Muhammad Khalid*ab,
Memoona Arshadab,
Muhammad Haroonc,
Ataualpa Albert Carmo Braga
d and
Norah Alhokbanye
aInstitute of Chemistry, Khwaja Fareed University of Engineering & Information Technology, Rahim Yar Khan, 64200, Pakistan. E-mail: muhammad.khalid@kfueit.edu.pk; khalid@iq.usp.br
bCentre for Theoretical and Computational Research, Khwaja Fareed University of Engineering & Information Technology, Rahim Yar Khan, 64200, Pakistan
cDepartment of Chemistry and Biochemistry, Miami University, Oxford, OH, USA
dDepartamento de Química Fundamental, Instituto de Química, Universidade de São Paulo, Av. Prof. Lineu Prestes, 748, São Paulo, 05508-000, Brazil
eDepartment of Chemistry, College of Science, King Saud University, P. O. Box 2455, Riyadh 11451, Saudi Arabia
First published on 25th February 2025
Organic molecules are considered important NLO materials in the modern age because of their potential optoelectronic features. Therefore, a series of organic molecules (DBTD1–DBTD7) with a D–π–A architecture was designed from the reference compound (DBTR) by structurally tailoring it with effective donors and making significant modifications to the π-spacers. All derivatives and references were initially optimized at the M06 functional with a 6-311G(d,p) basis set. Furthermore, the optimized structures (DBTR and DBTD1–DBTD7) were used to determine the density of states (DOS), frontier molecular orbitals (FMOs), natural bond orbital (NBO), UV-Visible, transition density matrix (TDM) analyses, and the most significant NLO properties. Among the compounds, DBTD4 had the smallest band gap (2.882 eV) which was further supported by DOS analysis. The global reactivity parameters were also related to the HOMO–LUMO band gap, where the compound with the lowest band gap showed a lower hardness value and higher softness value. NBO analysis was used to explain the molecular stability and hyperconjugation. DBTD4 and DBTD5 demonstrated comparable low bandgaps with the highest comparable NLO parameter values. However, we also observed efficient NLO characteristics for DBTD6 and DBTD7. The highest μtot value is observed in DBTD6 as 10.362 D, whereas the highest αtot (1.48 × 10−22 and 1.47 × 10−22 esu) is observed in DBTD6 and DBTD7. Further, βtot (6.68 × 10−28 and 6.23 × 10−28 esu) in DBTD4 and DBTD5 and γtot (6.20 × 10−33 and 6.59 × 10−33 esu) values are observed in DBTD5 and DBTD6. To acquire favorable NLO responses in molecules, structural modelling utilizing efficient donor units played significant role. Thus, current research insights encourage researches to develop efficient NLO materials for optoelectronic applications.
Among these, understanding and finding new and better NLO materials to fulfil the needs of advanced technologies is the most important research field. Some years ago, several manufactured polymers, coloured compounds, and inorganic or organic diodes were semi-conductors.5 Although each kind of NLO material has its merits and demerits. Currently, organic based non-linear optical materials are more popular in the field of NLO research due to high non-linear optical parameters and the versatility of structures.6,7 Recent findings show that organic chemicals are valued for their excellent non-linear optical characteristics, attributed to their application in contemporary technologies, and can be synthesized cost-effectively.8 Therefore, currently much experimental and theoretical attention has been paid to π-conjugated molecular systems, based on donor moiety (D) linked through a π-bridge to an acceptor moiety (A) due to their exceptional NLO response.9,10 The design strategy of NLO in the D–π–A arrangement is characterized as a push–pull mechanism, which further enhances the internal charge transfer. Theoretical and experimental investigations confirm that careful selection of the π-spacer, donor, and acceptor units plays a crucial role in tuning NLO efficiency.11 The literature reports indicate that π-conjugated moieties provide an efficient charge transfer pathway under an electric field.12,13
Moreover, it is observed that incorporating π-conjugated linkers of an appropriate length enhances the NLO characteristics.7,14 By enhancing charge transfer between donor and acceptor units, non-centrosymmetric π-linker derivatives exhibiting first, second and third-order nonlinear polarizabilities can be achieved.15,16 To design the push–pull mechanism various schemes, for example, donor–acceptor, donor–π–acceptor, acceptor–π–donor–π–acceptor, and donor–donor–π–acceptor, are reported.17 Among them D–π–A framework, which facilitates photoinduced charge transfer, has been extensively explored in the design of high-performance NLO materials. At higher wavelengths, the push–pull mechanism can increase the asymmetric electronic distribution, improve the NLO response, lower the HOMO/LUMO band gap, and increase the light penetration range.18,19 The diphenylborane moiety enhances electron affinity and charge transfer process through its vacant p-orbital, lowering the LUMO level and boosting NLO and optoelectronic performance.20 The thiophene-based π-spacers enhance charge transport due to their high conjugation efficiency and rigidity, which aid in stabilizing the charge-separated states.21 Similarly, phenoxazine, triphenylamine, and carbazole units, with their strong electron-donating capabilities, enhance charge asymmetry and boost hyperpolarizability.22
Given these advantages, we designed a series of D–π–A derivatives, DBTD1–DBTD7, from a synthesized compound 4-(5-(diphenylboraneyl)thiophen-2-yl)-N,N-diphenylaniline23 utilized as reference compound for current study. This reference compound consists of a diphenylborane acceptor and a diphenylaniline donor, bridged by phenylthiophene π-spacers to promote efficient charge transfer. To further enhance intramolecular charge transfer (ICT), and electronic properties, the phenylthiophene unit in DBTR was replaced with tolyl thiophene and diphenyl thieno thiophene rings. Additionally, strategic donor modifications and central core adjustments were introduced to optimize π-conjugation and improve NLO efficiency.
This study explores the structure–property relationship of entitled compounds with distinct π-spacers and donor groups. A literature review revealed that their NLO properties remain unexplored. Our structural modifications significantly enhance the NLO response, making these compounds promising candidates for optoelectronic applications such as electro-optic switching, second-harmonic generation, and photonic devices. Furthermore, these findings provide valuable insights into the rational design of next-generation NLO materials, contributing to the development of more efficient and high-performance optoelectronic devices. The chemical structures and schematic representation of the reference and investigated compounds are presented in Fig. 1 and 2.
Dipole moment35 was calculated by using following eqn (1)
μ = (μx2 + μy2 + μz2)½ | (1) |
Average polarizability (α)36 was examined using the eqn (2)
(α) = 1/3(axx + ayy + azz) | (2) |
The magnitude of βtot37 is defined in eqn (3).
βtot = (βx2 + βy2 + βz2)½ | (3) |
Eqn (4) was utilized to get the second hyperpolarizability (γtot)38
![]() | (4) |
Koopman's theorem39 was used to calculate global reactivity descriptors using the following eqn (5)–(9):
![]() | (5) |
![]() | (6) |
![]() | (7) |
![]() | (8) |
![]() | (9) |
The visual depiction of entitled chromophores depicts electronic distribution patterns in HOMO/LUMO during intramolecular charge transfer using red and blue areas in Fig. S1.† In the same way, the pictographs of charge transferred in other orbitals (HOMO−1/LUMO+1 and HOMO−2/LUMO+2) are represented in Table S10 and Fig. S1.†
The energy findings of FMOs as illustrated in Table S9† represented that EHOMO of reference compound and designed derivatives are calculated as −5.631, −5.711, −5.576, −5.662, −5.553, −5.583, −5.786 and −5.915 eV while ELUMO values are found as −2.106, −2.613, −2.663, −2.667, −2.671, −2.690, −2.688 and −2.706 eV. We calculated the band gaps of the studied compounds using the HOMO–LUMO energies. The band gaps of DBTR–DBTD7 are computed as follows: 3.525, 3.098, 2.913, 2.995, 2.882, 2.893, 3.098 and 3.209 eV. Here, we observed that all the designed compounds showed smaller band gaps than the reference compound. The insertion of cyano units in derivatives, replacement of thiophene with thieno[3,2-b]thiophene, and enhanced resonance effect cause this reduction and enhance charge transfer. DBTD4 has the smallest band gap because of the introduction of a donor (N-methyl-N-phenylaniline) with the highest electron-donating effect. The data of FMOs along with molecular orbital energies also tells us about internal charge transfer between orbitals.53 The compound DBTD4 with lowest band gap has efficient charge transfer from donor to acceptor through π-bridge. The decreasing trend of band gaps for tailored compounds and reference are as follows: DBTR (3.525) > DBTD7 (3.209) > DBTD6 > DBTD1 (3.098) > DBTD3 (2.995) > DBTD2 (2.913) > DBTD5 (2.893) > DBTD4 (2.882) in eV. DBTD6, DBTD1, and DBTD7 had the largest band gaps among all the derivatives. This might be due to a varied chemical framework and various associated donors with a lower electron-donating impact and hence does not cause effective charge transfer.
The pictograms of FMO indicate the contribution of charge density. In DBTR, the electronic clouds related to the HOMO are mainly present on the donor and π-spacer. In contrast, the acceptor primarily contributes to the LUMO and a minor contribution from the π-spacer. However, a diverse charge density distribution was observed for the tailored compounds (DBTD1–DBTD7). The HOMOs of all the developed compounds demonstrated that their electronic clouds were primarily centred on the donor portion, with a modest contribution to the π-spacer.
On the other hand, the LUMO has focused electron clouds toward the acceptor part while leaving a tiny area on the π-spacer portion. This suggests that π-spacers facilitate the charge transfer between the donor and acceptor units. All investigated compounds may be considered significant NLO components owing to this charge flow. The pictorial representation of band gaps is presented in Fig. 4.
To investigate the DOS, we further divided the reference compound and its derivatives into three segments: donor, π-spacer, and acceptor parts, represented by red, green, and blue lines, respectively. The density of states (DOS) examination at the M06 functional was conducted using the 6-311G(d,p) basis set to elucidate the findings of DBTR and DBTD1–DBTD7. The alteration of donors and π-spacers resulted in a change in the electronic distribution pattern. This transition can be explained very well by studying the percentages of moieties in HOMOs and LUMOs.60 Regarding DBTR, the findings indicated that the donor moiety showed contributions of 57.4% towards HOMO and 2.4% towards LUMO. Similarly, the acceptor moiety displayed substantial contributions of 53.8% towards the LUMO and 2.3% towards the HOMO. The π-spacer contributed to the LUMO by 43.7% and 40.3% towards the HOMO, respectively. In same way, in the case of DBTD1–DBTD7, the notable electronic charge distributions for the LUMO were 71.0, 74.1, 73.6, 73.4, 72.4, 72.8, 72.9% for the acceptor moiety. The electron density in the HOMO is distributed over the acceptor is noted as 2.5, 0.8, 1.1, 0.7, 0.3, 1.0, and 1.2%. Similarly, the charge contributions on the donors for the LUMO are as follows: 1.4, 0.3, 0.1, 0.2, 0.1, 0.0, 0.1%, and 27.7, 25.5, 26.3, 26.4, 27.6, 27.1, and 27.0% on π-spacer, respectively. Whereas in HOMO, the charge contributions on donor are 58.1, 49.5, 25.2, 42.9, 50.1, 17.8, 35.4% and 39.4, 18.7, 73.7, 56.4, 49.5, 81.2, 63.5% on π-spacer, accordingly. The acceptor significantly contributes to the LUMO. In the HOMO, all compounds displayed a significant contribution towards the donor, except for DBTD3, DBTD4, DBTD6, and DBTD7. However, these compounds have a major contribution to the π-spacers. The recorded images for studied compounds are presented in Fig. 5, and their percentages are presented in Table S11.†
Compounds | IP | EA | X | η | μ | ω | σ | ΔNmax |
---|---|---|---|---|---|---|---|---|
a Ionization potential (I), electron affinity (A), electronegativity (X), global hardness (η), chemical potential (μ), global electrophilicity (ω), global softness (σ), and maximum charge transfer index (ΔNmax). | ||||||||
DBTR | 5.631 | 2.106 | 3.868 | 1.762 | −3.868 | 4.246 | 0.284 | 2.195 |
DBTD1 | 5.711 | 2.613 | 4.162 | 1.549 | −4.162 | 5.592 | 0.323 | 2.687 |
DBTD2 | 5.576 | 2.663 | 4.119 | 1.456 | −4.119 | 5.826 | 0.343 | 2.828 |
DBTD3 | 5.662 | 2.667 | 4.164 | 1.497 | −4.164 | 5.791 | 0.334 | 2.781 |
DBTD4 | 5.553 | 2.671 | 4.112 | 1.441 | −4.112 | 5.867 | 0.347 | 2.854 |
DBTD5 | 5.583 | 2.690 | 4.136 | 1.446 | −4.136 | 5.915 | 0.346 | 2.859 |
DBTD6 | 5.786 | 2.688 | 4.237 | 1.549 | −4.237 | 5.795 | 0.323 | 2.736 |
DBTD7 | 5.915 | 2.706 | 4.310 | 1.604 | −4.310 | 5.790 | 0.312 | 2.686 |
The association between the hardness (η), energy gaps, stability of a molecule, and chemical potential (μ) is direct. However, the relationship between reactivity and softness is inverse.45
Based on the GRPs, we observed that the electron affinity values were higher in derivatives than the reference chromophore. The hardness values of DBTR–DBTD7 are 1.762, 1.549, 1.456, 1.497, 1.441, 1.446, 1.549, and 1.604 eV. The softness values of the tailored compounds were greater than those of the reference compound: 0.284, 0.323, 0.343, 0.334, 0.347, 0.346, 0.323, and 0.312 eV−1. Another parameter used to examine the stability and reactivity of the substances under study is the chemical potential (μ). It varies directly with stability and inversely with the compound reactivity. A D–π–A system is predicted to be more reactive and less stable because its μ values are more negative.68 The chemical potential values are presented in Table 2. Moreover, the decreasing trend in hardness was DBTR > DBTD7 > DBTD6 = DBTD1 > DBTD3 > DBTD2 > DBTD5 > DBTD4. The order of increasing softness was DBTR < DBTD7 < DBTD6 = DBTD1< DBTD3 < DBTD2 < DBTD5 < DBTD4. The values of the global electrophilicity index of DBTR–DBTD7 were 4.246, 5.592, 5.826, 5.791, 5.867, 5.915, 5.795, and 5.790 eV, respectively. Since the compound (DBTD4) has the lowest hardness (1.441 eV) and the highest softness (0.347 eV−1), which matches its lower energy band gap (2.882 eV), it is expected to be the most advantageous designed compound with good NLO properties.
Comp. | DFT λ (nm) | E (eV) | fos | MO contributions |
---|---|---|---|---|
a MO = molecular orbital, H = HOMO, L = LUMO, fos = oscillator strength, DFT = density functional theory. | ||||
DBTR | 430.097 | 2.883 | 0.921 | H → L (95%) H−1 → L (2%) H → L+1 (2%) |
DBTD1 | 485.242 | 2.555 | 0.758 | H → L (96%) |
DBTD2 | 502.265 | 2.469 | 0.741 | H → L (93%) H−1 → L (3%) H → L+1 (3%) |
DBTD3 | 490.657 | 2.527 | 0.984 | H → L (93%) H−1 → L (3%) H → L+1 (3%) |
DBTD4 | 502.591 | 2.467 | 0.885 | H → L (92%) H−1 → L (4%) H → L+1 (3%) |
DBTD5 | 488.088 | 2.540 | 1.024 | H−1 → L (14%) H → L (81%), H → L+1 (3%) |
DBTD6 | 473.276 | 2.620 | 1.248 | H → L (87%) H−1 → L (8%) H → L+1 (3%) |
DBTD7 | 459.116 | 2.701 | 1.163 | H → L (86%) H−1 → L (9%) H → L+1 (3%) |
The values of λmax, along with the associated aspects, such as transition energy, oscillation strength, and molecular contributions in the solvent and gaseous phases, are presented in Tables 2 and 3. In contrast, Tables S12–S27† indicate the lowest six transitions. The data show that all of the compounds under investigation have absorption levels in the visible region of the electromagnetic spectrum, except DBTD2, DBTD3, and DBTD4 in tetrahydrofuran solvent. They displayed low absorption values in the ultraviolet region, which might be due to the presence of double thiophene, which can lead to increased steric hindrance, conjugational distortion, and complex geometry alterations. When the oscillation strength was preferred over the wavelength, the absorption maxima (λmax) in THF solvent were as follows: DBTD5 (488.088) > DBTD1 (485.242) > DBTD6 (473.276) > DBTD7 (459.116) > DBTR (430.097) > DBTD4 (394.013) > DBTD2 (387.693), and DBTD3 (377.333) (Tables S12–S19†). Here, DBTD5 has the highest wavelength, which might be due to the presence of auxochrome, leading to a redshift with lowest excitation energy of 2.540 eV.
Comp. | DFT λ (nm) | E (eV) | fos | MO contributions |
---|---|---|---|---|
a MO = molecular orbital, H = HOMO, L = LUMO, fos = oscillator strength, DFT = density functional theory. | ||||
DBTR | 413.391 | 2.999 | 0.836 | H → L (95%) |
DBTD1 | 472.555 | 2.624 | 0.667 | H → L (97%) |
DBTD2 | 506.203 | 2.449 | 0.600 | H → L (96%) |
DBTD3 | 496.056 | 2.499 | 0.810 | H → L (96%) H−1 → L (2%) |
DBTD4 | 509.468 | 2.434 | 0.710 | H → L (95%) H−1 → L (3%) |
DBTD5 | 502.754 | 2.466 | 0.667 | H → L (90%) H−1 → L (7%) |
DBTD6 | 481.380 | 2.576 | 0.988 | H → L (90%) H−1 → L (6%) |
DBTD7 | 463.579 | 2.675 | 0.867 | H−1 → L (10%) H → L (86%) H → L+1 (2%) |
However, there are variations in the wavelength values if we favor wavelength over oscillation strength in the THF solvent. In this case, the absorption maxima in THF decreased in the following order in nm: DBTD4 (502.591) > DBTD2 (502.265) > DBTD3 (490.657) > DBTD5 (488.088) > DBTD1 (485.242) > DBTD6 (473.276) > DBTD7 (473.276), and DBTR (459.116).
At the lowest excitation energy of 2.467 eV, the maximum absorption value was recorded for DBTD4 at 502.591 nm. Because of its narrowest band gap, it can take less energy to transition, and more electrons might move into excited states, accounting for its highest wavelength.
All derivatives exhibited higher absorption values in the gaseous phase than the reference compound. The compounds exhibited absorption values in the range from 413.391 nm to 509.468 nm. The DBTD4 had the highest wavelength with an excitation energy of 2.434 eV in the gas phase. This might be due to the development of a more efficient push–pull mechanism. These findings were obtained when we prioritized wavelengths above oscillation strength. The gaseous phase exhibited a declining trend in λmax values measured in nm, as follows: DBTD4 (509.468) > DBTD2 (506.203) > DBTD5 (502.754) > DBTD3 (496.056) > DBTD6 (481.380) > DBTD1 (472.555) > DBTD7 (463.579), and DBTR (413.391). When we prefer oscillation frequency to wavelength, the proposed compounds exhibit absorption values ranging from 370 to 496 nm regarding oscillation strength preference. All the substances exhibited UV absorption values except for DBTD3, DBTD7,DBTD1, and DBTR. (Tables S20–S36†).
In summary, DBTD4 had the strongest affinity for decreasing the transition energy, the smallest energy difference, and exhibited a bathochromic shift. These features make it a desirable optically active material for nonlinear optics applications.
Light-harvesting efficiency (LHE) is a crucial factor that determines the optical performance of the designed compounds. It can be evaluated using the oscillator strength (fos) obtained from TD-DFT calculations. A higher LHE value indicates an improved photocurrent response of the compound. The LHE for the designed chromophores was calculated using eqn (10) (ref. 74) and the results are presented in Table S36.†
LHE = 1 − 10−f | (10) |
Oudar and Chemla introduced a two-state model35 that has been widely used in the literature to analyze the NLO response, incorporating the key contributions of the ground and excited states in a sum-over-states approach. This model establishes a relationship between charge transfer transitions and second-order polarizability, which serves as the foundation for the push–pull architecture in designing high-performance NLO compounds.
βCT = (Δμgm × fgm)/Egm3 | (11) |
Eqn (11) represents the two-state model for calculating first hyperpolarizability (βCT). In this model, βCT depends on three key factors: the dipole moment difference between the ground and excited states (Δμgm), the oscillator strength of the transition (fgm), and the cube of the transition energy (Egm3). According to this model, βCT is directly proportional to Δμgm and fgm, meaning that a larger dipole moment difference and higher oscillator strength enhance the NLO response. Conversely, βCT is inversely proportional to Egm3, indicating that a lower transition energy leads to better NLO performance. Therefore, molecules with strong NLO properties should have a significant dipole moment difference, strong oscillator strength, and lower transition energy. In this study, these parameters for entitled compounds were determined and are presented in Table S36.†
The results indicate that among the studied compounds, reference compound exhibits the highest Egm3 value at 3.862 eV, while the lowest value is observed for DBTD4 at 2.435 eV. This suggests that DBTD4 compound might exhibit a stronger NLO response due to its lower transition energy. Additionally, to further analyze the NLO properties, the relationship between the two-state model (βCT) and the total hyperpolarizability (βtot) for all compounds (DBTR, DBTD1–DBTD7) is illustrated in Fig. 6. The variations in the graph reflect the influence of different π-spacers and terminal acceptor units on βtot, which align with the trends predicted by the two-state model (βCT), supporting their potential as efficient NLO materials.
![]() | (12) |
Different transitions are observed in the natural bond orbital analysis, such as π → π*, σ → σ*, LP → π*, and LP → σ*. The π–π * transition is the most significant of all the transitions. Charge transfer in the π → π* transition is more dominant than that in σ → σ*. The stabilization energy is the most important parameter to consider in NBO analysis. Table 4 illustrates the selected NBO values, whereas the other values are presented in Tables S37–S44.†
Compounds | Donor(i) | Type | Acceptor(j) | Type | E(2) [kcal mol−1] | E(j) − E(i) [a.u] | F(i,j) [a.u] |
---|---|---|---|---|---|---|---|
DBTR | C2–C4 | π | C7–C9 | π* | 24.54 | 0.3 | 0.077 |
C35–C37 | π | C32–C33 | π* | 15.88 | 0.3 | 0.064 | |
C37–B39 | σ | C35–C37 | σ* | 7.21 | 1.17 | 0.082 | |
S34–C37 | σ | C33–C35 | σ* | 0.52 | 1.22 | 0.023 | |
S34 | LP (2) | C32–C33 | π* | 27.12 | 0.27 | 0.078 | |
N1 | LP (1) | C22–C23 | σ* | 3.55 | 0.85 | 0.052 | |
DBTD1 | C50–C52 | π | C55–C57 | π* | 26.37 | 0.28 | 0.077 |
C32–C33 | π | C32–C33 | π* | 0.55 | 0.3 | 0.012 | |
C57–C64 | σ | C64–N65 | σ* | 8.85 | 1.61 | 0.107 | |
C37–B39 | σ | S34–C37 | σ* | 0.52 | 0.8 | 0.018 | |
S34 | LP (2) | C32–C33 | π* | 27.7 | 0.27 | 0.079 | |
N63 | LP (1) | C47–C62 | σ* | 11.87 | 1.06 | 0.1 | |
DBTD2 | C18–C20 | π | C23–C25 | π* | 26.26 | 0.3 | 0.079 |
C23–C25 | π | C23–C25 | π* | 0.64 | 0.29 | 0.012 | |
C66–C72 | σ | C72–N73 | σ* | 8.86 | 1.61 | 0.107 | |
B49–C50 | σ | C25–B49 | σ* | 0.57 | 1.08 | 0.022 | |
N28 | LP (1) | C13–C15 | π* | 25.54 | 0.29 | 0.08 | |
N73 | LP (1) | C66–C72 | σ* | 11.87 | 1.06 | 0.1 | |
DBTD3 | C11–C12 | π | C14–C17 | π* | 26.02 | 0.28 | 0.076 |
C1–C3 | π | C1–C3 | π* | 0.64 | 0.29 | 0.012 | |
C40–N41 | σ | C26–C40 | σ* | 8.2 | 1.58 | 0.102 | |
C42–S44 | σ | C42–C43 | σ* | 0.52 | 1.25 | 0.023 | |
N74 | LP (1) | C50–C52 | π* | 32.39 | 0.32 | 0.094 | |
N39 | LP (1) | C17–C38 | σ* | 11.87 | 1.06 | 0.1 | |
DBTD4 | C43–C45 | π | C48–C50 | π* | 26.1 | 0.3 | 0.079 |
C48–C50 | π | C48–C50 | π* | 0.64 | 0.29 | 0.012 | |
C71–C76 | σ | C76–N77 | σ* | 8.86 | 1.61 | 0.107 | |
C32–S34 | σ | C32–C33 | σ* | 0.52 | 1.25 | 0.023 | |
N1 | LP (1) | C2–C4 | π* | 28.23 | 0.29 | 0.084 | |
N75 | LP (1) | C61–C74 | σ* | 11.87 | 1.06 | 0.1 | |
DBTD5 | C43–C45 | π | C48–C50 | π* | 25.9 | 0.3 | 0.079 |
C35–C37 | π | C40–C41 | π* | 17.08 | 0.32 | 0.067 | |
C61–C74 | σ | C74–N75 | σ* | 8.86 | 1.61 | 0.107 | |
C81–H86 | σ | C81–C82 | σ* | 0.96 | 1.12 | 0.029 | |
S42 | LP (2) | C35–C37 | π* | 25.38 | 0.26 | 0.075 | |
N75 | LP (1) | C61–C74 | σ* | 11.87 | 1.06 | 0.1 | |
DBTD6 | C41–C43 | π | C46–C48 | π* | 25.88 | 0.3 | 0.079 |
C21–C22 | π | C21–C22 | π* | 0.69 | 0.29 | 0.013 | |
C74–N75 | σ | C69–C74 | σ* | 8.21 | 1.58 | 0.102 | |
C30–S32 | σ | C30–C31 | σ* | 0.54 | 1.25 | 0.023 | |
N1 | LP (1) | C2–C3 | π* | 36.4 | 0.31 | 0.097 | |
N73 | LP (1) | C59–C72 | σ* | 11.87 | 1.06 | 0.1 | |
DBTD7 | C39–C41 | π | C44–C46 | π* | 25.71 | 0.3 | 0.079 |
C44–C46 | π | C44–C46 | π* | 0.57 | 0.29 | 0.011 | |
C72–N73 | σ | C67–C72 | σ* | 8.21 | 1.58 | 0.102 | |
C2–C4 | σ | N1–C10 | σ* | 0.5 | 1.16 | 0.022 | |
N1 | LP (1) | C19–C20 | π* | 40.32 | 0.31 | 0.101 | |
N71 | LP (1) | C57–C70 | σ* | 11.87 | 1.06 | 0.1 |
In the case of reference compound (DBTR), the value of the highest stabilization energy for π (C2–C4) → π* (C2–C4) transition is 24.4 kcal mol−1 while the lowest energy for π (C35–C37) → π* (C32–C33) is about 15.88 kcal mol−1. The weak transitions σ → σ* show greatest and lowest stabilization energy as 7.21 and 0.52 for σ (C37–B39) → σ* (S35–C37) and σ (S34–C37) → σ* (C33–C35). The lone pair transitions show values as 27.12 and 3.55 kcal mol−1 for LP (S34) → π* (C32–C33) and LP (N1) → σ* (C22–B23).
The stabilization energy values for DBTD1 are 26.37, 0.55, 8.85, 0.52, 27.7, and 11.87 kcal mol−1 with corresponding transition as π (C50–C52) → π* (C55–C57), π (C32–C33) → π*(C32–C33), σ (C57–C64) → σ* (C64–N65), σ (C37–B39) → σ* (S34–C37), LP (S34) → π* (C32–C33), LP (N63) → σ*(C47–C62). We observed that DBTD1 had a higher stabilization energy than the reference compound.
The designed compound (DBTD2) has a maximum stabilization energy of 26.26 kcal mol−1 for the π (C18–C20) → π* (C23–C25) transition and a minimum of 0.64 kcal mol−1 for the π (C23–C25) → π* (C23–C25) transition. The weak σ → σ* transitions exhibit the highest and lowest stabilization energies of 8.86 and 0.57, respectively, for σ (C66–C72) → σ* (C72–N73) and σ (B49–C50) → σ* (C25–B49). The lone pair transitions for LP (N28) → π* (C13–C15) and LP(N73) → σ* (C66–C72) gave values of 25.54 and 11.87, respectively.
For the π (C11–C12) → π* (C14–C17) transition, the proposed compound (DBTD3) had a maximum stabilization energy of 26.02 kcal mol−1, whereas for the π (C1–C3) → π* (C1–C3) transition, it had a minimum stabilization energy of 0.64 kcal mol−1. The stabilization energies of the weak transitions σ → σ* are the lowest at 0.52 and the highest at 8.2 for σ (C42–S44) → σ* (C42–C43) and σ (C40–N41) → σ* (C26–C40). For LP (N74) → π* (C50–C52) and LP(N39) → σ* (C17–C38), the lone pair transitions provided values of 32.39 and 2.34, respectively.
Regarding DBTD4, there is a minimum stabilization energy of 0.64 kcal mol−1 for the π (C48–C50) → π* (C48–C50) transition and a maximum stabilization energy of 26.1 kcal mol−1 for the π (C43–C45) → π* (C48–C50) transition. When discussing weak transitions σ (C71–C76) → σ* (C76–N77) and σ (C32–S34) → σ* (C32–C33), the highest and lowest stabilization energies of the transitions σ → σ* are 8.86 and 0.52 kcal mol−1. Analysing further transitions for LP (N1) → π* (C2–C4) and LP (S42) → σ* (C33–C35), the values are 28.23 and 11.87 kcal mol−1.
With respect to DBTD5, the π (C35–C37) → π* (C40–C41) transition had a minimum stabilization energy of 17.08 kcal mol−1, whereas the π (C43–C45) → π* (C48–C50) transition had a maximum stabilization energy of 25.9 kcal mol−1. The weak transitions σ (C61–C74) → σ* (C74–N75) and σ (C81–H86) →σ* (C81–C82) have the greatest and lowest stabilization energy, respectively, at 8.86 and 0.96 kcal mol−1. The results of the analysis of lone pair transitions for LP (S42) → π* (C35–C37) and LP (N75) → σ* (C61–C74) are 25.38 and 11.87 kcal mol−1, respectively.
A comparison of the stabilization energies for all the electronic transitions occurring in the DBTD6 molecule shows that the energy transition π (C41–C43) →π* (C46–C48) has the maximum energy stabilized at 25.88 kcal mol−1. The transition π (C21–C22) →π* (C21–C22) requires a small energy of only 0.69 kcal mol−1. Quantitatively, we found that the maximum energy of σ → σ * transition is 8.21 kcal mol−1 for (C74–N75) and (C69–C74). Conversely, the minimum energy in the case of σ (C30–S32) → σ* (C30–C31) was 0.54 kcal mol−1. It was noted that LP → π* held the maximum value of 36.4 kcal mol−1 in LP (N1) and (C2–C3) transition, while a minimum value of 11.87 kcal mol−1 was found in LP (N73) → σ * (C59–C72) transition.
Analysis of the DBTD7 molecule showed that the π (C39–C41) → π* (C44–C46) transitions have the largest stabilization energy of 25.71 kcal mol−1, while the π (C44–C46) → π* (C44–C46) transitions have the lowest value, 0.57 kcal mol−1. The weak interaction transitions σ (C72–N73) →σ* (C67–C72) and σ (C2–C4) → σ*(N1–C10) have the highest and lowest stabilization energies of 8.21 and 0.5 kcal mol−1, respectively. Furthermore, the LP (N71) → σ* (C57–C70) and LP (N1) → π* (C19–C20) transitions, which have stabilization energies of 11.87 and 40.32 kcal mol−1, respectively, are caused by resonance effects in the molecule. For DBTD7, many other transitions are shown in Table S37.†
The above discussion indicates that the NBO findings of all the chemicals examined are interrelated. However, it should be noted that, unlike other designed compounds, DBTD1 gave the highest stabilization energies. This could be due to the presence of cyano groups, as they often enhance electron delocalization, leading to molecular stability. The stability of the compound is also evidenced by the charges of the NBOs. We can observe that ICT and hyperconjugation play a vital role in developing materials with strong NLO characteristics.
An essential component for determining the polarizability of organic compounds is the dipole moment (μ).91 For entitled molecules, their μ values are calculated in Debye (D), in solvent and gas phase shown in Tables S45 and S52,† respectively. As observed here, all developed compounds exhibited higher values than the reference compound. The highest μtot was found in DBTD6 and DBTD3 (10.362 and 10.223D, respectively) in solvent phase because of the strong donor moieties, which may be due to their non-centrosymmetric structure. Urea was used as a reference compound for comparison analysis in NLO studies, according to the literature data its dipole moment is 1.3732D.16 The μtot in Debye for all other compounds and references in tetrahydrofuran are as the following: DBTD1 (9.920D), DBTD4 (9.610), DBTD2 (9.090), DBTD5 (8.828), DBTD7 (5.402), DBTR (2.328). All the designed derivatives had higher values than urea. Higher values of μtot indicate high polarizability. For all the studied compounds, larger polarizability values were analyzed along the x-axis (DBTD1 = −9.573, DBTD2 = 8.874, DBTD3 = 10.167, DBTD4 = 9.610, DBTD5 = −8.827, DBTD6 = −10.258, and DBTD7 = −5.401 in D).
For DBTR and DBTD1–DBTD7, Table S46† displays the 〈α〉 values in solvent phase. It also reveals crucial 3-D tensors for polarizability. The linear polarizability value of DBTR was the lowest, and all other designed compounds had the highest values. Along the z-axis all compounds showed major contributing tensor in the (αtot) value compared with the x- and y-axes with comparable values. The 〈α〉 values for entitled molecules in declining order are represented as follows in solvent phase: DBTD5 (1.48 × 10−22) > DBTD6 (1.47 × 10−22) > DBTD4 (1.31 × 10−22) > DBTD7 (1.29 × 10−22) = DBTD3 (1.29 × 10−22) > DBTD2 (1.19 × 10−22) > DBTD1 (1.02 × 10−22) and DBTR (0.916 × 10−22) in esu. The computed linear polarizability in gas phase is presented in Table S53.†
According to the literature, the FMO bandgap has a distinct impact on the polarizability of compounds; that is, the narrower the bandgap, the higher the polarizability values, and vice versa.50 The compounds to be studied in our case have a similar pattern: DBTD4 has the highest βtot value (6.68 × 10−28 esu) with the smallest band gap of 2.882 eV due to the extended conjugated system and strong push–pull framework. A decreasing trend in all entitled compounds in solvent phase is as follows: DBTD4 (6.68 × 10−28) > DBTD3 (6.23 × 10−28) > DBTD5 (6.04 × 10−28) > DBTD2 (5.33 × 10−28) = DBTD6 (5.33 × 10−28) > DBTD1 (4.23 × 10−28) > DBTD7 (3.73 × 10−28) and DBTR (2.17 × 10−28) Table S47.† The calculated βtot in gas phase of all examined compounds is presented in Table S55.†
In the second hyper-polarizability, the values of γtot are the highest for the DBTD5 and DBTD4 compounds for solvent phase, which have the lowest band gaps. All values of γtot were comparable to each other and greater than that of the reference compound (Table S48†). The order of declining trend of γtot in solvent medium in esu is following: DBTD5 (6.59 × 10−33) > DBTD4 (6.20 × 10−33) > DBTD6 (5.32 × 10−33) > DBTD3 (5.07 × 10−33) > DBTD7 (3.54 × 10−33) > DBTD2 (4.42 × 10−33) > DBTD1 (2.75 × 10−33) and DBTR (1.43 × 10−33). The second hyper-polarizability of examined compounds in gas phase is presented in Table S56.†
We computed frequency-dependent first hyperpolarizability coefficients for electro-optic Pockel's effect (EOPE) with β(ω;ω,0) and second-harmonic generation of first hyperpolarizability (SHG) with β(−2ω;ω,ω) at commonly used laser wavelengths of 532 nm and 1064 nm to give insights for experimental studies. At 532 nm, the EOPE values fall between 2.84 × 10−28 to 10.1 × 10−28, whereas at 1064 nm, the response is in the range of 20.5 × 10−28 to 773 × 10−28 in solvent phase. We found significant EOPE values at 1064 nm, which might be attributed to an enhancement of resonance. Similarly, the SHG values are high for all derivatives except DBTD5 and DBTD6 at both 532 and 1064 as shown in the Table S49.† Therefore, it can be concluded that increased wavelength produces significant SHG responses.
At 532 nm and 1064 nm, third order nonlinear optical response coefficients, such as the electric field induced second harmonic generation (ESHG) γ(−2ω;ω,ω,0) and the dc-Kerr effect γ(−ω;ω,0,0) were also calculated. Higher dc-Kerr effect and ESHG values are mostly seen at 1064 nm, showing that the entitled compounds can significantly improve their performance at higher wavelengths as depicted in Table S50† might be due to increased polarizability.
Overall, it has been noted that all derivatives have polarizable characteristics and a smaller band gap than DBTR. Among all derivatives, DBTD4, DBTD5, and DBTD6 had the highest NLO values. Furthermore, these compounds have low hardness and high softness values, indicating that these are more polarizable. Therefore, the entitled compounds are considered as good NLO materials for optoelectronic devices. The frequency dependent second order and third order optical response in gas phase is illustrated in Tables S56–S57.†
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra08662g |
This journal is © The Royal Society of Chemistry 2025 |