Eman Nabil* and
Mohamed Zakaria
Department of Chemistry, Faculty of Science, Alexandria University, Alexandria, Egypt. E-mail: eman.nabil@alexu.edu.eg
First published on 7th August 2025
Conjugation elongation of squaraines is a potential approach to optimize their performance as nonlinear optical (NLO) chromophores and TiO2-photosensitizers in dye-sensitized solar cells (DSCs). This study investigates the impact of integrating π-conjugated heteroaromatic spacers on the optoelectronic properties of four modeled unsymmetrical squaraine derivatives. Density functional theory (DFT) and time-dependent TD-DFT computations revealed the dual functionality of all four π-extended squaraine dyes, with the capability to sensitize TiO2 for far-red light harvesting and amplify second-order NLO response at the molecular-level. Dye SQ-N incorporating an ethyl-dithienopyrrole π-spacer emerged as the optimal photosensitizer for TiO2-based DSCs, exhibiting a hyperchromic S1 transition and a light harvesting efficiency (LHE) of 98% at 697 nm (λmax), the most thermodynamically driven electron injection (ΔGinj), robust adsorption (Eads) onto the TiO2 nanocluster, enhanced orbital coupling (ΔEoi) and hybridization between virtual molecular π* orbitals of SQ-N and 3d-orbitals of Ti atoms, in addition to superior charge transfer at the SQ-N–TiO2 interface, under deep red-to-NIR photoexcitation. Conversely, the P-acetyl dithienophosphole oxide π-linker in SQ-P led to a sixfold enhancement in off-resonant hyperpolarizability (β0) compared to the π-spacer-free parent dye, in addition to manifesting the maximal dynamic electro-optic Pockels (EOP) β1064 and β1460, second-harmonic generation (SHG) β1064 and hyper-Rayleigh scattering β1064. Analytical DFT-predicted SHG activity of the modeled dyes showed simultaneous potential for NIR-to-green and telecom E-band (1460 nm) to red light conversion. β scans demonstrated dual EOP and optical rectification functionality, while dyes SQ-N and SQ-Th further displayed significant sum/difference frequency generation (SFG/DFG) output at ω1 ± ω2. Polarization-resolved SHG analysis revealed a hybrid dipolar–octupolar NLO symmetry across all the dyes, with maximal harmonic intensity at Ψ = ±90°—a signature of synergistic dipole alignment and 3D charge delocalization.
Design, System, ApplicationSquaraine dyes, with their intense far-red absorption and tunable electronic properties, are propitious candidates for dye-sensitized solar cells (DSCs) and nonlinear optical (NLO) applications. However, their limited conjugation often restricts their charge transfer efficiency and light-harvesting range. In this work, we designed π-extended unsymmetrical benzoindole-based squaraine dyes by incorporating four π-conjugated spacers: P-acetyl dithienophosphole oxide (SQ-P), ethyl-dithienopyrrole (SQ-N), dimethyl-silolo-dithiophene (SQ-Si) and thienothiophene (SQ-Th). As computationally assessed for the free dyes and those adsorbed on a TiO2 (38-atom) nanocluster model, this molecular engineering enhanced the intramolecular charge transfer (ICT) and redshifted absorption toward the far-red/NIR domain, improving TiO2 sensitization. Additionally, their strong ICT character, EW-driven electron displacement and tunable dipole moments make them ideal for NLO applications, with computed first hyperpolarizabilities surpassing conventional squaraine chromophores. |
Dye-sensitized solar cells (DSCs), a third-generation photovoltaic technology, are distinguished by their cost-effectiveness and ambient sunlight harvesting, besides aesthetic appeal, making them applicable to portable electronics and indoor building integration.8 The three main components of a DSC device are: a dye-sensitized mesoporous semiconducting photoanode, typically TiO2; a Pt counter electrode; and an electrolyte packed in between the two electrodes. The photosensitizer anchored to the TiO2 nanoparticles absorbs photons and injects the photoexcited electrons into the conduction band (CB) of TiO2. A critical barrier for maximizing photon harvesting efficacy is the inherent limitation of single-dye-sensitized TiO2 in attaining panchromatic solar absorption. This limitation can be strategically mitigated by extending squaraine conjugation with π-spacers, which synergistically broaden spectral response into NIR and enhance electron injection dynamics. For instance, π-extension with each –CC– unit has been reported to exhibit ∼100 nm of redshift in the charge transfer transition for a series of unsymmetrical squaraine dyes, while maintaining high molar extinction coefficients.9 Similarly, thiophene-based spacers have been shown to contribute to the panchromatic response of unsymmetrical squaraine dye, with an onset of 850 nm.10 The role of π-spacers extends beyond mere conjugation lengthening; boosted molecular conjugation improves the electronic coupling between the dye's LUMO and CB of TiO2, thus accelerating electron injection kinetics.11 The π-spacers, incorporated with alkyl chains, can also suppress the charge recombination rate, thereby enabling smoother electron flow through the semiconductor material.6 Moreover, a π-spacer can introduce steric effects that prevent random dye aggregation on TiO2 surfaces—a common issue that reduces photocurrent generation due to exciton quenching, further enhancing the DSC device's long-term stability.12
In the context of NLO, extended molecular conjugation increases the π-electron density participating in ICT, thereby elevating the nonlinear susceptibility.13–15 NLO phenomena encompass a range of physical processes, including harmonic generation, electro-optic modulation, parametric frequency conversion (sum/difference frequency generation), and optical rectification. One fundamental second-order NLO phenomenon is the second-harmonic generation (SHG), a frequency-doubling effect (and consequently a wavelength-halved), which is utilized in laser technology to tune the colour of their outcoming light beam. Molecular-scale control of SHG has been previously established by several favourable structural axioms, including intramolecular polarizability, planar molecular geometry, a ‘push–pull’ system, and an effective ICT facilitated by an extended π-conjugated bridge.16 The NLO activity of squaraine chromophores with push–pull architecture has garnered sustained scholarly interest. Their high polarizability, amplified under high-intensity laser irradiation, renders them promising candidates for numerous advanced photonic technologies. In 1994, Chen et al. reported that extended π-conjugation of an unsymmetrical squaraine molecule resulted in an absorption maximum (λmax) at 732 nm, a red-shift of over 160 nm, and a five-fold increase in the static first hyperpolarizability (β0), a critical metric for NLO performance.17 Beverina et al. have synthesized π-extended symmetric squaraine derivatives as two-photon absorption (2PA) active photosensitizers, exhibiting a strong 2PA cross-section at 806 nm, attributed to π-rich heterocycles.18 Bondar et al. have presented new symmetrical squaraine molecules showing 2PA, with a maximum 2PA cross-section (δ2PA max) of approximately 400 GM at around 800 nm, in addition to efficient NIR-superluminescence and high photostability.19 Such multifunctional NLO performance is important for emerging technologies in telecommunications and biomedical imaging, where NIR compatibility and high nonlinearity are paramount.
This study presents a comprehensive quantum chemical investigation to elucidate the impact of π-extended conjugation on the performance of four modelled squaraine dyes as NLO-active chromophores and photosensitizers for TiO2-based DSCs. A parent squaraine dye, coded as SQ-140,20 featuring a benzoindole-squaraine-indole scaffold with a cyanoacrylic acid acceptor/anchoring group, and devoid of π-spacers, served as the reference system. The four modelled unsymmetrical squaraine derivatives (SQ-P, SQ-N, SQ-Si and SQ-Th) were computationally designed through structural tailoring of four π-conjugated spacers. The optoelectronic and NLO properties of the modelled dyes were rigorously evaluated at the molecular level via (time-dependent) density functional theory computations. Our dyes exhibited synergistic dipole alignment and 3D charge delocalization, i.e., a mixed dipolar–octupolar symmetry. Accordingly, this work establishes structure–property correlations demonstrating that squaraine dyes with mixed dipolar–octupolar symmetry enable simultaneous enhancement of NLO efficiency and TiO2 charge injection. The study recommends an ethyl-dithienopyrrole π-linker as optimal for TiO2-photosensitization and a P-acetyl dithienophosphole oxide π-linker for squaraine-based NLO chromophores, providing a rational molecular modeling blueprint for dual-functional organic materials in energy conversion and photonic technologies.
B3LYP is a hybrid functional that combines a limited amount of exact exchange (20% HFX) with a gradient-corrected correlation functional. The functional is known for its problematic over-polarization when a significant charge separation appears, which prompts errors in CT transition energies.50 Herein, the superior performance showed by B3LYP could be attributed to the nature of transition in SQ-140 reference dye, being reminiscent of typical cyanine topology with a partial CT character. Indeed, the simulated charge-transfer spectrum of SQ-140 reference dye, as obtained by the Multiwfn code, revealed that the electronic excitation at 668 nm is a hybridized transition with π → π* local excitation (64%) and charge transfer (36%) characters. Accordingly, the balanced treatment of B3LYP between exact and dynamic exchange or local and non-local interactions could deliver a reasonable accuracy. In the same way, the RSH functionals are designed to partition the exchange–correlation energy into short-range and long-range components, with increased exact HFX% in long ranges to correct the dynamic behaviour of the exchange potential. This is crucial for pure long-range CT transition with significant extent of charge separation. However, for a hybrid excitation, as in the current system, the range-separation amplifies the CT component disproportionately, and neglects the significant local excitation contribution, leading to the observed overestimated excitation energies and large blue shifts in absorption wavelengths.
Considering that the four investigated dyes in the present study are the π-spacer modified analogues to SQ-140 reference dye, particular attention was given to the performance of B3LYP and CAM-B3LYP in predicting the resulting changes in excitation energies. Since the CT character is anticipated to become more prominent upon π-spacer insertion, CAM-B3LYP was selected for comparison due to its relative efficacy among the examined RSH functionals (Fig. 1), given that RSHs are generally considered suitable for describing CT transitions. A bathochromic shift in absorption spectra of modelled dyes is also anticipated owing to introducing a π-spacer reduces the energy gap by boosting π-conjugation. However, CAM-B3LYP predicted a counterintuitive blueshift, for all four modelled dyes (SQ-P, SQ-N, SQ-Si and SQ-Th) with calculated λmax values of 595, 592, 594 and 589 nm, respectively. These values are physically improbable considering the extended π-systems compared to the parent SQ-140 dye (668 nm). Conversely, B3LYP maintained reasonable accuracy, with the corresponding wavelengths of 850, 695, 746 and 728 nm. This suggests that B3LYP benefits from its lower HF exchange and empirical parameterization in capturing the electronic interactions in the examined squaraine systems, thus effectively describing their red-shifted hybridized transitions, even when the CT character dominates. Indeed, a good agreement between the theoretical and experimental UV-vis spectra was reported for TD-B3LYP simulated structurally congeneric squaraine dyes.22 Lastly, the impact of augmenting the 6-311G(d,p) basis set with diffuse functions was evaluated for the best-performing B3LYP functional. As revealed in Fig. 1, upon utilizing the TD-B3LYP/6-311+G(d,p) level of theory, the absolute deviation attained is limited to ca. 0.016 eV (5.6 nm). It is worth mentioning that this result (0.016 eV) has attained the chemical accuracy, defined as an absolute error of less than 0.043 eV.51 Accordingly, this computational level can be deemed reliable for simulating both linear and nonlinear optical phenomena of the modelled squaraine dyes.
JSC = ∫LHE(λ)Φinj(λ)ηcollectdλ | (1) |
LHE(λ) = 1 – 10–f | (2) |
![]() | (3) |
![]() | (4) |
ΔGreg = E(I−/I3−) − Eox | (5) |
Ground state (S0) | Excited state (S1) | |||||||
---|---|---|---|---|---|---|---|---|
φ1 | φ2 | β1 | β2 | φ1 | φ2 | β1 | β2 | |
SQ-N | 2 | 25 | 1 | 1 | 3 | 0.3 | 1 | 1 |
SQ-P | 1 | 25 | 1 | 1 | 2 | 24 | 1 | 1 |
SQ-Si | 2 | 23 | 2 | 2 | 3 | 4 | 0.6 | 0.4 |
SQ-Th | 0.3 | 25 | 1 | 1 | 0.3 | 12 | 1 | 1 |
DFT calculations showed that conjugation is exhibited via four of the π-conjugated spacers under study with vanishing bond length alteration (BLA) and quasi-planar conjugated pathway. This should allow electrons to move freely, creating a continuous conjugation pathway for π-electron delocalization. In particular, the calculated BLA for successive C–C bonds of the entire π-bridge and the vicinal C–C bond have shown an increasing trend: SQ-N (0.017 Å) < SQ-Si (0.030 Å) < SQ-Th (0.033 Å) < SQ-P (0.034 Å), which implies an improved degree of π-electron delocalization (smallest BLA) for the dithienopyrrole π-linker. A closely related pattern was found for the BLA of successive C–C bonds calculated over the entire length of the conjugated alkyl chain of molecules, namely SQ-N (0.009 Å) < SQ-Th (0.012 Å) < SQ-Si (0.013 Å) < SQ-P (0.014 Å). Further analysis of cyclic delocalization of mobile electrons over fused rings of the π-bridges was conducted by means of: ring-size normalized multi-centre index (MCI),60 harmonic oscillator measure of aromaticity (HOMA)61 and Shannon aromaticity (SA)62 index at the position of bond critical points (3, −1). All indices were computed using the Multiwfn code based on G09-optimized geometries. Unlike MCI and HOMA indices, the smaller the SA index, the more aromatic the ring is. According to the geometry-based HOMA indicator, electron delocalization decreases in the following sequence: SQ-N (0.69) > SQ-Th (0.65) > SQ-P (0.42) > SQ-Si (0.38). Similarly, SQ-N (0.017) > SQ-Th (0.021) > SQ-P (0.027) > SQ-Si (0.051) was the order that the SA index indicated. Meanwhile, the MCI indicator has predicted the trend: SQ-Th (0.48) > SQ-N (0.46) > SQ-Si (0.41) > SQ-P (0.31). It is credible to draw the conclusion that a comparable electron delocalization potential is exhibited by dithienopyrrole and thienothiophene π-bridges, which is superior to those of dithienophosphole and dithienosilole. A promoted π-conjugation is anticipated for SQ-N and SQ-Th, accordingly. Regarding the conjugation pathway and molecular planarity, the D1-A1-D2 segment of all the molecules has attained planarity in both ground-state and excited-state geometries. A maximum deviation of ca. 1.7° was noted in SQ-Si ground-state for both β1 and β2 dihedral twist angles. These features indicate that donor moieties' orientation and thus, π-conjugation with the squaric ring is not affected by the insertion of π-bridges at the terminals. Planarity along π-linker rings was preserved with a maximum deviation of ca. 0.82°, 0.79° and 0.56° and 0.62° for the inner thiophene ring in SQ-P, SQ-N, SQ-Si and SQ-Th, respectively. Upon photoexcitation, this planarity was boosted with the corresponding values being 0.56°, 0.001°, 0.02° and 0.43°, implying enhanced conjugation for SQ-N due to the presence of the dithienopyrrole π-linker. It also can be detected that planarity and thus conjugation was conserved between the introduced π-spacer and terminal cyanoacrylic acid acceptor CAA (A2) in all the molecules, with maximum deviation recorded (φ1 ≃ 3°) for excited geometries. On the other hand, the dihedral twist angle (φ2) in ground-state geometries recorded the largest values (φ2 = 23–25°). For excited-state geometries, results have shown a geometrical relaxation (φ2 = 0.3–12°) concerning the photo-excited rotation about the C–C single bond, connecting the indole ring and π-spacer, with a tendency to align with the molecular plane. One exception is SQ-P dye (φ2 = 24°), attributable to the relative bulkiness of the P-acetyl dithienophosphole oxide linker, and hence the space around the indole ring could not accommodate a planar alignment. This twist angle (φ2) can be considered the most sensitive (φ2 = 0.3–24°) to the nature of π-linkers during such molecular structure perturbations of relaxed optically excited geometries. For SQ-N dye, the greatest rotation recorded (25° → 0.3°) for the dithienopyrrole linker implies that the resonance effect dominates over the steric hindrance competing effect upon optical excitation. This is consistent with the dye displaying the lowest BLA (0.005 Å), corresponding to a higher degree of delocalization. Since enlarging the photo-induced geometrical relaxation ensures a large amount of torsional work and boosts the Stokes shift,63 it is therefore reasonable to predict that SQ-N dye would display the largest Stokes shift and, hence, minimum self-quenching. Considering the negligible twisting about the spacer–acceptor dihedral angle (φ1 = 0.3–3°), which reflects their maintained planarity and effective p-orbital overlap, one could argue that ease of charge flow and electron injection into the CB of TiO2 are estimated for all modelled dyes.
![]() | ||
Fig. 3 Calculated orbital energy diagram and FMO contour plots (isovalue = 0.02) of the modelled squaraine dyes. |
Dye | State | λ (nm) | E (eV) | ƒ | τe (ns) | Major MO contributions | Dominant character of transitions |
---|---|---|---|---|---|---|---|
SQ-N | S1 | 697 | 1.78 | 1.79 | 4.07 | H → L (99.3%) | HLCT: ICT (64%) + π → π* LE (28%) |
S2 | 586 | 2.12 | 1.08 | 4.74 | H → L + 1 (93.5%); H − 1 → L (5.9%) | HLCT: π → π * LE (64%) + ICT (26%) | |
SQ-P | S1 | 845 | 1.47 | 0.83 | 12.86 | H → L (99.5%) | HLCT: ICT (86%) + π → π* LE (10%) |
S2 | 599 | 2.07 | 2.01 | 2.68 | H → L + 1 (89.3%); H − 1 → L (9.1%) | HLCT: π → π* LE (78%) + ICT (13%) | |
SQ-Si | S1 | 745 | 1.66 | 1.21 | 6.88 | H → L (99.3%) | HLCT: ICT (77%) + π → π* LE (16%) |
S2 | 596 | 2.08 | 1.45 | 3.68 | H → L + 1 (91.1%); H − 1 → L (8.1%) | HLCT: π → π* LE (71%) + ICT (17%) | |
SQ-Th | S1 | 728 | 1.70 | 1.25 | 6.35 | H → L (99.6%) | HLCT: ICT (73%) + π → π* LE (23%) |
S2 | 581 | 2.13 | 1.37 | 3.69 | H → L + 1 (95.9%); H − 1 → L (3.3%) | HLCT: π → π* LE (76%) + ICT (19%) |
The dominant character of main electronic transitions (S1 and S2) was assigned according to relocations of natural transition orbitals (Fig. 4c) and their calculated charge-transfer spectra (Fig. S2†). Natural transition orbitals (NTO) were simulated as a compact orbital representation of the electronic transition density matrix, since several MO pairs contribute non-negligibly to each electronic transition at the same time (Fig. S3†). Also, the charge-transfer spectrum (CTS) can intuitively reveal characters of various peaks of the UV-vis spectrum. The three types of excited states (ES) were estimated as percentages according to the quantitative inter fragment charge transfer (IFCT) analysis of CTS (Table 2), namely, intramolecular charge transfer (ICT), π → π* local excitation (LE) and hybridized local excitation-charge transfer (HLCT) states, wherein the percentage of π → π* LE accounts for the magnitude of intrafragment electron redistribution within the dye fragments. Meanwhile, the interfragment CT from D1-A1-D2 to the π-A2 fragment is denoted by the percentage of ICT (Table 2). It can be observed that all four of the modelled SQ dyes exhibit analogous patterns during the electronic transitions to the first and the second excited states. The first electronic transition S1 ← S0 is primarily (99.3–99.6%) contributed by HOMO → LUMO photoexcitation. In the same manner, two MO pairs were identified to be accountable for S2 ← S0 transition in all the SQ dyes, specifically HOMO → LUMO+1 (89.3–95.9%) and HOMO−1 → LUMO (3.3–9.1%). The molecular moieties D1-A1-D2 and π-A2 are correspondingly where HOMO and LUMO orbitals primarily spatially located in all the SQ dyes (vide supra). For the HOMO−1 orbital (Fig. S3†), the electron density in SQ-P and SQ-Th is distributed throughout the entire molecules, but in SQ-N and SQ-Si, it is mostly located over indole and π-A2 moieties. In turn, the LUMO+1 orbital is majorly confined to D1-A1-D2 in SQ-P and SQ-Si molecules, while being dispersed on entire SQ-N and SQ-Th molecular systems. Thus, all the examined S1 and S2 states can be characterized as hybridized states (HLCT), with contrary hybridization statuses. In other words, the dominance of ICT nature over π → π* LE for S1 and S2 states is in opposition to one another. Specifically, the S1 electronic state of SQ-P, SQ-Si, SQ-Th and SQ-N dyes is a CT-dominated hybridized state exhibiting 86, 77, 73 and 64% ICT character, alongside minor π → π* LE fractions of 10, 16, 23 and 28%, respectively. In contrast, their corresponding S2 states are LE-dominated hybridized, featuring predominantly (78, 71, 76 and 64%) π → π* LE, alongside minor ICT percentages of 14, 17, 19 and 26%. The approximate quantified amount of charge transferred to the π-A2 fragment upon S1 ← S0 CT-dominated photoexcitation is 0.62, 0.85, 0.76 and 0.71 |e−| for the sensitizers SQ-N, SQ-P, SQ-Si and SQ-Th, respectively. Alternatively, the corresponding quantity of transported electrons has diminished to 0.17, 0.06, 0.11 and 0.14 |e−| during S2 ← S0 LE-dominated transition. Remarkably, back charge transfer from the terminal cyanoacrylic acid acceptor (A2) to the π-spacer was found to be negligible in both transitions, with a maximum value of ca. 0.01 |e−| recorded for SQ-N dye's S1 state. These HLCT states have been regarded as unique excited states since they could harvest both high photoluminescence quantum yield sourced from the LE constituent, in addition to exciton utilization contributed by the ICT component.69,70 The comprehensive CT metrics, obtained from both intramolecular (isolated dyes) and interfacial (dye–semiconductor) calculations, are compiled in Table 5, to enable a multiscale investigation of photoinduced electron transport.
Visualizing both hybridization statuses through depicted NTOs (Fig. 4c) validates the previous quantitative IFCT analysis of charge-transfer spectra. Upon red-NIR light photoexcitation to S1 of all the SQ dyes, the occupied NTOs are delocalized throughout the D1-A1-D2 horizontal backbone, exhibiting a slight shift towards thiophene rings of the π-spacers adjacent to the indole donor. Alternatively, the virtual NTOs are mostly localized over the π-spacer and CAA terminal acceptor (A2), with minor contributions originating from the benzoindole-squaraine core-indole (D1-A1-D2) unit. Furthermore, the shown minor contributions of the D1-A1-D2 unit to the virtual NTOs are much pronounced in SQ-N compared to SQ-P, in agreement with their corresponding calculated largest (28%) and smallest (10%) components of π → π* LE. It is therefore highly anticipated that the lowest energy HOMO → LUMO vertical excitation is accompanied by a considerable CT from all SQ sensitizers to the TiO2 substrate upon red-NIR light photoexcitation, with an efficient promotion of exciton dissociation. Considering the second electronic transitions (S2), it is clear that the D1-A1-D2 molecular backbone is the foremost region for both occupied and virtual NTOs. This coincides with the approximated LE-dominant (64–78%) hybridization for S0 → S2 transitions in all four SQ dyes. Additionally, the displayed minor existence of the virtual NTOs over π-A2 fragments, following the order SQ-N > SQ-Th > SQ-Si > SQ-P, is in line with their computed ICT% being 26, 19, 17 and 13%. This implies the capability of photoinduced ICT to be exhibited from the D1-A1-D2 molecular unit to the π-spacer and CAA terminal acceptor upon light-harvesting in the yellow-orange range, especially for SQ-N dye. Moreover, a considerable blue-green light-induced ICT is also predictable for the third electronic transition S0 → S3 of all the SQ dyes, except SQ-Th. Specifically, the simulated UV-vis spectra of SQ-N, SQ-P, and SQ-Si dyes revealed their third absorptions peaking at 491 nm (ε491 = 3.8 × 104 M−1 cm−1), 523 nm (ε523 = 3.4 × 104 M−1 cm−1) and 510 nm (ε510 = 5.7 × 104 M−1 cm−1), respectively. The corresponding ICT features of this S3 state were nearly 5% (0.05 |e−|), 17% (0.15 |e−|) and 13% (0.11 |e−|) from D1-A1-D2 to the CAA terminal acceptor (A2), in addition to 10% (0.11 |e−|), 11% (0.12 |e−|) and 10% (|e−|) from the π-spacers to the CAA acceptor (A2). Again, superior intramolecular CT accompanying photoexcitation to the S3 state is predictable for SQ-P and SQ-Si sensitizers featuring the dithienophosphole oxide and silolodithiophene π-spacers.
Regarding the optimal lifetime for the excited state, it should be sufficiently long to allow for electron injection (∼fs–ps) into the CB of the semiconductor, while being short enough to suppress recombination (∼μs) of the injected electrons with the cationic dye. Here, the lifetime of the excited state (τe) was estimated via: τe = 1.499/(f × E2), in which f is the electronic state's oscillator strength and E is the wavenumber in cm−1 correspondingly.52 For SQ-N, SQ-P, SQ-Si, and SQ-Th sensitizers, the computed excitation energies to the S1 state were ca. 14345.4, 11
837.8, 13
414.6 and 13
742.9 cm−1, respectively. Calculated excitation energies of ca. 17
108.6, 16
681.2, 16
768.3, and 17
216.7 cm−1 corresponded to the S2 state. Generally, the nanosecond scale of calculated τe for the four designed SQ dyes is in good agreement with the fs–ps time scale of charge injection, in addition to being faster than the μs time scale of the unwanted injected charge recombination. Table 2 shows τe decreasing in the order of SQ-P (12.68 ns) > SQ-Si (6.88 ns) > SQ-Th (6.35 ns) > SQ-N (4.07 ns) for S1 states, whereas SQ-N (4.74 ns) > SQ-Th (3.69 ns) ≥ SQ-Si (3.68 ns) > SQ-P (2.68 ns) for S2 states. The longer lifetime for the S1 state advocates a higher probability of radiative decay and reduced non-radiative decay pathways, making SQ-P a good candidate for fluorescent applications. On the other hand, the prolonged S2 of dye SQ-N could imply greater potential for effective electron transfer or charge separation processes before relaxing to S1, which are crucial for DSCs and photocatalysis. Comparison of τe of the modelled dyes with the parent dye, with a calculated τe of ca. 3.44 ns, suggests an increase in excited-state lifetime (4.07–12.86 ns) upon incorporating π-spacers. This could be attributed to increased π-conjugation and enhanced delocalization of excited-state π-electrons. For dye SQ-P (12.86 ns), the highly stabilized LUMO (−3.4 vs. −3.1 eV, Fig. 3) provided an additional advantage, contributing to its greater lifetime.
In order to pursue precise energetics and density-of-states, more sophisticated single point calculations at the B3LYP/6-311G(d,p) level were achieved with G09 software for the SQ@(TiO2)38 optimized geometries. The adsorption energy (Eads, kcal mol−1) was calculated as Eads = E(dye@TiO2) − (Edye + ETiO2). In which, Edye, ETiO2 and E(dye@TiO2) are the energies of isolated SQ dye, bare (TiO2)38 anatase nanocluster and SQ@(TiO2)38 systems, respectively. For all the studied adsorbed squaraines, the large negative values of Eads authenticate spontaneity of dye coverage. Greater negative Eads (−109.9 kcal mol−1) for the SQ-N@(TiO2)38 complex implies relatively higher energy release as the SQ-N dye relaxes on the titania surface, thus stronger electronic coupling. It should be noted that, while the omission of explicit basis-set-superposition error (BSSE) correction could overestimate the values of Eads, the consistent computational protocol applied across all SQ⋯TiO2 complexes guarantees the validity of the comparative trend. The Ti–O bond interaction lengths spanning from 2.0 to 2.2 Å advocate the formation of stable complexes of tightly-adsorbed squaraines onto the TiO2 surface. The bond lengths of O–H bonds due to a carboxylate anchor dissociation are ca. 1.02 Å, and the bond angles of the C–O–C bonds are ca. 126° for all the studied squaraines. Considering the expected deterioration of adsorbed molecular planarity due to induced steric hindrance, all dye segments have attained planarity upon adsorption except for the torsional angle φ2 connecting π-linkers and the indole donor. In particular, φ2 was found to increase in the order of SQ-N (3°) < SQ-Th (8°) < SQ-P (13°) < SQ-Si (17°), which implies that the dithienopyrrole linker provides a rigidified backbone that is less prone to bending, thus upholding the optoelectronic properties of the isolated SQ-N dye. For all the dyes, the geometric features (Fig. 5) reveal their vertical orientation above the TiO2 cluster surface. The maximum inclination, and hence minimum distance between the donor and cluster surface, is exhibited by SQ-Th dye, making it more susceptible for the undesired charge recombination back to the oxidized dye. Therefore, SQ-P and SQ-N dyes, with the beneficial elongated distance between the cation hole and cluster surface, are projected to more strongly suppress this undesirable charge recombination. Regarding the simulated FMO topologies, for all the studied systems, the spatial distribution of electrons in the HOMO is predominantly delocalized over the benzoindole-squaraine core-indole (D1-A1-D2) molecular fragment. Alternatively, the LUMO's electron dispersion is confined to TiO2 nanoparticles. Such an orbital relocation should construct the advantageous photogenerated exciton dissociation at the composite dye/TiO2 interface.
Further elucidation of the electronic coupling and band alignment of modelled SQ⋯TiO2 interfaces can be revealed via projected density-of-states (PDOS). As depicted in Fig. 6, sensitization of TiO2 with squaraines has introduced HOMO and some π-occupied orbitals (magenta discrete lines) within the cluster band-gap. In particular, upon adsorption of SQ-N and SQ-Si dyes, there exist more plainly visible localized states within the band-gap between HOMO (−4.92 eV and −4.90 eV) to HOMO−12 (−7.86 eV) and HOMO−11 (−7.67 eV), respectively. Simulated PDOS for s, p and d orbitals (Fig. S5†) clearly reveals that HOMO corresponds to a purely dye p-orbital and is well distinguished from other TiO2 valence band (VB) states. Predictably, it is also shown that Ti 3p and O 2p orbitals dominate the VB, while Ti 3d orbitals dominate the conduction band (CB). The inserted dye p-orbitals within the band-gap are discrete due to their “molecular” features, and their energies signify the new upper-domain of the VB, thus, minimizing the band-gap energy and broadening the absorption profile. Band gaps for SQ-P, SQ-Si, SQ-N and SQ-Th are 1.19, 1.21, 1.23 and 1.25, respectively. In comparison to that of isolated dye sensitizers (Fig. 3), there is a reduction in energy of the band-gap subsequent to dye adsorption, following the sequence of SQ-N (0.77 eV) > SQ-Th (0.69 eV) ≥ SQ-Si (0.68 eV) > SQ-P (0.49 eV). The new trend of band-gap energies for SQ@(TiO2)32 systems was SQ-P (1.19 eV) < SQ-Si (1.21 eV) < SQ-N (1.24 eV) ≤ SQ-Th (1.25 eV), implying a longer bathochromic shift in the absorption spectrum of SQ-P@(TiO2)32. The most pronounced reduction observed in the band-gap of SQ-N@(TiO2)32 (0.77 eV) can be attributed to stabilization of LUMO (−3.68 vs. −3.04 eV of free dye) upon adsorption, which suggests a greater redistribution of the negative charge over the titania nanocluster. Furthermore, in all the systems, virtual orbitals of the SQ adsorbates were coupled and overlapped within the manifold of titania CB states. The hybridized orbitals have predominant projections onto the p-orbitals of π-spacers as well as the 3d-orbitals of Ti4+ (Fig. S5†). Energetically, the LUMO of free SQ-P, SQ-Th and SQ-Si molecules corresponded to the energy of LUMO+13, LUMO+22 and LUMO+25 levels for the adsorbed dye-semiconductor systems. Meanwhile for SQ-N, the LUMO energy (−3.04 eV) of the free molecule goes much deeper in the CB of the TiO2 supercell and transformed into LUMO+31 of the SQ-N@(TiO2)38 adsorbed complex. This validates the aforesaid stronger SQ-N⋯TiO2 electronic coupling.
An insightful comprehension of the intermolecular chemical bonding at the surface was subsequently attained via bond energy decomposition (BED) analysis. Results are tabulated in Table 4, as obtained by the ADF code, in which the bonding energy is expressed as: ΔEb = ΔEoi + ΔE0. Herein, ΔEoi is the attractive orbital interaction energy, while ΔE0 corresponds to the total steric repulsion. The latter consists of a repulsive term originating from the Pauli antisymmetrization energy (ΔEPauli) for fermion wavefunctions, in addition to a classical electrostatic attraction term (ΔVelstat). In particular, ΔE0 = ΔVelstat + ΔEPauli = ΔVelstat + ΔVPauli + ΔT0, where the potential energy (ΔVPauli) and the kinetic energy (ΔT0) parts are summated into ΔEPauli. Of note, the repulsive character of ΔEPauli is usually attributed to the strong positive kinetic energy ΔT0, and hence the steric repulsion can be considered a kinetic repulsion.72 The computed contributions of steric ΔE0, presented in Table 4, clearly reveal that its positive repulsive character (114.01–121.30 Ha) is owing to the dominance of repulsive destabilizing ΔEPauli (160.47–169.71 Ha) over electrostatic attraction ΔVelstat (−46.46 to −48.40 Ha). Nonetheless, the strong orbital attraction ΔEoi (−169.14 to −178.72 Ha) prevailed Pauli repulsion, resulting in a considerable stabilization of the attractive binding energy ΔEb, with a high negative value (−55.13 to −57.46 Ha). Importantly, ΔEoi accounts for the combined effects of charge transfer (CT) and polarization interactions between two interacting closed-shell systems. Charge transfer involves electron donation from occupied orbitals of the donor to virtual orbitals of the acceptor, while polarization is attributed to the occupied-virtual orbital interactions within each fragment due to the influence of the neighbouring fragment. For the present adsorbed systems, the calculated ΔEoi values (Ha) followed the order of −178.7 (SQ-N) > −177.8 (SQ-Si) > −177.5 (SQ-P) > −169.1 (SQ-Th). Consequently, it can be claimed that the SQ-N@(TiO2)38 adsorbed complex will exhibit the most effective polarization and interfacial CT characteristics.
Complex | ΔE0 | ΔEPauli | ΔVelstat | ΔEoi |
---|---|---|---|---|
SQ-P@(TiO2)38 | 120.57 | 168.94 | −48.37 | −177.49 |
SQ-N@(TiO2)38 | 121.30 | 169.71 | −48.40 | −178.72 |
SQ-Si@(TiO2)38 | 120.35 | 168.51 | −48.16 | −177.80 |
SQ-Th@(TiO2)38 | 114.01 | 160.47 | −46.46 | −169.14 |
Dye | State | Isolated dye | State | Adsorbed dye onto TiO2 | Net CT (Δq |e−|) | ||||||
---|---|---|---|---|---|---|---|---|---|---|---|
DCT (Å) | t-Index (Å) | Sr (a.u.) | EC (eV) | DCT (Å) | t-Index (Å) | Sr (a.u.) | EC (eV) | SQ → (TiO2)38 | |||
SQ-P | S1 | 12.96 | 8.29 | 0.33 | 1.52 | S10 | 22.03 | 18.26 | 0.08 | 0.61 | 0.92 |
S2 | 0.84 | −4.92 | 0.75 | 3.06 | S12 | 21.38 | 16.68 | 0.16 | 0.69 | 0.76 | |
S3 | 6.27 | −0.78 | 0.67 | 2.41 | S15 | 22.58 | 18.53 | 0.09 | 0.64 | 0.90 | |
SQ-N | S1 | 9.79 | 3.49 | 0.53 | 2.17 | S19 | 22.18 | 18.22 | 0.07 | 0.63 | 0.97 |
S2 | 2.73 | −3.99 | 0.73 | 2.79 | S22 | 21.84 | 17.50 | 0.11 | 0.66 | 0.92 | |
S3 | 0.90 | −5.88 | 0.72 | 2.85 | S25 | 21.77 | 16.35 | 0.13 | 0.64 | 0.90 | |
SQ-Si | S1 | 11.24 | 6.05 | 0.43 | 1.78 | S19 | 21.60 | 17.76 | 0.08 | 0.60 | 0.94 |
S2 | 1.09 | −4.52 | 0.75 | 2.84 | S22 | 19.37 | 14.58 | 0.18 | 0.77 | 0.74 | |
S3 | 2.94 | −3.10 | 0.73 | 2.70 | S24 | 23.17 | 19.03 | 0.09 | 0.60 | 0.94 | |
SQ-Th | S1 | 10.47 | 5.13 | 0.45 | 2.04 | S22 | 18.02 | 13.93 | 0.15 | 0.79 | 0.83 |
S2 | 1.87 | −3.81 | 0.73 | 3.06 | S25 | 20.47 | 16.14 | 0.13 | 0.68 | 0.89 | |
S4 | 4.64 | −2.32 | 0.70 | 2.69 | S30 | 18.55 | 13.96 | 0.18 | 0.74 | 0.84 |
![]() | ||
Fig. 7 Charge density difference (CDD) plots for adsorbed systems along with the corresponding contribution of each dye to hole (cyan) and electron (purple) distributions. |
Explicitly, for the SQ-P@(TiO2)38 complex, the dominant electronic transitions (S10, S12 and S15 states) corresponded to photoabsorption at 903, 885 and 840 nm with an oscillator strength (f) of 0.08, 0.18 and 0.06, respectively. All excited states were characterized by separated electron–hole populations with a positive t-index (16.7–18.5 Å), low Sr index (0.09–0.16 a.u.), and EC (0.59–0.64 eV), in addition to an average DCT of ca. 21.99 Å, being about 14 times longer than the C–C bond length. Concerning SQ-N@(TiO2)38, three main absorption peaks at 777 nm (f = 0.057), 738 nm (f = 0.151) and 710 nm (f = 0.263) arise from the S0 to S19, S22 and S25 electronic transitions, respectively. Akin to SQ-P@(TiO2)38, CT descriptors demonstrated an evident CT nature of these states having a positive t-index (16.4–18.2 Å), a low Sr index (0.07–0.13 a.u.), and an average DCT of ca. 21.93 Å, which is in close proximity to the physical distance separating the squaraine core from the surface Ti atoms. The drastic 71% extent of decrease in EC for its major state S25 (0.64 eV) in comparison to the free molecule, being ca. 2.6 eV, implies that the strong orbital interaction with titania considerably improves the dye's CT characteristics. The S19, S22 and S24 dominant CT states in the SQ-Si@(TiO2)38 composite are due to photoexcitation at 794 nm (f = 0.096), 762 nm (f = 0.203) and 736 nm (f = 0.062). The corresponding values of DCT were 21.6, 19.4 and 23.2 Å, each of which is longer than the average separation degree HCT (5.4–5.9 Å), which was translated into a positive t-index (14.6–19.0 Å) and low overlapping Sr index (0.08–0.18 a.u.). For SQ-Th(TiO2)38, the excitation characteristics of S0 → S22 (λ = 749 nm, f = 0.166), S0 → S25 (λ = 719 nm, f = 0.115) and S0 → S30 (λ = 682 nm, f = 0.081) featured DCT falling within the range of 18.2–20.5 Å, Sr of ca. 0.13–0.18 a.u. and EC of 0.68–0.79 eV, a clear indication of outward-directed CT at the dye⋯TiO2 interface as well.
A detailed analysis was conducted for the net charge Δq collected at the TiO2 nanocluster due to the primary dominant state, namely S12 (f = 0.18) for SQ-P@(TiO2)38, S25 (f = 0.26) for SQ-N@(TiO2)38, S22 (f = 0.20) for SQ-N@(TiO2)38, and S22 (f = 0.17) for SQ-Th@(TiO2)38. Results revealed that the actual charge injected into the substrate, SQ → (TiO2)38, decreased in the following order: SQ-N (0.90 |e−|) > SQ-Th (0.83 |e−|) > SQ-P (0.76 |e−|) > SQ-Si (0.74 |e−|). Thus, the most pronounced interfacial CT can be predicted for SQ-N@(TiO2)38, exhibiting the maximum f and Δq. Additionally, SQ-N had the largest dipole moment, both as free (17.58 D) and adsorbed (13.95 D). This maximally allowed photoexcitation of SQ-N to be initiated from both HOMO and HOMO−1, where the electron density is distributed over D1-A1-D2 and π-spacers, respectively. This is in line with the anticipated electron-donating capability of the dithienopyrrole π-spacer, where HOMO−1 does not participate in the transitions that make up the main band for the remaining three dyes. For further details, Fig. S6† depicts all the simulated orbital contributions of the studied SQ@(TiO2)38 complexes upon photoexcitation. On a related aspect, the unpredicted outperformance of SQ-Th dye over SQ-P and SQ-Si, despite exhibiting promising intramolecular CT features, could be attributed to the stronger electron-withdrawing capability of P-acetyl dithienophosphole oxide (SQ-P) as well as dithienosilole (SQ-Si), which partially snatches the excited states and deteriorates the interfacial CT to TiO2. This is consistent with the CDD maps in Fig. 7, with π-A2 fragments exhibiting greater accumulation of ca. 24% and 26% of the photoexcited electrons (purple isosurface), correspondingly, to S12 of SQ-P@(TiO2)38 and S22 of SQ-Si@(TiO2)38, compared to 17% in the SQ-Th@(TiO2)38 composite. Fig. 7 also evidences that photoexcitation of SQ-N@(TiO2)38, in the deep red-to-NIR range, triggers the most substantial depletion of electron density in SQ-N dye, with calculated hole densities (cyan isosurface) of ca. 90–97%. This could be attributable to a complex interplay of multiple factors: the strongest coupling of SQ-N with the TiO2 surface, including the strongest spontaneous adsorption Eads (−109.9 kcal mol−1), orbital ΔEoi (−178.72 Ha) and electrostatic ΔVelstat (−48.40 Ha) attractive interactions; both HOMO and HOMO−1 contributing to its most intense transition; better alignment of energy levels with the CB of TiO2, with deeper positioning of the dye's LUMO within the CB; besides the most favourable adsorption geometry and spatial orientation, with minimal adsorption-induced deterioration of SQ-N molecular planarity.
![]() | ||
Fig. 8 Shaded surface map with projection for electron localization function (ELF) at the adsorption site of the SQ–TiO2 complex. |
Static β0 at λ = ∞ (ℏω = 0 eV) | ||||||||
---|---|---|---|---|---|---|---|---|
βx | βy | βz | βtot | βprj | β‖ | β‖(z) | β⊥(z) | |
SQ-P | 2584.6 | −140.8 | 57.4 | 2589.1 | 2414.2 | 1448.5 | 34.5 | 11.5 |
SQ-N | −1038.0 | 112.9 | 26.5 | 1044.5 | 875.6 | 525.3 | 15.9 | 5.3 |
SQ-Si | 1364.9 | −276.4 | 3.4 | 1392.6 | 1375.0 | 825.0 | 2.0 | 0.7 |
SQ-Th | −1159.6 | 152.3 | 4.4 | 1169.6 | 1074.2 | 644.5 | 2.6 | 0.9 |
Dynamic dc-Pockels β (−ω; ω, 0) at λ = 1460 nm, ℏω = 0.84 eV and (λ = 1064 nm, ℏω = 1.18 eV) | ||||||||
---|---|---|---|---|---|---|---|---|
βx | βy | βz | βtot | βprj | β‖ | β‖(z) | β⊥(z) | |
SQ-P | 6585.6 | −273.1 | −103.3 | 6592.1 | 6155.9 | 3693.6 | 62.0 | 17.3 |
(29![]() |
(−849.0) | (310.5) | (29![]() |
(27![]() |
(16![]() |
(186.3) | (23.6) | |
SQ-N | −1855.0 | 163.3 | 38.6 | 1862.6 | 1576.4 | 945.8 | 23.2 | 7.7 |
(−3716.8) | (249.2) | (60.4) | (3725.6) | (3182.4) | (1909.4) | (36.2) | (11.8) | |
SQ-Si | 2767.9 | −451.3 | 3.5 | 2804.5 | 2753.7 | 1652.2 | 2.1 | −0.9 |
(6877.3) | (−873.9) | (5.3) | (6932.6) | (6763.4) | (4058.0) | (3.2) | (−7.4) | |
SQ-Th | −2120.5 | 226.4 | 5.7 | 2132.5 | 1975.8 | 1185.5 | 3.4 | 1.0 |
(−4421.5) | (366.6) | (6.8) | (4436.7) | (4142.9) | (2485.8) | (4.1) | (0.9) |
Dynamic SHG β (−2ω; ω, ω) at λ = 1460 nm, ℏω = 0.84 eV and (λ = 1064 nm, ℏω = 1.18 eV) | ||||||||
---|---|---|---|---|---|---|---|---|
βx | βy | βz | βtot | βprj | β‖ | β‖(z) | β⊥(z) | |
SQ-P | −17![]() |
103.3 | −40.7 | 17![]() |
−16![]() |
−9821.3 | −24.4 | −10.3 |
(10![]() |
(1272.3) | (−126.3) | (10![]() |
(10![]() |
(6105.6) | (−75.8) | (−425.7) | |
SQ-N | −25![]() |
844.87 | 231.0 | 25![]() |
22![]() |
13![]() |
138.6 | 18.9 |
(−9470.2) | (−1067.8) | (−153.8) | (9531.5) | (8648.4) | (5189.1) | (−92.3) | (−178.3) | |
SQ-Si | −42![]() |
3051.1 | −41.4 | 42![]() |
−41![]() |
−24![]() |
−24.8 | −121.5 |
(7322.8) | (949.9) | (98.3) | (7384.8) | (6616.8) | (3970.1) | (59.0) | (111.0) | |
SQ-Th | −65![]() |
2976.8 | −18.0 | 65![]() |
61![]() |
37![]() |
−10.8 | −12.7 |
(−9169.0) | (−974.3) | (−60.2) | (9220.8) | (8973.2) | (5383.9) | (−36.1) | (−29.4) |
Dye | λ = ∞ (ℏω = 0 eV) | λ = 1460 nm (ℏω = 0.84 eV) | λ = 1064 nm (ℏω = 1.18 eV) | ||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
αx | αy | αz | αiso | αaniso | αx | αy | αz | αiso | αaniso | αx | αy | αz | αiso | αaniso | |
SQ-P | 3.88 | 1.21 | 0.64 | 1.79 | 2.89 | 4.72 | 1.26 | 0.65 | 2.06 | 3.68 | 6.59 | 1.31 | 0.66 | 2.68 | 5.49 |
SQ-N | 3.65 | 1.15 | 0.51 | 1.73 | 2.81 | 4.22 | 1.21 | 0.51 | 1.92 | 3.34 | 5.02 | 1.28 | 0.51 | 2.17 | 4.07 |
SQ-Si | 3.41 | 1.28 | 0.57 | 1.75 | 2.56 | 3.99 | 1.33 | 0.58 | 1.95 | 3.09 | 4.89 | 1.39 | 0.59 | 2.26 | 3.94 |
SQ-Th | 3.24 | 1.02 | 0.46 | 1.56 | 2.50 | 3.75 | 1.07 | 0.46 | 1.73 | 2.97 | 4.48 | 1.13 | 0.46 | 1.97 | 3.66 |
The static isotropic average polarizability αiso(∞) followed a similar pattern in the order of SQ-N and SQ-Th inverted. Namely, SQ-P (1204.8 a.u.) > SQ-Si (1178.4 a.u.) > SQ-N (1167.1 a.u.) > SQ-Th (1053.3 a.u.). Again, the conjugation length (SQ-N and SQ-Th) contributed less to α than the EW-driven electron displacement (in SQ-P and SQ-Si). Molecular volume offers another perspective on this trend, as polarizability in homologous molecules is predicted to increase with volume. An examination of the relatively higher eigenvalues of polarizability tensors αx = αxx + αxy + αxz (3.24 to 3.88 × 10−22 esu, Table 7) suggests that the molecule's principal polarizability axis is along the x-direction in all the dyes. In addition, the static and dynamic isotropic polarizability (αiso ≈ 1.56 to 2.68 × 10−22 esu), exhibiting comparatively low values relative to polarizability anisotropy (αaniso ≈ 2.50 to 5.49 × 10−22 esu), reflects the molecular anisotropy and electronic asymmetry with enhanced CT directionality. That is, the electron density is more easily delocalized along one molecular axis. A dominant hyperpolarizability (Table 6) along the x-axis was also anticipated for all the modelled dyes, with βx = βxxx + βxyy + βxzz being vastly larger (−1038.0 to 2584.6 × 10−30 esu) than βy (112.9 to −276.4 × 10−30 esu) and βz (3.4 to 57.4 × 10−30 esu). This highly anisotropic NLO response can be attributed to the molecular quasi-planarity in the xy-plane, thus, the strongest π-delocalization and CT occur along x, making βx the primary axial tensorial component. Notably, the permanent dipole moment, being mostly aligned along the x-axis (μx ≈ 88–96% of the total ||, Table S1†), also coincides with the primary polarizability and NLO-active direction (x-axis), which advocates maximization of β, efficient unidirectional CT, and robust macroscopic NLO responses.
Upon comparing the dominant diagonal components βxxx (1035.2 to 2584.8 × 10−30 esu) and the maximum off-diagonal β tensors in all the dyes; βxxy (111.6 to −247.9 × 10−30 esu, Table S2†), it is evident that minority of NLO response is associated with CT transitions that are polarized orthogonal to the molecular dipolar axis. Accordingly, βprj (Table 6), which estimates the projection of the β tensor onto the dipole direction, is respectively 93%, 98%, 92% and 84% of βtot for SQ-P, SQ-Si, SQ-Th and SQ-N, indicating strong alignment. Since β‖ is scaled for experimental purposes (β‖ = 3/5βprj) with statistical averaging over random molecular orientations in isotropic media, β‖ is then reduced, ending up near 50–59% of βtot, implying high efficiency, with the solution-phase β retaining more than half of the intrinsic molecular β. Also, the synergy between the x-axis dipole moment and (hyper)polarizability makes the modelled SQ dyes highly responsive to electric fields applied along this direction. It suggests that orienting the electric field along the x-axis would harness the full potential of the molecule's intrinsic CT efficiency, and hence ensures minimal orientational losses and optimal NLO performance. Analysis of the dynamic βtot/static βtot ratios (Table 6) reveals a highly frequency-dependent polarization and dispersion-driven responses. The values of β(ω) and β(2ω) are greater than the static values for all the dyes, with β(ω) increasing with frequency and β(2ω) decreasing with frequency, reflecting opposing resonance alignment.
For the dc-Pockels form β(ω) at λ = 1460 nm (ℏω = 0.84 eV), the dynamic β enhancement ratios were modest (β(ω)/β0 ≈ 1.8 to 2.5), signifying no direct one-photon resonance with S1/S2 states (all >1.47 eV, Table 2). Meanwhile at λ = 1064 nm (ℏω = 1.18 eV), the ratios were much higher (∼3.8 to 11.5), with dye SQ-P exhibiting the largest augmentation (11.5 × β0), attributable to a stronger near-resonant electro-optic response at 1064 nm. For the SHG form, β(2ω) is massively amplified at lower frequency (ℏω = 0.84 eV, λ = 1460 nm), reaching 31 and 56 × β0 for SQ-Si and SQ-Th, respectively, which validates their optimal applicability for frequency doubling telecom wavelengths (E-band) to red light (730 nm). The exact resonance between the second-harmonic frequency (2ℏω = 1.68 eV) and S1 (1.70 eV) of SQ-Th explains its SHG overestimation (β(2ω)/β0 ≈ 56.1). At higher frequency (ℏω = 1.18 eV, λ = 1064 nm), SHG enhancements are generally lower (∼4 to 9) but still significant, suggesting the suitability of SQ-P (β(2ω) ≈ 10693.1 × 10−30 esu) and SQ-N (β(2ω) ≈ 9531.5 × 10−30 esu) for converting near-infrared to green light (532 nm). The dominance of the axial tensorial βx over βy and βz along with the consistent ratio of βprj/βtot(84–98%), for both the Pockels effect (β(ω)) and SHG (β(2ω)) across both frequencies (0.84 and 1.18 eV), confirms the anisotropic electronic structure and alignment of μ with the primary CT axis and (hyper)polarizability along the x-direction The polarization-angle-resolved hyper-Rayleigh scattering (HRS) intensity profiles at λ = 1064 nm are depicted in Fig. 9. In which the radial distance of the pink curve corresponds to the intensity of scattered harmonic light (I2ωΨV) with respect to the incident polarization angle (Ψ) scanned from −180° to +180°. It can be found that all the dyes exhibit HRS intensity peaks at Ψ = ±90° and minima at Ψ = 0°/180°.
The symmetrical decay of intensity between these angles reflects the alignment-dependent scattering, advocating a maximized scattering efficiency when the external field is optimally aligned with the molecule's dominant NLO axis. The HRS maximum at ±90° is attributable to the dominant influence of 〈β2ZZZ〉 over 〈β2XZZ〉 rotational averages in Bersohn's expression:
I2ωΨV ∝ 〈β2XZZ〉cos4![]() ![]() ![]() ![]() ![]() |
Scanning ω1 and ω2 of β(−(ω1 + ω2); ω1, ω2), depicted in Fig. 10, can serve as a spectral signature of how the molecules respond to any pair of input frequencies. The discrepancy between the previous analytical calculations (Table 6), which predicted the existence of strong SHG peaks satisfying ω1 = ω2, and their absence in this plot is attributed to the insufficient number of excited states (20) for the sum-over-states (SOS) calculation owing to computational constraints, thereby failing to capture the full first hyperpolarizability response. However, the scanning still provides valuable insights into the frequency-dependent electro-optic Pockels (EOP) and other NLO properties. In particular, the abundant peaks appearing for SQ-N and SQ-Th at ω1 + ω2 and ω1 − ω2 imply the two photon processes: sum frequency generation (SFG) and difference frequency generation (DFG), respectively. For instance, SQ-Th exhibits two strong (β ≈ 5.32 × 108 a.u.) peaks at (0.014, 0.064 a.u.) and (0.064, −0.078 a.u.), signifying SFG and DFG output, respectively. That is, SFG involves combining the two photons to create a visible green light with a higher-frequency of ω ≈ 0.078 a.u. (∼580 nm), while DFG generates a lower-frequency photon with ω ≈ 0.014 a.u. (∼3230 nm) corresponding to mid-IR. For SQ-N, two SFG and DFG peaks of β ≈ 1.95 × 108 a.u., with input frequencies of (0.092, 0.016 a.u.) and (0.092, −0.108 a.u.), advocate violet-blue light (ω ≈ 0.108 a.u., λ ≈ 418 nm) and mid-IR (ω ≈ 0.016 a.u., λ ≈ 2830 nm) generation. The six most dominant peaks far exceeding others in height of dyes SQ-P, SQ-Si, SQ-N and SQ-Th exhibited a β of approximately 3.08 × 1012, 3.43 × 109, 8.12 × 108 and 3.47 × 108 a.u. In each dye, four of the six highest peaks correspond to ω1 or ω2 values of 0, implying a significant electro-optics Pockels effect. Meanwhile, the remaining two correspond to ω1 = −ω2, which indicates a remarkable optical rectification (OR) effect i.e., rectifying the optical field into a DC signal. As demonstrated in Fig. 10, for dye SQ-P, the four EOP peaks are observable at the frequencies (a.u.) of (0, ±0.08) and (±0.08, 0), while the two OR peaks appear at (−0.08, 0.08) and (0.08, −0.08). Similarly, dye SQ-Si exhibits EOP at (±0.14, 0) and (0, ±0.14), with OR occurring at (−0.14, 0.14) and (0.14, −0.14). The EOP response of SQ-N is detectable at (±0.11, 0) and (0, ±0.11), while the OR is located at (−0.11, 0.11) and (0.11, −0.11). Dye SQ-Th displays EOP at (±0.06, 0) and (0, ±0.06) and OR output at (−0.06, 0.06) and (0.06, −0.06). These findings advocate the suitability of SQ-P, SQ-Si, SQ-N and SQ-Th dyes for correspondingly processing green light (ω ≈ 0.08 a.u., λ ≈ 566 nm), UV light (ω ≈ 0.14 a.u., λ ≈ 323 nm), violet light (ω ≈ 0.11 a.u., λ ≈ 414 nm) and NIR (ω ≈ 0.06 a.u., λ ≈ 754 nm) in both electro-optic modulation and optical rectification.
![]() | ||
Fig. 10 A scan of the first hyperpolarizability (β) along the direction of μ as a function of the two external field frequencies (ω1, ω2). |
Lastly, the unit sphere representation of static α and β, as developed by Tuer et al.,74 is displayed in Fig. 11 so as to rationalize the directionality of the tensors of (hyper)polarizabilities in a static field. It is perceptible that the elongated α ellipsoids and the β spheres clearly show the anisotropy of polarizability and hyperpolarizability. That is, the (hyper)polarizability tensorial components are primarily dispersed in the molecular plane (xy-plane), while the z-axis components, orthogonal to the molecular plane, are negligible (blue radial arrows), in agreement with the domination of αx and βx tensors, as presented in Tables 6 and 7. For all the dyes, the β spheres highlight, with thick arrows, the induced SHG regions if subjected to an external electric field (EEF). Upon applying two EEFs along the cyan arrow, their combination effect will result in the occurrence of a substantial SHG dipole in the same direction, as exhibited by the red radial arrows. Meanwhile, an induced SHG dipole will appear directed oppositely if two EEFs are exerted simultaneously along the green arrow. If the two EEFs are imposed from top to bottom, as illustrated by the pink arrow, their combination effect will lead to a generated SHG dipole pointing to the left. Overall, the directional maxima of all unit sphere representations reveal the non-centrosymmetry of second-order NLO response and the pronounced push–pull character, which are critical for high NLO efficiency.
The computational investigation of the modelled NLO-chromophores was achieved through a dual approach: analytical (DFT-based) and numerical sum-over-states (TD-DFT-based) methodologies. The π-spacer-functionalized modelled SQ dyes exhibited a twofold up to sixfold enhancement of the off-resonant β0, when compared to dyes' counterpart lacking the π-spacer. The intrinsic second-order NLO response, as evaluated by total β0 (×10−30 esu), revealed a decrease in the sequence of SQ-P (2589.1) > SQ-Si (1392.6) > SQ-Th (1169.6) > SQ-N (1044.5). The permanent dipole, charge-transfer, polarizability and hyperpolarizability were all predominantly aligned along the molecular plane (x-axis), creating highly anisotropic molecules with strong electronic and optical responses along this direction, as evidenced through both calculated numerical parameters and the visualized unit sphere representations. The larger values of EOP β (−ω; ω, 0) and SHG β(−2ω; ω, ω) relative to the static β0 values, with EOP β(ω) increasing and SHG β(2ω) decreasing with frequency, reflected a dispersion-driven response with opposing resonance alignment. Dye SQ-P exhibited the greatest EOP response at both 1064 nm (β(ω) ≈ 29781.7 × 10−30 esu) and 1460 nm (β(ω) ≈ 6592.1 × 10−30 esu), and the strongest SHG signal (β(2ω) ≈ 10
693.1 × 10−30 esu at 1064 nm). The SHG activity of the modelled dyes, as estimated by analytical-DFT, showed a simultaneous potential for NIR-to-green and telecom E-band (1460 nm) to red light conversion. Polarization-angle-resolved HRS intensity profiles predicted the maximum intensity of harmonic light peaking at Ψ = ±90° for all the dyes, with SQ-P exerting maximum off-resonant βHRS of ca. 2.2 × 1011 a.u. For all the dyes, the molecular symmetry of the nonlinear response displayed both dipolar and octupolar characters, with SQ-Si having more pronounced octupolar response (Φ(βJ=3) = 61%). Scanning ω1 and ω2 of β(−(ω1 + ω2); ω1, ω2) predicted the suitability of all the dyes in both electro-optic modulation (EOP) and optical rectification (ω1 = −ω2) effects, with the highest signal intensity (β ≈ 3.08 × 1012 a.u.) exhibited by SQ-P, if the nonzero external wavelength corresponds to the green light (ω ≈ 0.08 a.u., λ ≈ 566 nm). Additional abundant peaks appeared for SQ-N and SQ-Th at ω1 + ω2 and ω1 − ω2 corresponding to sum frequency generation (SFG) and difference frequency generation (DFG), respectively. The divergent trends in β0, α, and μ addressed the need for multifactorial optimization in dye design, with SQ-P exceling in β0 and α owing to strong EW-driven polarization, while SQ-N leveraging asymmetry for high μ and significantly advantageous interfacial CT features. Hence, incorporation of the P-acetyl dithienophosphole oxide spacer and ethyl-dithienopyrrole spacer into squaraines predicts a dependable performance in nonlinear optics and dye-sensitized solar cells, respectively. In conclusion, the current study highlights the role of extended π-conjugation as a critical design strategy in optimizing linear and nonlinear optical activity of squaraine chromophores. While limited to four π-spacers, our quantum mechanical framework established here is readily extensible to broader spacer classes. Future work could explore strong donor–acceptor systems (e.g., carbazole derivatives) and planar π-extended architectures (e.g., indacenodithiophene) to further generalize the derived structure–property relationships and design principles for squaraine-based optoelectronics.
Any data not included in the ESI† is available upon request to E. N. at E-mail: eman.nabil@alexu.edu.eg.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5me00071h |
This journal is © The Royal Society of Chemistry 2025 |