Rohini R. Suradkar,
Dnyaneshwar P. Gholap,
Aarti V. Belambe and
Machhindra. K. Lande*
Department of Chemistry, Dr Babasaheb Ambedkar Marathwada University, Chhatrapati Sambhajinagar, Maharashtra, India. E-mail: mkl_chem@yahoo.com
First published on 29th September 2025
Herein, the synthesis of novel organic molecules 2,6-bis-(4,5-diphenyl-1-imidazole-2-yl)pyridine (3A), 2,6-bis-(7H-acenaphtho[1,2-d]imidazole-8-yl)pyridine (3B) and 2,6-bis-(1H-phenanthro[9,10-d]imidazole-2-yl)pyridine (3C) was reported. They were synthesized by the Debus–Radziszewski imidazole synthetic method and characterized by FTIR, UV-vis, 1H NMR, 13C NMR and mass spectrometry. A density functional theory (DFT) approach was used to compute optical analysis, as well as the study of vibrational, frontier molecular orbitals (FMOs) and global indices of reactivity. The electronic transition was explored through the TD-DFT/B3LYP method, which employs time-dependent density functional theory calculations. The recently synthesized compounds were assessed for their fluorescence characteristics, and encouraging findings indicated that the emission efficiency was enhanced through the modulation of conjugation within a molecule. A highly sensitive and selective fluorescent chemosensor exhibited an “on-off” fluorescence response to Fe3+ with a 1:
1 binding ratio in ethanol.
Based on our literature search, no existing studies have used quantum chemical analysis to examine the reactivity, charge distribution, or geometric and spectroscopic properties of the synthesized molecule. The main aim of the present research is to synthesize compounds and comparative studies through experimental and computational analysis of various parameters and their importance in metal ion detection.
The use of B3LYP/6-311G(d,p) for DFT and TD-DFT calculations provides a consistent framework, but its effectiveness in accurately describing excited-state properties raises concerns. While B3LYP is reliable for ground-state geometries, it struggles with charge-transfer excitations and predicting λmax values for systems with significant delocalization due to its lack of long-range correction. To enhance accuracy and gain deeper insights into photophysical properties, exploring more advanced functionals like M06-2X, CAM-B3LYP, or ωB97XD, which better address delocalization errors and long-range interactions, is essential. Although discussing B3LYP's limitations and conducting preliminary comparisons with these functionals would strengthen the computational methodology.
The current research also aims to refine chemical concepts related to molecular structure by optimizing parameters using efficient and cost-effective computational methods instead of expensive and labor-intensive experimental approaches.20,21 In this manuscript, we present a design and synthesis of a novel molecule synthesized by the Debus–Radziszewski imidazole synthesis method, and characterized by 1H NMR 13C NMR, GC-MS the definitive structure of the molecule was validated using powerful analytical instruments and computational analysis.22 DFT was used to computationally analyze the molecule, revealing their reactivity and active sites.23 This study is highly valuable for metal ion detection. DFT-based simulations of IR and UV-vis spectra enhance its practical application in experimental research, providing a powerful tool for data analysis. The reason for conducting this study is that it yields accurate results for compounds while maintaining a low computational expense. Computational analyses have been performed using DFT/B3LYP/6-311G basis sets at the ground-state level. To explore molecular structures, computational researchers calculated quantum chemical descriptors. Additionally, TD-DFT was used to simulate the UV-vis spectra of each compound.24 Spectroscopic characteristics have additionally been explored by refining their FT-IR vibrational parameters using different basis sets.25–27 This study investigates the photophysical properties of three compounds containing benzil, acenaphthoquinone, and phenanthroline groups by analyzing their optical absorption and photoluminescence. Particular attention is paid to comparing the calculated structural and spectral properties of these ligands with experimental findings.
![]() | ||
Scheme 1 : Synthetic pathway of imidazole-based ligand by using Debus–Radziszewski synthesis method. |
Yield: 76.80%.
Melting point: 118–120 °C.
FTIR (υmax cm−1): 3574.26(N–H), 1422.98(C–N), 1582.88 (CN).
1H NMR (500 MHz, DMSO δ ppm): 8.35 (s, 1H), 7.93–7.92 (m, 10H), 7.82 (t, J = 8.5 Hz, 1H), 7.79 (d, J = 8.2 Hz, 2H), 7.64–7.62 (m, 10H).
13C NMR (500 MHz, DMSO δ ppm): 194.7, 135.4, 132.1, 129.5, 129.4.
Mass C33H20N8O, M+: 516.21.
Yield: 70.16%.
Melting point: 133–135 °C.
FTIR (υmax cm−1): 3746.19(N–H), 1582.52(C–N), 1497.52(CN).
1H NMR (500 MHz, DMSO δ ppm): 8.55 (s, 1H), 8.45–8.43 (m, 3H), 8.09–8.07 (m, 6H), 7.94–7.92 (m, 6H).
13C NMR (500 MHz, DMSO δ ppm): 187.5, 144.2, 135.5, 132.2, 130.4, 128.8, 128.4, 121.1.
Mass C33H20N8O, M+: 459.13.
Yield: 69.20%.
Melting point: 148–150 °C.
FTIR (υmax cm−1): 3756.66(N–H), 1583.11(C–N), 1440.52(CN).
1H NMR (500 MHZ, DMSO δ ppm): 8.30 (s, 3H), 8.03–8.02 (m, 4H), 7.88–7.77 (m, 6H), 7.55–7.52 (m, 6H).
13C NMR – (500 MHz, DMSO δ ppm): 178.9, 135.4, 135.2, 131.1, 129.2, 129.0, 124.3.
Mass C33H20N8O, M+: 512.15.
![]() | ||
Fig. 1 The optimized theoretical geometric structures of the synthesized derivates at DFT/B3LYP/6-311G(d,p). |
These synthesized compounds display behavior indicative of double bonds. However, C–N bond lengths are observed at 1.3818 Å for 3A and 1.3805 Å for 3B, and 3C bond lengths are observed at 1.3720 Å. The observed C–N bond lengths are shorter than typical single bonds, indicating resonance effects within the region of the molecule.33 The N–H bond lengths observed for 3A, 3B and 3C are 1.008, 1.0073 and 1.0069 Å respectively. The variations in bond length and bond angle within the derivatives can be attributed to the existence of intermolecular interactions, alongside lone pair electrons, electronegativity, and conjugation, which significantly influence the molecular framework. As the conjugation increases, a shorter bond length is observed in N–H bonds. The selected bond angles and bond lengths of all optimized molecules are shown in Table 1.
3A | 3B | 3C | |||
---|---|---|---|---|---|
Bond length (Å) | Bond length (Å) | Bond length (Å) | |||
9C–12N | 1.3618 | 10C–50N | 1.3805 | 47C–50N | 1.3720 |
9C–11N | 1.3195 | 10C–51N | 1.3307 | 47C–51N | 1.3204 |
12N–64H | 1.0080 | 50N–11H | 1.0073 | 50N–57H | 1.0069 |
10C–13N | 1.3618 | 13C–49H | 1.3805 | 28C–48N | 1.3720 |
10C–14N | 1.3195 | 13C–48C | 1.3307 | 28C–49N | 1.3204 |
13N–65H | 1.0080 | 49N–14H | 1.0073 | 48N–56H | 1.0069 |
5C–15N | 1.3435 | 1C–6N | 1.3446 | 24N–23C | 1.3433 |
1C–15N | 1.3435 | 5C–6N | 1.3446 | 24N–19C | 1.3433 |
Bond angle (°) | Bond angle (°) | Bond angle (°) | |||
---|---|---|---|---|---|
5C,9C,12N | 122.4747 | 2C,1C,6N | 122.9008 | 20C,19C,24N | 122.9967 |
5C,9C,11N | 126.7058 | 4C,5C,6N | 122.9008 | 22C,23C,24N | 122.9963 |
4C,5C,15N | 122.9302 | 1C,10C,50N | 121.6378 | 19C,47C,50N | 121.8955 |
2C,1C,15N | 122.9302 | 1C,10C,51N | 126.2930 | 19C,47C,51N | 126.1542 |
14N,10C,13N | 110.8191 | 50N,10C,51N | 112.0692 | 51N,47C,50N | 111.9504 |
10C,13N,65H | 124.1463 | 10C,50N,11H | 123.0698 | 47C,50N,57H | 123.4681 |
Assignment | 3A | 3B | 3C | |||
---|---|---|---|---|---|---|
Experimental (cm−1) | Theoretical (cm−1) | Experimental (cm−1) | Theoretical (cm−1) | Experimental (cm−1) | Theoretical (cm−1) | |
N–H | 3574.26 | 3648.24 | 3746.19 | 3658.21 | 3756.66 | 3661.13 |
C–N | 1422.98 | 1448.07 | 1582.52 | 1512.86 | 1583.11 | 1575.56 |
C![]() |
1582.88 | 1590.07 | 1497.52 | 1496.10 | 1440.52 | 1437.10 |
The DFT computations serve as an effective method for distinguishing between C–N (single bond) and CN (double bond) vibrational bands by examining their symmetric and asymmetric stretching modes. Owing to variations in bond order, length, and strength, C–N and C
N bonds display unique vibrational frequencies. In a molecule featuring C–N or C
N groups, these computations can uncover both symmetric and asymmetric stretching vibrations. Typically, asymmetric stretching occurs at higher frequencies compared to symmetric stretching for a particular bond type. By visualizing the displacement vectors for each calculated vibrational mode, one can accurately associate a computed frequency with a specific C–N or C
N stretching motion and further differentiate between its symmetric and asymmetric variations.
The mean absolute deviations for all computed properties compared to the experimental data gathered using the DFT/B3LYP/6-311G(d,p) basis set are detailed in the SI file (Fig. 11). A remarkable correlation between theoretical predictions and experimental findings is confirmed for approximately 75% of the results within a specified deviation range. The most significant mean absolute deviations were noted for the NH in relation to the experimental data.
The determined frequencies were contrasted with their corresponding experimental values and demonstrated a strong correlation between the two. The calculated values exhibited a close alignment with the related experimental figures, while the discrepancies can be attributed to phase transitions between the calculated and experimental vibrational frequencies.
![]() | ||
Fig. 2 FMO's, molecular electrostatic potential and contour image of 3A, 3B and 3C calculated at DFT/B3LYP/6-311G(d,p). |
Based on the EHOMO values, the compound reactivity trend will follow compound 3B > compound 3A > compound 3C. Although lower ELUMO values suggest that compounds have strong electron-accepting abilities and act as reactive entities for electron-donating compounds, the reactivity follows the trend in this sequence, compound 3B > compound 3C > compound 3A. Also, the energy band gap of the synthesized compound follows the order 3A (3.8564) > 3C (3.8257) > 3B (3.2569). When π-orbitals extensively interact throughout a molecule, they create delocalized molecular orbitals. As the degree of π-conjugation rises, the number of molecular orbitals also increases, and the energy differences between them typically diminish. The 3B molecule is likely to exhibit the most extensive and effective π-conjugation, featuring a distinctive five-membered ring linked to a naphthalene unit, which imparts a slightly strained and non-planar nature in contrast to entirely planar structures. The existence of the double bond within the five-membered ring can greatly influence conjugation. The 3C molecule demonstrates high levels of conjugation, likely enhancing efficient π-delocalization and resulting in a potentially narrower gap when compared to systems with less extensive conjugation, whereas the 3A molecule presumably showcases the least efficient π-conjugation. This may stem from the fact that the conjugation in benzil is somewhat disrupted by the single bonds linking the phenyl rings, which allows for a degree of rotational flexibility.
As shown in the SI file (Table 5), the difference between frontier molecular orbitals not only indicates the stability of the molecule but is also connected to its interactions with other entities. Smaller energy differences lead to softer molecules, which are associated with higher reactivity and reduced kinetic stability, thus enhancing the fluorescence activity of the compounds as the energy gaps decrease.
Red < orange < yellow < green < blue.
The presence of a blue hue localized around the hydrogen atoms of imidazole rings contributes to the optimal radical scavenging activity.42,43 All three molecules' electrostatic potential values were assessed (shown in Table 6 in the SI file), revealing that the reactivity order of the compounds exhibits a specific trend.
Compound 3C > compound 3B > compound 3A.
I = −EHOMO |
A = −ELUMO |
EGAP = ELUMO − EHOMO |
η = (I − A)/2 |
χ = (I − A)/2 |
The trend in reactivity for the computed values of compounds should also be considered as an additional characteristic linked to FMOs (Fig. 4); this includes chemical hardness and softness, which are inversely related. A chemical hardness of zero indicates the highest level of chemical softness.42,43 In this scenario, the 3B compound is less rigid in comparison to other molecules, and this characteristic can be understood through the HSAB theory, which states that ‘‘Hard molecules possess greater energy gaps than soft molecules.’’ Since biological systems are made up of soft cells and enzymes, they are more likely to interact and bond with soft molecules rather than with those that are harder.44,45 The substances that exhibit greater dipole moment values (μ) indicate longer bond lengths and enhanced charge distribution, making them more localized molecules or more electrophilic systems. This characteristic is associated with increased conductivity during oxidation processes. Conversely, molecules with lower dipole moments tend to accept a smaller amount of electronic charge compared to those with higher dipole moments.
Compound 3A > compound 3B > compound 3C.
The global reactivity parameters are linked to the fluorescence behavior of the molecule and are clarified through Koopman's theorem, which offers a practical and swift method as illustrated in Table 3.46
Global reactivity descriptors | Compound 3A | Compound 3B | Compound 3C |
---|---|---|---|
Optimization energy (Hartree) | −1622.9826 | −1465.6627 | −1684.7703 |
HOMO (eV) | −5.5092 | −5.3868 | −5.9805 |
LUMO (eV) | −1.6528 | −2.1298 | −2.1549 |
Band gap (eV) | 3.8564 | 3.2569 | 3.8257 |
Ionization potential (eV) | 5.5092 | 5.3868 | 5.9805 |
Electron affinity (eV) | 1.6528 | 2.1298 | 2.1549 |
Absolute hardness (η) | 1.9282 | 1.6285 | 1.9128 |
Absolute softness (σ) | 0.9641 | 0.8142 | 0.9564 |
Absolute electronegativity (χ) | 3.5810 | 3.7583 | 4.0677 |
Higher electronegativity and reduced chemical potential values enhance electronic transfer since compounds with delocalized electron clouds can readily interact with other molecules. Therefore, the ranking of activity should be compound 3C > compound 3B > compound 3A conclusively, we could say that high electronegativity is directly proportional to emission, and low chemical potential is also directly proportional to emission. The energy gap between the valence and conduction bands increases leading to a reduction in the emission intensity however, variations can also occur due to the reduced conjugation in the molecules. This study showed that the molecule possessed a greater number of functional groups, such as CN/C–N and N–H groups, on its surface, and the enhanced conjugation effect contributed to the increased fluorescence activity of the molecule.
This study investigates the structure–activity association by analyzing delocalized π-orbitals and electronic transition energies. Peak absorption wavelengths and their oscillation strengths, which correspond to electronic transitions between frontier molecular orbitals were examined. The HOMO is delocalized across the entire π-conjugated ring system, while the LUMO is primarily located on the pyridine ring. Table 4 summarizes the maximum absorption wavelengths (λmax), oscillation strengths (f), excitation energies (E), electronic transition contributions (ETC %), and the deviation between experimental and theoretical values for all compounds. The calculated λmax values mostly fall within the range of lower-lying singlet electronic transitions (HOMO → LUMO). Each molecule exhibited three excited states with varying transitions. ETC % values highlight the dominant electronic transition for each excited state shown in Fig. 3. The absorption data were compared to experimental values to confirm the computational outcomes.
State | Assignment | Coefficient | Energy of transition (eV) | Wavelength (nm) | Oscillator strength | |
---|---|---|---|---|---|---|
Excitation energies and oscillator strengths of 3A molecule | ||||||
From | To | |||||
S0–S1 | HOMO | LUMO | 0.70015 (98.04%) | 3.3359 | 371.66 | 0.3141 |
S0–S2 | HOMO−1 | LUMO | 0.50879 (51.77%) | 3.6186 | 342.63 | 0.0123 |
HOMO | LUMO+1 | 0.48543 (47.12%) | 3.6790 | 337.01 | 0.2605 | |
S0–S3 | HOMO−1 | LUMO+1 | 0.70062 (98.17%) | |||
![]() |
||||||
Excitation energies and oscillator strengths of 3B molecule | ||||||
From | To | |||||
S0–S1 | HOMO | LUMO | 0.63066 (79.54%) | 2.6722 | 463.98 | 0.0645 |
S0–S2 | HOMO | LUMO+1 | 0.59602 (71.04%) | 2.7176 | 456.22 | 0.0334 |
S0–S3 | HOMO−1 | LUMO | 0.60521 (73.25%) | 3.1374 | 395.18 | 00 |
HOMO | LUMO+1 | 0.35801 (25.63%) | ||||
![]() |
||||||
Excitation energies and oscillator strengths of 3C molecule | ||||||
From | To | |||||
S0–S1 | HOMO | LUMO | 0.70123 (98.34%) | 3.3169 | 373.78 | 0.5240 |
S0–S2 | HOMO−1 | LUMO | 0.38383 (29.64%) | 3.5932 | 345.05 | 0.0525 |
HOMO | LUMO+1 | 0.58229 (67.81%) | ||||
S0–S3 | HOMO−1 | LUMO+1 | 0.63417 (80.43%) | 3.6410 | 340.53 | 0.3492 |
HOMO−1 | LUMO+3 | 0.11034 (24.34%) | ||||
HOMO | LUMO+2 | 0.25695 (13.20%) |
![]() | ||
Fig. 3 Frontier molecular orbitals involved in the electronic absorption transitions of the compounds 3A–3C calculated at TD-DFT/B3LYP/6-311G(d,p). |
TD-DFT calculations were performed on molecules 3A, 3B, and 3C. Three bands at 342.63, 371.66, and 337.01 nm, which correspond to HOMO → LUMO, HOMO−1 →LUMO, and HOMO−1 → LUMO+1 in that order, define the theoretical range of 3A. The shift observed at 342.63 nm results from a 98.04% contribution from the HOMO to LUMO transition. In contrast, the subsequent excitation band at 371.66 nm arises from a 51.77% contribution from the HOMO−1 to LUMO transition. Furthermore, the third excitation band at 337.01 nm corresponds to a 98.17% contribution from the HOMO−1 to LUMO+1 transition.49 As a result, the sole transition states that have effective oscillator strengths of 0.3141, 0.0123, and 0.2605 are S0 → S1, S0 → S2, and S0 → S3, in that order. Fig. 3 illustrates how the FMO orbitals of compound 3A and the movement of electron density influence the electronic transitions. Compound 3B's TD-DFT Fig. 3 shows three bands at wavelengths of 463.98, 456.22, and 395.18 nm that are caused by the transitions HOMO → LUMO, HOMO → LUMO+1, HOMO−1 → LUMO+1 and HOMO → LUMO+1.
TD-DFT simulations forecasted one strong band and two weaker bands overall. From Table 4, it can be inferred that in the gas phase, bands II and III of all three molecules exhibited low oscillator strength (f) values. This suggests that these two bands are of low intensity and involve a forbidden transition. Conversely, band I across all solvents showed moderate oscillator strength values. This implies that band I is more intense compared to bands II and III, indicating it is an allowed transition.
The initial excitation band is observed at 463.98 nm, which is attributed to a 79.54% contribution from the transition HOMO → LUMO. The subsequent excitation band at 456.22 nm shows a 71.04% contribution from the transition HOMO → LUMO+1. The third band, located at 395.18 nm, is associated with contributions of 73.54% from the transition HOMO−1 → LUMO+1 and 25.63% from the HOMO → LUMO+1 transition. The only allowed transition states that exhibit significant oscillator strengths of 0.0645, 0.0334, and 0 were the vertical excitation energy states S0 → S1, S0 → S2, and S0 → S3, respectively. For the theoretical spectrum of compound 3C, the transition occurring at 373.78 nm is attributed to a 98.34% contribution from the HOMO → LUMO transition. In contrast, the second excitation band at 345.05 nm is due to contributions of 29.64% from the HOMO−1 → LUMO transition and 67.81% from the HOMO → LUMO+1 transition. The third excitation band, found at 340.53 nm, corresponds to contributions of 80.43%, 24.34%, and 13.20% from the transitions HOMO−1 → LUMO+1, HOMO−1 → LUMO+3, and HOMO → LUMO+2, respectively. The only valid transition states with relevant oscillator strengths of 0.5240, 0.0525, and 0.3492 correspond to the vertical excitation energy states S0 → S1, S0 → S2, and S0 → S3. In Fig. 3, the orbitals of the frontier molecular orbitals (FMO) and the electron density transfer of compound 3C, which are pertinent to the electronic transitions, are discussed.
In this context, EH and EL represent the energy levels of the Highest Occupied Molecular Orbital (HOMO) and Lowest Unoccupied Molecular Orbital (LUMO) measured in electron volts (eV). The term EH−1 refers to the energy of the orbital that lies one level below the HOMO, while EL+1 indicates the energy of the orbital that is one level above the LUMO. Additionally, it presents the transfer integral for electrons (te) and holes (th), highlighting an improved rate of electron mobility for the molecule (shown in SI file Table 7). The table presents the calculated charge-transfer integral values (te and th) and the ground state dipole moments for three different molecules, 3A, 3B, and 3C. The electron transfer integral (te) for molecule 3A (0.0551) is notably higher than that for 3B (0.0276) and 3C (0.0376), suggesting potentially more efficient electron transport in 3A. Conversely, the hole transfer integral (th) is highest for molecule 3B (0.1512), indicating a more facile hole transfer compared to 3A (0.1132) and 3C (0.1263). This suggests that 3B might be a better hole transporter, while 3A could be more suited for electron transport. Regarding the ground state dipole moments, molecule 3A exhibits the largest value (3.7459 Debye), followed by 3B (3.4124 Debye), and 3C has the smallest dipole moment (1.2903 Debye). The significant difference in dipole moments, particularly the much lower value for 3C, could indicate differences in molecular polarity and charge distribution, which impact their interactions with polar environments or their self-assembly properties.
This spectrum visually represents electron behavior within the conduction and valence bands, revealing the distribution of energy states. The segments at the beginning of the energy axis of the graph, ranging from −20 eV to −5 eV, are referred to as filled orbitals, while the range from −5 eV to 0 eV is known as virtual orbitals. Virtual orbitals are unoccupied and are often termed acceptor orbitals. On the other hand, filled orbitals are identified as donor orbitals. A pronounced density of states (DOS) at certain energy points indicates a high availability of states for occupation. Conversely, a DOS of zero intensity signifies that no states are available for the system to occupy. The overarching blue curve, the DOS spectrum, indicates the density of states at each energy level. A complete analysis would involve discussing the orbital density distribution for key orbitals, particularly the HOMO and LUMO. For instance, if the HOMO is localized on a specific part of the molecule and the LUMO on another, it signifies a charge-transfer character. Such a spatial separation between donor and acceptor regions, visible from the orbital density, is directly linked to the efficiency of intramolecular charge transfer. This efficiency, in turn, impacts the molecule's fluorescence behavior53 DOS diagrams (Fig. 4) display molecular orbital energy values on the x-axis and the relative strength of states on the y-axis. These diagrams for compounds 3A–3C indicate that the HOMO and LUMO are primarily derived from the donor portion, with a minor contribution from the acceptor. The band gap values shown are consistent with those calculated and shown in the SI file (Table 8). A notable overlap between the donor-localized highest occupied molecular orbital and the acceptor-localized lowest unoccupied molecular orbital, combined with a suitable energy difference, promotes effective electron transfer when excited, which has a direct effect on phenomena such as fluorescence. A pronounced charge-transfer characteristic in the excited state, which can be recognized by specific orbital distribution on the donor and acceptor fragments in the density of states (DOS)-projected molecular orbitals, typically results in a diminished oscillator strength and, as a result, weaker or redshifted fluorescence, since the emission involves a transition from a significantly charge-separated state. In contrast, if the excited states maintain considerable local excitation characteristics (less charge separation), one might observe more intense fluorescence.
![]() | ||
Fig. 7 UV-vis spectra of 3A, 3B and 3C with the addition of different metal ions (100 ppm) in ethanol at pH 7.5. Inset: colour change of the probe in the absence and presence of Fe3+ ions. |
All absorption spectra showed no notable changes, except for the addition of Fe3+, which caused a blue shift from 290 nm to 385 nm and a colour transition from colourless to yellow. Initially, when we introduced 2.0 equivalents of Fe3+ ions to a bare receptor 3A, a new complex absorption peak at 345 nm was detected. Similarly, we assessed the selectivity and sensitivity of 3B and 3C towards the aforementioned active metal ions. Both ligands showed considerable responses to Fe3+ ions, with complex absorption peaks observed at 320 nm and 353 nm, while no notable changes were seen in the presence of other tested competitive ions. Clear isosbestic points centered at 345 nm for the 3A ligand, 320 nm, and 353 nm for 3B and 3C respectively, indicate that the metal complexes formation as shown in Fig. 7. All three ligands possessed appropriate requisite sites like amine, nitrogen from the pyridine ring, and a cavity that allows for metal ion complexation, controlling ion selectivity.55,56
![]() | ||
Fig. 8 Fluorescence spectra of all sensor 3A–3C with the adding of diverse metal ions in ethanol (pH-7.5) and the ligands with different metal ions (100 ppm), under UV light. |
The fluorescence spectra were recorded between 250 nm and 350 nm (λex: 300 nm). Each of the three sensors exhibited significant fluorescence quenching in the presence of Fe3+ when contrasted with other metal ions, leading to a distinct colour alteration from colourless to vibrant yellow. These results indicate that the ligand sensor exhibits a significant sensitivity and specificity for identifying Fe3+ ions, accompanied by a noticeable colour change that is likely beneficial for detection by the unaided eye. The probe's emission was determined to arise from the enhanced conjugation within the ring, which features the highest π-conjugation and the lowest HOMO–LUMO energy gap. Since molecules with delocalized electronic clouds can readily coordinate with metal ion systems, the hierarchy should be 3C > 3B > 3A. Ultimately, we can conclude that the highest level of conjugation in the ligand is directly related to the probe's emission. The emission intensity of the probes in an ethanol solution increased following the introduction of the metal cation, and the probe solution changed to a vibrant yellow due to the formation of a ligand to metal charge-transfer complex with Fe3+. It also demonstrated a strong selectivity for Fe3+ and an impressive fluorescence “turn-on” response.
Once the selectivity of the ligand has been established, it is crucial to examine the coordination mode between the ligand and Fe3+ to better understand how the ligand recognizes Fe3+. Job's method was utilized to analyze the binding stoichiometry between the ligand and Fe3+. Solutions containing varying ratios of the ligand and Fe3+ were prepared. The emission spectrum for each solution was recorded, and from the resulting emission spectra, a Job's plot was generated. The Job plots for all three ligands obtained through fluorescence titrations demonstrated a peak emission intensity at approximately 0.5 mol fractions, suggesting that the sensor forms a 1:
1 complex with Fe3+.59–61 (provided in the SI file Spectrum 17–19).
![]() | ||
Fig. 9 Optimized geometries and FMO orbital of studied proposed ferric-based complexes at the DFT/B3LYP/LanL2DZ level of theory. |
In complexes 3D, 3E, and 3F specifically, there are three Fe–N bonds and two Fe–Cl bonds. The measured bond lengths for complex 3D fall between 1.8749 Å and 2.3072 Å. Similarly, complex 3E exhibited bond lengths ranging from 1.8762 Å to 2.2795 Å. Furthermore, the bond lengths in complex 3F were found to range from 1.8734 Å to 2.2866 Å. The HOMO energies for the complexes examined range from −5.6716 eV to −5.8346 eV, whereas the LUMO energies fall between −4.9412 eV and −5.0425 eV. The observed energy gaps (ΔE) for complexes 3D, 3E, and 3F are 0.7303, 0.8147, and 0.7921 eV, respectively, with complex 3D having the smallest energy gap as shown in Table 5.
Molecules | 3D | 3E | 3F |
---|---|---|---|
Optimized energy (Hartree) | −1774.4996 | −1617.2582 | −1772.1424 |
Dipole moment (D) | 1.2338 | 0.6349 | 0.9235 |
HOMO (eV) | −5.6716 | −5.7570 | −5.8346 |
LUMO (eV) | −4.9412 | −4.9423 | −5.0425 |
Band gap (eV) | 0.7303 | 0.8147 | 0.7921 |
Ionization energy (eV) | 5.6716 | 5.7570 | 5.8346 |
Electron affinity (eV) | 4.9412 | 4.9423 | 5.0425 |
Chemical hardness (eV) | 0.3652 | 0.4073 | 0.3961 |
Global softness (eV) | 1.3692 | 1.2274 | 1.2624 |
Chemical potential (eV) | −5.3064 | −5.3497 | −5.4385 |
Electronegativity (eV) | 5.3064 | 5.3497 | 5.4385 |
For complexes 3D and 3F, the HOMO cloud density is uniformly spread across the molecule, while in complex 3E, the HOMO cloud is predominantly concentrated on the imidazole. On the other hand, the LUMO cloud density across all examined complexes is located on the imidazole rings, suggesting the electroactive or reactive properties of the ring.
The transition that occurs from the HOMO to the LUMO is associated with the π–π* transitions of the imidazole ring and Fe3+ ions, respectively. However, in complex 3E, the HOMO is mainly situated on the imidazole ring, whereas the LUMO is present on both the Fe3+ ion and the imidazole ring, suggesting an n–π* transition. The overall charge transfer in all the complexes analyzed moves from the ligand to the metal. The MEP of all three complexes was illustrated in Fig. 10; the prominent red sphere at the center of the image signifies a zone of high electron density. This area is rich in electrons and is likely linked to the central metal atom and the ligands that are directly attached to it. This zone would be the most appealing to positively charged entities, which could be regarded as the most nucleophilic section of the molecule. The blue regions, especially at the upper part of the molecule near the hydrogen atoms, denote an area of low electron density. These regions are deficient in electrons and are likely where a positive charge is concentrated. This section of the molecule would be the most attractive to negatively charged entities, which could be viewed as the most electrophilic section of the molecule.
The synthesized three ligands were used for the detection of the Fe3+ metal ions. The pH assessment revealed that the optimal range lies between 4 and 8, suggesting that the sensors can function effectively in both acidic and neutral environments. The absorption spectra exhibited no significant alterations upon introducing 10 equivalents of numerous metal ions, except that the addition of Fe3+ caused a blue shift from 250 nm to 350 nm, accompanied by a colour transition from colourless to vibrant yellow. The fluorescence spectra demonstrated a clear quenching effect in response to Fe3+. Job plots indicated that the sensors establish a 1:
1 binding ratio with Fe3+. Furthermore, these sensors can facilitate visual detection of Fe3+, significantly enhancing their potential applications.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra04242a.
This journal is © The Royal Society of Chemistry 2025 |