planar systems
[43] M. O’Neill, S.M. Kelly, Ordered Materials for Organic Electronics and Photonics, and a protocol of studying π electronic structure, Theor Chem Acc 139 (2020) 25.
Advanced Materials 23 (2011) 566–584. [75] G. Zhang, C.B. Musgrave, Comparison of DFT Methods for Molecular Orbital
[44] P.K. Hegde, A. Vasudeva Adhikari, M.G. Manjunatha, C. Suchand Sandeep, Eigenvalue Calculations, J Phys Chem A 111 (2007) 1554–1561.
R. Philip, Novel poly(3,4-dialkoxythiophene)s carrying 1,3,4-oxadiazolyl- [76] Lu Tian, Chen Fei-Wu, Calculation of Molecular Orbital Composition, Acta Chimi
biphenyl moieties: synthesis and nonlinear optical studies, Polym Int 60 (2011) Sin 69 (2011) 2393–2406.
112–118. [77] Polar Covalent Bonds - Electronegativity, https://chem.libretexts.org/@go/p
[45] H.R. Manjunath, P.C. Rajesh Kumar, S. Naveen, V. Ravindrachary, M.A. Sridhar, age/31383, (accessed 12 October 2023).
J. Shashidhara Prasad, P. Karegoudar, Growth, characterization, crystal and [78] F. Sessa, M. Rahm, Electronegativity Equilibration, J Phys Chem A 126 (2022)
molecular structure studies of, J Cryst Growth 327 (2011) 161–166. 5472–5482.
[46] A. Bruker, Instrument Service v4. 2.7, APEX2, SADABS, SAINT-Plus & XPREP. [79] W. Leupin, J. Wirz, Low-lying electronically excited states of cycl[3.3.3]azine, a
[47] G.M. Sheldrick, A short history of SHELX, Acta Crystallogr A 64 (2008) 112–122. bridged 12.pi.-perimeter, J Am Chem Soc 102 (1980) 6068–6075.
[48] A.L. Spek, Structure validation in chemical crystallography, Acta Crystallogr D [80] H. Watanabe, K. Tanaka, Y. Chujo, Independently Tuned Frontier Orbital Energy
Biol Crystallogr 65 (2009) 148–155. Levels of 1,3,4,6,9b-Pentaazaphenalene Derivatives by the Conjugation Effect,
[49] C.F. Macrae, I.J. Bruno, J.A. Chisholm, P.R. Edgington, P. McCabe, E. Pidcock, J Org Chem 84 (2019) 2768–2778.
L. Rodriguez-Monge, R. Taylor, J. van de Streek, P.A. Wood, Mercury CSD 2.0– [81] T. Zhang, I.E. Brumboiu, C. Grazioli, A. Guarnaccio, M. Coreno, M. de Simone,
new features for the visualization and investigation of crystal structures, J Appl A. Santagata, H. Rensmo, B. Brena, V. Lanzilotto, C. Puglia, Lone-Pair
Crystallogr 41 (2008) 466–470. Delocalization Effects within Electron Donor Molecules: The Case of
[50] P.R. Spackman, M.J. Turner, J.J. McKinnon, S.K. Wolff, D.J. Grimwood, Triphenylamine and Its Thiophene-Analog, The Journal of Physical Chemistry C
D. Jayatilaka, M.A. Spackman, CrystalExplorer: a program for Hirshfeld surface 122 (2018) 17706–17717.

19
N.A. Mohamad Daud et al. Chemical Physics Impact 8 (2024) 100565

[82] L.W. Wijayanti, R.T. Swasono, W. Lee, J. Jumina, Synthesis and Evaluation of [102] A. Ekbote, P.S. Patil, S.R. Maidur, T.S. Chia, C.K. Quah, Structure and nonlinear
Chalcone Derivatives as Novel Sunscreen Agent, Molecules 26 (2021) 2698. optical properties of (E)-1-(4-aminophenyl)-3-(3-chlorophenyl) prop-2-en-1-one:
[83] A. Rammohan, J.S. Reddy, G. Sravya, C.N. Rao, G.V. Zyryanov, Chalcone A promising new D-π-A-π-D type chalcone derivative crystal for nonlinear optical
synthesis, properties and medicinal applications: a review, Environ Chem Lett 18 devices, J Mol Struct 1129 (2017) 239–247.
(2020) 433–458. [103] V.S. Naik, P.S. Patil, N.B. Gummagol, Q.A. Wong, C.K. Quah, H.S. Jayanna,
[84] Z. Liu, T. Lu, Q. Chen, An sp-hybridized all-carboatomic ring, cyclo[18]carbon: Crystal structure, linear and nonlinear optical properties of three thiophenyl
Electronic structure, electronic spectrum, and optical nonlinearity, Carbon N Y chalcone derivatives: A combined experimental and computational study, Opt
165 (2020) 461–467. Mater (Amst) 110 (2020) 110462.
[85] Z. Liu, X. Wang, T. Lu, A. Yuan, X. Yan, Potential optical molecular switch: [104] H.C. Kwong, M.S. Rakesh, C.S. Chidan Kumar, S.R. Maidur, P.S. Patil, C.K. Quah,
Lithium@cyclo[18]carbon complex transforming between two stable Y.-F. Win, C. Parlak, S. Chandraju, Structure–property relation and third-order
configurations, Carbon N Y 187 (2022) 78–85. nonlinear optical studies of two new halogenated chalcones, Z Kristallogr Cryst
[86] C.A. Guido, P. Cortona, B. Mennucci, C. Adamo, On the Metric of Charge Transfer Mater 233 (2018) 349–360.
Molecular Excitations: A Simple Chemical Descriptor, J Chem Theory Comput 9 [105] M. Kamiya, H. Sekino, T. Tsuneda and K. Hirao, Nonlinear optical property
(2013) 3118–3126. calculations by the long-range-corrected coupled-perturbed Kohn–Sham method,
[87] T. Le Bahers, C. Adamo, I. Ciofini, A Qualitative Index of Spatial Extent in Charge- J Chem Phys, 10.1063/1.1935514.
Transfer Excitations, J Chem Theory Comput 7 (2011) 2498–2506. [106] K. Anitha, V. Balachandran, Assessment of long-range corrected and conventional
[88] J. Wang, Y. Chen, R. Li, H. Dong, L. Zhang, M. Lotya, J. N, W. J, Carbon DFT functional for the prediction of second – Order NLO properties and other
Nanotubes - Synthesis, Characterization, Applications, InTech, 2011. molecular properties of N-(2-cyanoethyl)-N-butylaniline – A vibrational
[89] H. Liu, Z. Li, Y. Yu, J. Lin, S. Liu, F. Pang, T. Wang, Nonlinear optical properties of spectroscopy study, Spectrochim Acta A Mol Biomol Spectrosc 146 (2015) 66–79.
anisotropic two-dimensional layered materials for ultrafast photonics, [107] Y. Erande, M.C. Sreenath, S. Chitrambalam, I.H. Joe, N. Sekar, Spectroscopic, DFT
Nanophotonics 9 (2020) 1651–1673. and Z-scan supported investigation of dicyanoisophorone based push-pull
[90] C. Babeela, M.A. Assiri, A.G. Al-Sehemi, M. Pannipara, T.C.S. Girisun, Excited NLOphoric styryl dyes, Opt Mater (Amst) 66 (2017) 494–511.
state absorption of Cu-doped barium borate nanostructures under nanopulsed [108] Y. Shao, Y. Mei, D. Sundholm, V.R.I. Kaila, Benchmarking the Performance of
laser excitation, The European Physical Journal D 75 (2021) 102. Time-Dependent Density Functional Theory Methods on Biochromophores,
[91] B. Anand, A. Kaniyoor, S.S.S. Sai, R. Philip, S. Ramaprabhu, Enhanced optical J Chem Theory Comput 16 (2020) 587–600.
limiting in functionalized hydrogen exfoliated graphene and its metal hybrids, [109] M.S. Kodikara, R. Stranger, M.G. Humphrey, Long-Range Corrected DFT
J Mater Chem C Mater 1 (2013) 2773. Calculations of First Hyperpolarizabilities and Excitation Energies of Metal
[92] B. Binish, M. Durairaj, T.C.S. Girisun, K.M. Rahulan, Engineering the nonlinear Alkynyl Complexes, ChemPhysChem 19 (2018) 1537–1546.
optical properties of barium molybdate by doping Sn4+ ions for optical limiting [110] S. Hua, X. Wang, Z. Liu, T. Lu, M. Zhao, Effects of external field wavelength and
device applications, Ceram Int 49 (2023) 17629–17638. solvation on the photophysical property and optical nonlinearity of 1,3-thiazo­
[93] H.Y. Tan, G.L. Ong, C.H. Nee, S.L. Yap, H.S. Poh, T.Y. Tou, B.L. Lan, S.F. Lee, S. lium-5-thiolates mesoionic compound, Spectrochim Acta A Mol Biomol Spectrosc
S. Yap, Thermal-induced effects on ultrafast laser filamentation in ethanol, Opt 289 (2023) 122227.
Laser Technol 163 (2023) 109350. [111] T. Sugino, N. Kambe, N. Sonoda, T. Sakaguchi, K. Ohta, Ab initio molecular
[94] S.K. Maurya, D. Yadav, D. Goswami, Effect of femtosecond laser pulse repetition orbital calculations of the static polarizabilities of xanthone analogues, Chem
rate on nonlinear optical properties of organic liquids, PeerJ Phys Chem 1 (2019) Phys Lett 251 (1996) 125–131.
e1. [112] Saswati. Ghosal, Marek. Samoc, P.N. Prasad, J.J. Tufariello, Optical nonlinearities
[95] S.M. Mian, S.B. McGee, N. Melikechi, Experimental and theoretical investigation of organometallic structures: aryl and vinyl derivatives of ferrocene, J Phys Chem
of thermal lensing effects in mode-locked femtosecond Z-scan experiments, Opt 94 (1990) 2847–2851.
Commun 207 (2002) 339–345. [113] S. Muhammad, A.G. Al-Sehemi, A. Irfan, A.R. Chaudhry, Tuning the push–pull
[96] J.Z. Low, B. Capozzi, J. Cui, S. Wei, L. Venkataraman, L.M. Campos, Tuning the configuration for efficient second-order nonlinear optical properties in some
polarity of charge carriers using electron deficient thiophenes, Chem Sci 8 (2017) chalcone derivatives, J Mol Graph Model 68 (2016) 95–105.
3254–3259. [114] M. Pawlicki, H.A. Collins, R.G. Denning, H.L. Anderson, Two-Photon Absorption
[97] G. Vinitha, A. Ramalingam, Third-order optical nonlinearities and optical- and the Design of Two-Photon Dyes, Angewandte Chemie International Edition
limiting properties of a Pararosanilin dye in liquid and solid media, Laser Phys 18 48 (2009) 3244–3266.
(2008) 1070–1073. [115] X. Yin, T. Feng, Z. Liang, J. Li, Artificial Kerr-type medium using metamaterials,
[98] L.W. Tutt, T.F. Boggess, A review of optical limiting mechanisms and devices Opt Express 20 (2012) 8543.
using organics, fullerenes, semiconductors and other materials, Prog Quantum [116] C. Valverde, A. de L. E. S. Castro, R. G. Vaz, L. de A. J. Ferreira, B. Baseia and A. P.
Electron 17 (1993) 299–338. F. Osório, Third-Order Nonlinear Optical Properties of a Carboxylic Acid
[99] T.C. Sabari Girisun, M. Saravanan, V.R. Soma, Wavelength-Dependent Nonlinear Derivative., Acta Chim Slov, 65, 739–749.
Optical Absorption and Broadband Optical Limiting in Au-Fe 2 O 3 -rGO [117] S. Marques, M.A. Castro, S.A. Leão, T.L. Fonseca, Second hyperpolarizability of
Nanocomposites, ACS Appl Nano Mater 1 (2018) 6337–6348. the calcium-doped lithium salt of pyridazine Li–H3C4N2 ⋯ Ca, Chem Phys Lett
[100] A. Ekbote, P.S. Patil, S.R. Maidur, T.S. Chia, C.K. Quah, Structural, third-order 659 (2016) 76–79.
optical nonlinearities and figures of merit of (E)-1-(3-substituted phenyl)-3-(4- [118] J. Zhang, 2017, p. 020025.
fluorophenyl) prop-2-en-1-one under CW regime: New chalcone derivatives for [119] Kuzyk, Characterization Techniques and Tabulations for Organic Nonlinear
optical limiting applications, Dyes and Pigments 139 (2017) 720–729. Optical Materials, Routledge, 2018.
[101] J.R. Jahagirdar, S.R. Maidur, P.S. Patil, T.S. Chia, C.K. Quah, Growth,
characterizations and nonlinear optical studies of dimethylamine substituted
anthracene chalcone single crystals, J Mol Struct 1278 (2023) 134897.

20

You might also like