Open Access Article
José Guadalupe
Hernández
a,
Carlos Alberto Huerta
Aguilar
b,
Jayanthi
Narayanan
*c,
Eduardo Daniel Tecuapa
Flores
c,
Pandiyan
Thangarasu
*d,
Aldo Hernández
Ramírez
c,
Karthika
Shanmugam
e and
Mayra Margarita Luna
Martinez
c
aCentro Tecnológico, Facultad de Estudios Superiores (FES-Aragón), Universidad Nacional Autónoma de México (UNAM), Estado de México, CP 57130, Mexico
bInstituto Tecnológico de Estudios Superiores de Monterrey (ITESM), Puebla, CP 72453, Mexico
cDivisión de Ingeniería en Nanotecnología, Universidad Politécnica del Valle de México, Av. Mexiquense s/n esquina Av. Universidad Politécnica, Estado de México, CP 54910, Mexico. E-mail: jnarayanan@upvm.edu.mx
dFacultad de Química, Universidad Nacional Autónoma de México (UNAM), Ciudad Universitaria, CDMX, CP 04510, Mexico. E-mail: pandiyan@unam.mx
eSchool of Chemical Engineering, Vellore Institute of Technology, Vellore 632014, Tamil Nadu, India
First published on 22nd February 2024
Herein, we report the synthesis and characterization of metal complexes of Fe(III), Co(II), and Cu(II) with SALPHEN (N,N-bis(salicylimine)-o-phenyldiammine) and their potential application as sensitizers in dye-sensitized solar cells (DSSCs). The high thermal stability up to 550 °C observed for these compounds suggests that they may be possible candidates for the fabrication of solar cell devices without pyrolysis or thermo-oxidative degradation during the solar cell operation. The photovoltaic parameters of SALPHEN and its metal complexes were investigated experimentally, and it was observed that the metal ion coordination generally enhanced the power conversion efficiency (η) where the maximum η of 46.2% was obtained for the Co(II) complex due to its high molar extinction coefficient and enhanced charge transfer by improved π-conjugation and electron-withdrawing capability. In the theoretical calculations, the B3LYP functional was employed to compute the ground-state geometries and the frontier molecular orbitals for all the complexes. The ground-state and excited-state energy transfer, which specifically contribute to the light harvesting properties of SALPHEN and its metal complexes, were analyzed using the functionals with corrected dispersion CAM-B3LYP/dgdzvp basis set (double zeta) and functional (B3PW91, B3P86/6-311++G(d,p) basis set (triple-zeta)). It was observed that the HOMOs of the metal complexes were localized on the d orbitals with π(SALPHEN nitrogen and O− atoms) orbital character, whereas their LUMOs have π*(O−) orbital character. The light-harvesting efficiency (LHE) was analyzed using the theoretically obtained band-gap values for SALPHEN and its metal complexes. In general, metal ion chelation showed increased charge transfer transitions in both the ground and excited states. Moreover, the charge density difference in each compound was observed through three-dimensional (3D) visualization of its electron density isosurface (contour 0.05 e Å−3), which supports understanding the mechanism involved in the energy transfer.
N–) group act as D–π–A charge-transfer chromophores, showing excellent photo-harvesting properties, which is attributed to the wide-ranging delocalized π-systems between the acceptor (A) and donor (D) units through the azo group.5–7 Furthermore, their photophysical activity can often be enhanced by metal ion coordination, which influences their solar light-harvesting behavior. In particular, metal ions with incomplete valence shell configurations show intrinsic chemical characteristics when there are multi-dentate ligands in the metal complex, which has attracted considerable attention in the development of renewable energy technologies.8–11 The incorporation of suitable metal ions with a suitable ligand usually increases the HOMO–LUMO energy gap, which allows the excited-state electron to have a longer lifetime even at high temperatures, increasing the overall luminescence quantum yield of DSSCs.12 It is known that the efficiency of a DSSC is dependent on the photosensitizers used between the electron donor and acceptor electrodes, which should have a broad and strong light absorption capable of covering the entire visible and near-infrared regions (NIR) and be stable to a certain temperature. In this case, the understanding of the power conversion efficiency (PCE), spectral response range, and stability of the DSSC performance are dependent on the molecular-level study of the photosensitizers.13–16 Specifically, dynamic photo-induced intramolecular charge transfer (ICT) in the excited state reflects the change in both the electronic and structural properties of photosensitizers, which involve π-conjugation extending from the donor to the acceptor and often differ significantly from the ground-state molecular structure.17–20 Hence, it is important to explore the electronic and optoelectronic characteristics of sensitizer molecules to reveal their electron-donating and electron-withdrawing behavior, given that they play a vital role in photo-induced ICT after photoexcitation. Density functional theory (DFT) calculations can be employed to elucidate the role of LMCT (ligand-to-metal charge transfer), MLCT, and MC (metal-centered) in the excited states of photosensitizers present in DSSCs and to corroborate experimental interpretation.21–23 Time-dependent density functional theory (TD-DFT) studies have been carried out to understand the various multifaceted processes in DSSCs and metal–organic complexes.24–26 For instance, the DFT and TD-DFT methods were used to explore the light-harvesting efficiency (LHE), open circuit voltage (Voc), and driving force (ΔGinject) for Ru-based and Fe-based dye molecules, which depend on the relative energy levels of the HOMOs of Ru and Fe with respect to the π (dye) orbital and the LUMOs with π*(dye) orbital.27 A similar study was performed for diamine-dithiolate-based Pt complexes, together with an NBO (natural bond orbital) analysis, helping to identify suitable anchoring groups that enhance the overall electronic and optical properties of sensitizers.27–29 In another study, the enhancement in solar cell efficiency was explored by determining the excited-state energies by DFT for porphyrin analogues over Zn-meso-tetraphenylporphyrin complexes having different β-substituents given that an increase in π-conjugation on porphyrin impacted the electron-withdrawing capability of the substituents.30 SALPHEN-based metal complexes have interesting properties because their conjugation and rigidity are important structural features that their dictate molecular properties. Recently, these complexes have been applied in various fields of research, including metal–organic frameworks, fluorescent sensors, and potential catalysts, and even in biological applications.31 The adaptability of SALPHEN complexes has also been extended to the field of new materials, such as solar cells and scaffold materials.32 SALPHEN ligands can be easily fabricated and tuned and are alternatives to synthetically more challenging ligands, such as porphyrins, while conserving similar functionality.33 However, despite their benefits reported to date, there is limited information available on the effect of the conjugation and structural rigidity of these complexes on their optical performance.34–37 Therefore, SALPHEN-based metal complexes continue to attract attention for exploring new mechanistic insights to understand the process of photo-induced charge transfer during their DSSC performance, which may increase their applicability.35–37 Herein, we report the fabrication of a DSSC using metal complexes of an azo-Schiff base, i.e., N,N′-bis(salicylimine)-o-phenyldiammine [SALPHEN], with Fe, Co, and Cu and its photovoltaic parameters including short circuit current (Isc), open circuit voltage (Voc), and fill factor (FF) were explored. The proposed DSSC design approach is based on the use of earth-abundant metals, which can replace ruthenium; also, the multivalent nature of these metal ions can provide metal-to-ligand charge transfer in the presence of conjugated π-organic ligands, resulting in a high molar extinction coefficient. Furthermore, DFT and TD-DFT studies were performed for the metal complexes to determine their molecular and optical properties. With the use of natural transition orbitals (NTO), the nature of charge transfer in their electronic absorption and emissions was determined in detail. The obtained photovoltaic parameters for the DSSC with the metal complexes were compared with that from theoretical calculations.
B.M.38,39 A thermogravimetric study was conducted for all the compounds using a thermal analysis system consisting of a simultaneous thermogravimetric analysis (TGA)-differential scanning calorimeter (DSC) (NETZSCH model STA 449 F1 Jupiter coupled to a Bruker FT-IR spectrometer and a nitrogen-generating instrument by Parker Balston). The thermal behavior of all the compounds was analyzed using two techniques, i.e., TGA and DSC, which were carried out on a PerkinElmer TGA-4000 and DSC1 (Mettler Toledo) instrument, respectively. Experiments were carried out from 30 °C to 800 °C for TGA and 30 °C to 550 °C for DSC at a heating rate of 10 °C min−1 under a flowing nitrogen atmosphere (20 mL min−1). The molar conductance of the metal complexes in CH3CN, DMF, and DMSO (0.001 M) solutions at 25 °C was measured using a YK-43CD LUTRON digital conductivity meter. The current–voltage measurement of the solar cell was carried out using a SANWA RD701 digital multimeter.
N)] and 1560m [ν(C
C)]; benzene ring in aromatic compounds (ring-stretch, sharp bands) at 1478m [ν(C–C)] and 1278 [ν(C–N)]; O–H bending (in-plane) at 1330–1430s; α-CH2 bending at 1400–1450s; C–N in aromatic amines at 1280–1180s; C–O–H in alcohols (C–O stretch) at 1200–1015vs; ring vibrations in-plane (C–H) deformation at 1330–1300s; stretching vibrations at 1191m [ν(C–O)]. In addition, CH2
CH– as (–CH2 out-of-plane wag) at 950–900vs; pyridine ring (in-plane ring formation) at 757m [ν(C
N)] and 501s [ν(C
C)]; C–N–C in amines (bend) at 510–400s; ring in benzene derivatives (in-plane and out-of-plane and ring-deformations as two bands) at 580–420m-s.
:
1, v/v). Elemental analysis: calcd (%) for C20H16O3N2ClFe: C, 56.70; H, 3.81; N, 6.61; found (%): C, 56.31; H, 3.59; N, 6.59. FT-IR (νmax/cm−1) (Fig. 1): medium-sized stretching vibration in the range of 3458–3392 [ν(O–H)] due to the crystal waters in the Fe(III) complex. Stretching frequency in the range of 2160s, 2034s, and 1978s [ν(Fe–O)]; stretching frequency at 907s [ν(C–O), ν(C–O–Fe)]; bending vibration at 822m [δ(Fe–O)]; stretching frequency at 1496m [ν(C
N)]; bending vibration at 590 [δ (C
N)]. Molar conductance/DMF: 3.93–5.44 Ω−1 cm2 mol−1. Magnetic moment (μeff): 1.82 B.M.
:
1, v/v). Elemental analysis: calcd (%) for C20H16O3N2Co: C, 61.39; H, 4.12; N, 7.16; found (%): C, 61.01; H, 4.33; N, 7.03. FT-IR (νmax/cm−1) (Fig. 1): small bands of two stretching vibrations at 3480 and 3429 [ν(O–H)] due to the crystal waters in the Co(II) complex; stretching frequency in the range of 2160s, 2034s, and 1978s [ν(Co–O)]; stretching frequency at 907s [ν(C–O), ν(C–O–Co)]; bending vibration at 822m [δ(Co–O)]; stretching frequency at 1496m [ν(C
N)]; bending vibration at 590 [δ (C
N)]. Molar conductance/DMF: 3.93–5.44 Ω−1 cm2 mol−1. Magnetic moment (μeff): 1.78 B.M.
:
1, v/v). Elemental analysis: calcd (%) for C20H14O2N2Cu: C, 63.57; H, 3.73; N, 7.41; found (%): C, 63.17; H, 3.53; N, 7.01. FT-IR (νmax/cm−1) (Fig. 1): stretching frequencies at 2160s, 2034s, and 1978s [ν(Cu–O)]; stretching frequency at 907s [ν(C–O), ν(C–O–Cu)]; bending vibration at 822m [δ(Cu–O)]; stretching frequency at 1496m [ν(C
N)]; bending vibration at 590 [δ(C
N)]. Molar conductance/DMF: 3.93–5.44 Ω−1 cm2 mol−1. Magnetic moment (μeff): 1.91 B.M.
:
water (1
:
1) solution until complete desorption of the dye, and the desorbed dye solution was measured using a UV-visible spectrophotometer.
The vertical excitation energy and oscillation strengths were also calculated for the complex geometry in the ground state. The natural transition orbitals (NTOs) were derived for [Fe(SALPHEN)Cl(H2O)], [Co(SALPHEN)(H2O)], and [Cu(SALPHEN)] to determine the particle-hole position in the excited-state using B3LYP/DGDZVP.55,56 The charge difference density was also calculated by three-dimensional (3D) visualization of the electron density isosurface (contour 0.05 e Å−3). The HOMO and LUMO were derived for the complexes using B3LYP/DGDZVP to understand the influence of their structure on the performance of the DSSC.57 The visualization of the molecular system in the excited state can support the mechanism involved in the photo-induced charge transfer and energy transfer.58
N)] and 1560 cm−1 [ν(C
C)] and the bending vibrations at 757 cm−1 and 501 cm−1, which are characteristic of the Schiff base ligands.60 In addition, the stretching vibrations corresponding to [ν(C–C)] for the benzene ring, ν[C–O] for the phenolic, and [ν(C–N)] for the azomethine groups were observed at 1478, 1278, and 1191 cm−1, respectively.61
All the IR frequencies (νmax/cm−1) were consistent with the theoretical DFT-IR spectra (Fig. S1–S5, ESI†). Similarly, the experimental FTIR spectra of the metal complexes were compared with that of the ligand (SALPHEN), and it was observed that there was a remarkable shift in the frequency for the metal complexes compared to the ligand, especially the peaks for the azomethine ν[C
N] and phenolic ν[O–H] groups shifted to a lower frequency for the complexes, which suggests the involvement of the nitrogen and oxygen atoms in the coordination bond formed with the metal(II) ions. For example, for the Fe(III) complex (Fig. 1), two peaks appeared at 3458 and 3392 cm−1 due to the ν[–OH] stretching vibration, which correspond to the crystal water molecules, suggesting the presence of crystal water in the iron complex and confirmed by the thermal and elemental analyses.62 The coordination of water molecules with Fe(III) was confirmed by the FTIR signal at 600 cm−1.63 The appearance of a broad peak at around 2160–1978 cm−1 originated from the ν[M–O] and ν[M–N] vibrations as the characteristic of the metal coordination with the Schiff base ligand and confirms the involvement of the nitrogen and oxygen atoms of azomethine and phenolic (–OH) groups. Also, there was a red shift in the C
N stretching vibration, which was detected for the ligand at 1610 cm−1 and shifted to 1592 cm−1 for the complex due to a shift of the azomethine nitrogen lone pair density toward the metal ion during the coordination of the metal ion. This was further supported by the signals at 822, 717, and 671 cm−1, corresponding to M–N and M–O originating from the azomethine and phenolic groups, respectively. Furthermore, the peak for ν(C− ≪ O) observed at 1278 cm−1 for the free ligand shifted to 907 cm−1 for the complex due to the formation of the C–O–M bond.64 The same observation was found for the Co(II) complex (Fig. 1), where the ν[O–H] stretching vibration observed at around 3480–3429 cm−1 corresponds to the crystal water molecules in the coordination sphere. This was further confirmed by the bands corresponding to ν[O–H] (645 cm−1), which indicate the presence of coordinated water molecules.65 The coordination of azomethine and phenolic groups (from the ligand) with the Co(II) ion generated the IR peaks at 2156, 2031, and 1977 cm−1 assigned to the ν[M–O] and ν[M–N] bonds, which were further supported by the bands at 876, 822 and 670 cm−1.66 Alternatively, for the Cu(II) complex (Fig. 1), the vibrations corresponding to the azomethine nitrogen ν[C
N] and phenolic oxygen ν[O–H] to shifted to lower frequencies for the complex, which were observed at 1632, 1664 cm−1, and 3216 cm−1, while these peaks for the ligand were observed at 1680 cm−1 and 3208 cm−1, confirming the coordination of nitrogen and oxygen atoms with the Cu(II) ion.67 As observed for the other complexes, no ν[–OH] stretching vibration was observed in the range of 3480–3429 cm−1, suggesting the absence of water molecule coordination with Cu(II) in the complex.
In addition, the number of vibrational modes were compared with the theoretically obtained values. The number of vibrational modes for the Co complex is 120; for the Fe complex, it is 123; and for Cu, it is 111. However, according to the theoretical calculations by DFT, the vibrational mode was 129 for the Co complex, 132 for the Fe complex, and 120 for the Cu complex. There is was slight increase in the vibration frequency in the optimized geometry due to the gaseous state used for the theoretical calculations.
N) and phenolic oxygen (OH) of SALPHEN.68 In the UV spectra of the metal complexes, the transition at around 250 nm remained unchanged during the formation of the complex due to the intra-molecular π → π* transition within the ligand molecule. The transition at around 331–393 nm is associated with the n → π* transition originating from the phenolic groups involving a bathochromic shift during the complexation, indicating the ligand coordination with the metal ion. For example, the Fe(III) complex exhibited transitions at 299 and 288 nm originating from the LMCT and n → π*, which correspond to phenolic oxygen and coordinated water molecule present in the coordination sphere, respectively. In addition, bands were observed at 329 and 334 nm, corresponding to the 6A1g → 6T2g and 6A1g → 5T1g spin and multiplicity forbidden transitions, respectively, confirming the presence of a low-spin octahedral configuration for Fe(III) (d5).69 The low-spin nature of Fe(III)-SALPHEN was further supported by measuring the magnetic moment (μeff) of 1.82 B.M. at room temperature, showing its paramagnetic nature with one unpaired electron.
| Compound | (λ) nm | Electronic transitions | Molar extinction coefficient (ε) (M−1 cm−1) |
|---|---|---|---|
| SALPHEN | 247 | π → π*; n → π* | 4.91 × 103 |
| 319 | n → π* | 3.12 × 103 | |
| 387 | n → π* | 3.06 × 103 | |
| Fe(III)-SALPHEN | 249 | π → π*; n → π* | 2.04 × 105 |
| Low spin d5 | 288 | n → π* | 1.66 × 105 |
| Octahedral55 | 299 | LMCT | 1.57 × 105 |
| 335 | 6A1g → 6T2g | 1.21 × 105 | |
| 378 | 6A1g → 5T1g | 5.08 × 103 | |
| Co(II)-SALPHEN- | 250 | π → π*; n → π* | 7.83 × 104 |
| Low spin d7 | 294 | MLCT | 4.54 × 104 |
| Square pyramidal56 | 331 | 4A2 + 4E → 4A2(P) | 4.37 × 104 |
| 393 | 4A2 + 4E → 4E(P) | 1.41 × 104 | |
| 458 | 4A2 + 4E → 4B1 | 1.15 × 104 | |
| Cu(II)-SALPHEN | 287 | n → π* | 1.32 × 105 |
| Low spin d9 | 350 | LMCT | 3.28 × 104 |
| Square planar57 | 457 | 2B1g → 2B2g (dx2−y2 → dzy) | 3.02 × 104 |
| 657 | 2B1g → 2A1g (dx2−y2 → dz2) | 7.71 × 103 |
In the case of Co(II)-SALPHEN, a typical d–d transition at 458 nm corresponding to 4A2 + 4E → 4B1 was observed. The bands (319 and 387 nm) observed for the ligand shifted to 331 and 393 nm for Co(II)-SALPHEN, where the latter peaks are assigned to MLCT, 4A2 + 4E → 4E(P) and 4A2 + 4E → 4A2(P), respectively, which originated from the square pyramidal geometry of the Co(II) complex.70 Similar to the iron complex, the band at 294 nm corresponding to the n → π* transition originated from the coordinated water molecule present in the coordination sphere. The magnetic moment (μeff, 1.78 B.M.) determined for the Co(II) complex corresponds to one unpaired electron, suggesting its low spin and paramagnetic character.
The Cu(II) complex exhibited a n → π* transition at 287 nm and LMCT transition at 350 nm. The typical d–d absorption bands for Cu(II) in the square planar coordination were observed at 657 and 618 nm, which are generated from the 2B1g → 2B2g (dx2−y2 → dzy) 2B1g → 2A1g (dx2−y2 → dz2) transition, confirming its geometry with SALPHEN.57 Also, it was found to be paramagnetic with a magnetic moment, μeff, of 1.91 B.M., which is in good agreement with the spin-only value of the square planar complex.71,72 In addition, the corresponding molar extinction coefficients (ε) were determined and shown in Table 1. In general, the metal complexes exhibit higher molar extinction coefficients compared to SALPHEN. Among the studied metal complexes, Fe(III)-SALPHEN showed the highest molecular extinction coefficient at its maximum absorption values (2.04 × 105, 1.66 × 105, 1.57 × 105, and 1.21 × 105 M−1 cm−1), as well as broader and stronger UV-Vis absorption bands compared to other metal ions. However, Cu(II)-SALPHEN presented the maximum molar extinction coefficient of 1.32 × 105 M−1 cm−1 and the longest wavelength absorption band, reaching up to 657 nm. Furthermore, a very low molar conductance was calculated for all the complexes (Fe(III)-SALPHEN, 3.93; Co(II)-SALPHEN, 2.15; Cu(II)-SALPHEN, 5.44 Ω−1 cm2 mol−1 in DMF), which indicates the existence of neutral and non-electrolytic natures.
) (250–350 nm), and above 250 nm is normally from the oxide ion (O−)
;
;
(MLCT); O− (π) → Md (LMCT); and also from some weak d–d transitions.73
| Compounds | Ground state (GS) | Excited state (ES) |
|---|---|---|
| [Fe(SALPHEN)Cl(H2O)] | S = 1/2 | S = 3/2 |
| Doublet-zeta basis set | Wavelength (λ, nm) | |
| B3LYP/DGDZVP | 369, 420, 499, 1024 | 424, 556, 615, 766 |
| CAM-B3LYP/DGDZVP | 460, 530, 610, 746, 1090 | 468, 514, 645 |
| Triple-zeta basis set | ||
| B3LYP/6-311++g(d,p) | 495, 570, 698 | 537, 607, 738 |
| B3PW91/6-311++g(d,p) | 548, 608, 1056 | 541, 604, 746 |
| B3P86/6-311++g(d,p) | 494, 570, 709 | 520, 599, 761 |
| [Co(SALPHEN)(H2O)] | S = 1/2 | S = 3/2 |
| Doublet-zeta basis set | Wavelength (λ, nm) | |
| B3LYP/DGDZVP | 340, 385, 451 | 319, 342, 408, 493 |
| CAM-B3LYP/DGDZVP | 270, 340, 349 | 270, 291, 354 |
| Triple-zeta basis set | ||
| B3LYP/6-311++g(d,p) | 327, 386, 495 | 292, 339, 397 |
| B3PW91/6-311++g(d,p) | 318, 374, 518 | 304, 328, 377 |
| B3P86/6-311++g(d,p) | 326, 385, 504 | 300, 330, 387 |
| [Cu(SALPHEN)] | S = 1/2 | S = 3/2 |
| Doublet-zeta basis set | Wavelength (λ, nm) | |
| B3LYP/DGDZVP | 241, 289, 327, 439 | 404, 525 |
| CAM-B3LYP/DGDZVP | 268, 288, 377, 537 | 267, 294, 316, 366, 460, 691 |
| Triple-zeta basis set | ||
| B3LYP/6-311++g(d,p) | 297, 352, 408 | 377, 437, 527, 1013 |
| B3PW91/6-311++g(d,p) | 291, 352, 412, 439 | 340, 391, 402, 477, 802, 1312 |
| B3P86/6-311++g(d,p) | 291, 352, 406 | 378, 388, 473, 717, 1241 |
Nevertheless, there was an alteration in the band position for the different metal ions in [M-SALPHEN]n+ (n = 3, M = Fe; n = 2, M = Co and Cu), where the greater intensity of the higher-energy peaks originated from the metal–ligand charge transfer (MLCT). The following orbital electronic transitions were obtained in the ground state and in the excited state by employing the B3LYP/DGDZVP double-zeta basis set: for [Fe(SALPHEN)Cl(H2O)] in the ground state, its orbital transitions were HOMO−X to LUMO+Y (X = 0–6, 10–13; Y = 0 + 4); while in the excited state, they were (X = 0–12, Y = 0 + 5); similarly, for [Co(SALPHEN)(H2O)] in the ground state, its transitions of HOMO−X to LUMO+Y were (X = 0–7, 11; Y = 0 + 2, 4) and its excited-state transitions were (X = 0–7, 11; Y = 0 + 2); and for [Cu(SALPHEN)], its ground-state transitions were (X = 0–4, 6; 9, 14, 15, 21, Y = 0 + 2), its excited-state orbital transitions were (X = 0–4, 9–10, 13–14, 18, 21–22). After analyzing the above-mentioned transitions, the high energy band (413 nm) was assigned to the mixing of the metal orbital LUMO or LUMO+Y (Y = 0, 7) with the HOMOs of the donor (Nazo) in the excited state, and the peaks (997, 507, and 423 nm) for [Fe(SALPHEN)Cl(H2O)]; the signal (488 nm and 416 nm) for [Co(SALPHEN)(H2O)]; and (346 nm and 463 nm) for [Cu(SALPHEN)] (Fig. 3) confirm the coordination of nitrogen atoms with the metal ion.74
Similarly, for the complexes, the electronic orbital transitions such as HOMO−X to LUMO+Y using the B3PW91/6-311++G(d,p) triple-zeta basis set were obtained as follows: for [Fe(SALPHEN)Cl(H2O)] in the ground state, the transitions were (X = 0–7, 11–12; Y = 0 + 4), and in the excited state, they were (X = 0–7, 12; Y = 0 + 5, 11); for [Co(SALPHEN)(H2O)], its ground-state transitions were (X = 0–6, 11; Y = 0 + 2, 4) and its excited-state transitions were (X = 0–6, 8; Y = 0 + 3, 5, 11); and finally, for [Cu(SALPHEN)], its transitions in the ground state were (X = 0–8, 10-11, 13, 20, Y = 0 + 4, 6 + 7) and its transitions in the excited state were (X = 0–5, 7–10, 12–13); Y = 0 + 3, 6, 8 + 13.
As observed in Fig. 3(a)-I, in the double zeta basis set DGDZVP with B3LYP and CAM-B3LYP without diffuse and polarized functions, there is a slight shift in the absorption peak in the spectra. For example, with B3LYP/DGDZVP in the ground state, for [Co(SALPHEN)(H2O)], three peaks (340, 385, and 451 nm) appeared in the ground state spectra are converted to four peaks (319, 342, 408, and 493 nm) in the excited state. The accurate displacement of peaks was observed using CAM-B3LYP/DGDZVP with a dispersion of correlated data. This indicates that better displacement of the bands (270, 340, and 349 nm) was obtained for [Co(SALPHEN)(H2O)] in the ground-state spectrum compared to its excited-state signals (270, 291, and 354 nm). However, Table 2 and Fig. 3(a)-II show that with the use of a triple-zeta basis set 6-311++G(d,p) with B3LYP, B3PW91, and B3P86 with diffuse and polarized functions, a clear shift in the excited-state electronic transitions was obtained compared to its ground-state peaks. For example, for [Co(SALPHEN)(H2O)], with B3PW91/6-311++G(d,p), the best displacement of its electronic transitions was seen in the ground-state peak (318 nm), which appears at 304 nm in the excited state; similarly, the other ground state peaks (374 nm and 518 nm) shifted to 328 nm and 377 nm, respectively, in the excited state (see Table 2 and Fig. 3(a)-II).
The above-mentioned methods were employed to optimize the geometry of [Fe(SALPHEN)Cl(H2O)] and [Cu(SALPHEN)]. After analyzing the results, in contrast to the above-mentioned [Co(SALPHEN)(H2O)] complex, a mixture of peaks was obtained for [Fe(SALPHEN)Cl(H2O)] in the red region of the spectrum, showing a small displacement with a low absorption intensity both in the ground state and in the excited state for both the double-zeta basis set and triple-zeta basis set. For example, for [Fe(SALPHEN)Cl(H2O)], with the triple zeta B3PW91/6-311++G(g,p) basis set, a slight spectral shift to the blue region was found, namely, the ground state peak (548 nm) shifted to 541 nm (towards the blue region) in its spectrum in the excited state. Similarly, the other GS peaks (608 nm and 1056 nm) shifted to 604 nm and 746 nm, respectively, in the ES. In the case of the double zeta basis set, the absorption bands (364 nm, 420 nm, and 499 nm in the ground state) shifted to 420 nm, 556 nm, and 615 nm in the excited state, respectively (in the red region of the electronic spectrum; see Fig. 3(b)(I) and (II).
Similar results were obtained for [Cu(SALPHEN)], where its GS peaks (291 nm, 352 nm, and 406 nm) shifted to 378 nm, 388 nm, and 473 nm in the excited state, respectively. Additionally, two new bands (717 and 1241 nm) were also observed in the excited state in the red region of its spectrum, as shown in Fig. 3(c)(I) and (II).
| Double-zeta basis set | RMSD (Debye) | Dipole moment (μ, Debye) | Polarizability anisotropy (Δα, Debye) | Relative energy values (kcal mol−1) | ||||
|---|---|---|---|---|---|---|---|---|
| [Co(SALPHEN)(H2O)] | (GS) | (ES) | (GS) | (ES) | (GS) | (ES) | (GS) | (ES) |
| B3LYP/dgdzvp | 3.92 × 10−9 | 2.91 × 10−9 | 3.318 × 10−18 | 4.967 × 10−18 | 2.44 × 10−18 | 3.51 × 10−18 | 0 | 6.3 a |
| 3.270 × 10−18 | 0.93 × 10−18 | 62.7 | ||||||
| CAM-B3LYP/dgdzvp | 4.70 × 10−9 | 4.80 × 10−9 | 7.388 × 10−18 | 3.563 × 10−18 | 5.11 × 10−18 | 2.61 × 10−18 | 0 | 6.3 a |
| 1.359 × 10−18 | 1.17 × 10−18 | 69.0 | ||||||
| Triple zeta basis set | ||||||||
| B3LYP/6-311++G(d,p) | 9.0 × 10−9 | 5.45 × 10−9 | 3.342 × 10−18 | 4.141 × 10−18 | 2.24 × 10−18 | 2.22 × 10−18 | 0 | 6.3 a |
| 56.5 | ||||||||
| B3PW91/6-311++G(d,p) | 6.66 × 10−9 | 5.48 × 10−9 | 3.241 × 10−18 | 2.804 × 10−18 | 2.72 × 10−18 | 1.67 × 10−18 | 0 | 3.3 |
| B3P86/6-311++G(d,p) | 8.48 × 10−9 | 3.14 × 10−9 | 3.229 × 10−18 | 3.385 × 10−18 | 2.40 × 10−18 | 1.95 × 10−18 | 0 | 194.5 a |
| 2.499 × 10−18 | 1.17 × 10−18 | 238.5 | ||||||
![]() | ||
| Fig. 4 (a) Root-mean square of deviation (RMSD) value; (b) relative energy; and (c) mean polarizability (Δα) of [Co(SALPHEN)(H2O)] in the gaseous state. | ||
sin
θ) using the values of sin
2θ considering the h, k, and l data, and the correction of the d-values was carried out by comparing the observed density with that calculated using the X-ray powder diffractogram.77
| Lattice constants (Å) | Unit cell volume (Å3) | Crystal system | Space group | Angle (2θ) | d value (Å) | Density (g cm−3) | Axial angle (°) | |||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| a | b | c | α | β | γ | |||||||
| Co(II)-SALPHEN | 10.78 | 10.78 | 10.78 | 1252.73 | Cubic |
Fm![]() |
41.0815 | 2.1945 | 2.500 | 90 | 90 | 90 |
| Cu(II)-SALPHEN | 8.104 | 3.757 | 7.433 | 226.311 | Orthorhombic | Pmna | 21.9183 | 4.0503 | 2.500 | 90 | 90 | 90 |
| Fe(III)-SALPHEN | 5.34 | 5.20 | 15.02 | 417.71 | Monoclinic | P1c1 | 34.87 | 2.57 | 2.028 | 90 | 90 | |
In the case of the Fe(III) complex, 23 reflections with the maximum at 2θ = 34.87° were observed, with the following unit cell parameters: a = 5.34 Å, b = 5.20 Å, c = 15.02 Å and α = 90°, β = 90°, γ = 90°. Given that the complex satisfies the conditions of a ≠ b ≠ c and α = γ = 90° β ≠ 90°, it considered to be monoclinic. Alternatively, for the Co(II) complex, 27 reflections with the maximum at 2θ = 41.08° were detected, satisfying the conditions of a = b = c and α = β = γ = 90° with unit cell parameters of a = b = c = 10.78 Å and α = β = γ = 90° for the cubic system. Similarly, for the Cu(II) complex, it showed 22 reflections with the maximum at 2θ = 21.91° and it obeyed the crystallographic rule of a ≠ b ≠ c and α = β = γ = 90° for an orthorhombic structure, with the unit cell parameters of a = 8.104 Å, b = 3.757 Å, c = 7.433 Å and α = 90°, β = 90°, γ = 90°.78–80 The increase in the number of reflections for all the metal complexes indicates that these structures possess a large specific surface area and provide good surface permeability, and thus the irradiation of light on their planes is expected to be greater compared to other complexes.81
According to the results, it was observed that for the Co(II) complex, there are several crystal planes (311, 222, 400, and 331, appearing in the 2θ range of 25° to 40°) with significant intensities; in contrast, for the Cu or Fe complexes in the same 2θ range, they had a lower number of crystal planes (002, 011, and 202 for the Cu(II) complex and 111 113, and 114 for the Fe(II) complex). In addition, the inter-plane distance (d value) was determined to be 2.19, 4.05, and 2.57 for the Co, Cu and Fe compounds, respectively, showing that the Co(II) complex possesses a comparatively lower value than the other complexes, which indicates that the array of crystal planes is near to each other, facilitating the greater absorption of light.82
In general, the metal coordination increased the thermal stability of SALPHEN, suggesting that these metal complexes can be possible candidates for application in DSSCs. For example, the Co and Fe complexes remained stable up to 555 °C. However, a very small weight loss of approximately 3% was observed at around 250 °C for both complexes, which can be attributed to the evaporation of coordinated water molecules present in both complexes.83 The further mass loss of 23% (Fe-SALPHEN) and 30% (Co-SALPHEN) was observed at around 520 °C, which may be related to the decomposition of some of the fragments resulting from the complex at this temperature.72 In addition, Fe-SALPHEN showed a gradual decomposition between 520 °C and 730 °C, with the weight loss of 23% and 80%, respectively. In contrast, Co-SALPHEN presented a deep decomposition with 30% to 82% weight loss between 550 °C and 700 °C, respectively.
In contrast, Cu-SALPHEN only remained stable up to 285 °C; after that, it started to decompose with a rapid weight loss from 15% to 35% at 335 °C. This is attributed to the thermal degradation of Cu(II) complexes at higher temperatures,84,85 which also showed 75% weight loss at 798 °C.
According to the differential scanning calorimetry (DSC) results for SALPHEN and its metal complexes, as shown in Fig. 6(b), SALPHEN showed a glass transition peak at 133 °C and sharp endothermic peak at 160 °C; however, the presence of metal ions generally decreased the glass transition temperature, which was observed at 106 °C, 67 °C, and 53 °C for the Fe, Co, and Cu complexes, respectively. This is attributed to the decrease in the structural rigidity of the SALPHEN molecule in the presence of metal ions. Similarly, the metal ions expand the thermal-related phase changes in their crystal systems in comparison to the SALPHEN molecule, which can be observed as a decrease in the thermal hysteresis temperature peaks in the metal compounds, i.e., 119 °C, 112 °C, and 86 °C for Fe, Co, and Cu, respectively, whereas that for SALPHEN was at 160 °C, suggesting that these metal compounds can be used as phase change materials (PCM) for energy renewable applications.86 Alternatively, SALPHEN showed an exothermic melting temperature of 225 °C, whereas that of the Fe, Co, and Cu complexes was 424 °C, 395 °C, and 431 °C, respectively. This increased exothermic peak temperature is due to the increased crystalline nature of the metal complexes.87
![]() | ||
| Fig. 7 I–V curves of SALHEN and its metal complexes: (a) SALPHEN alone; (b) Cu-SALPHEN; (c) Co-SALPHEN; (d) Fe-SALPHEN. | ||
| V oc (mV) | I sc (μA) | FF (%) | η (%) | |
|---|---|---|---|---|
| where: Voc = open circuit voltage, Isc = short circuit current, FF = fill factor, η = efficiency of power conversion of DSSC. | ||||
| SALPHEN | 0.56 | 0.096 | 18.08 | 14.7 |
| [Cu(SALPHEN)(OH2)] | 1.1 | 0.544 | 11.78 | 13.3 |
| [Co(SALPHEN)(OH2)] | 2.2 | 0.620 | 15.58 | 46.2 |
| [Fe(SALPHEN)(OH2)Cl] | 0.88 | 0.420 | 15.82 | 15.0 |
The power conversion efficiency of the DSSCs was calculated using eqn (1), as follows:
![]() | (1) |
In addition, the fill factor, FF, is defined as the ratio of the maximum power output product (Pm) to the product of short circuit photocurrent and open circuit voltage, which can be derived from eqn (2), as follows:
![]() | (2) |
The data show that the I–V properties of M(SALPHEN) exhibited a significant change with a change in the metal ion of the complex, given that the nature of the metal ion coordination with SALPHEN differs, impacting the electron-transfer mechanism of the DSSC. Similarly, the metal coordination enhanced the power conversion efficiency (η) of SALPHEN, where the maximum (η) was obtained for the DSSC fabricated using the Co(II) complex (46.2%). In addition, to understand the power conversion efficiency and its corresponding electron transfer mechanism for the newly fabricated metal complex-based DSSCs, theoretical energy calculations of the various excited states of each metal complex were systematically performed by DFT.
This was further explained by the NTOs of the canonical base (Fig. 9) of each complex, which illustrate the hole contribution with the delocalization from the LUMO over HOMO, showing that the hole NTO localization on the metal center and the electron NTO delocalization over the σ* orbital for Cl, π* for the H2O moiety, and σ and π orbitals for the SALPHEN. For example, the absorption maximum (413.17 nm) obtained for [Fe(SALPHEN)Cl(H2O)] gives NTO pairs related to the MLCT states (charge-transfer), suggesting that the excitation electron moves from the occupied orbitals (hole) to the unoccupied orbitals. Thus, the transition would be 109 (HOMO) and 110 (LUMO) for [Fe(SALPHEN)Cl(H2O)] at 0.57251 eV, and its orbital energy is 13.17 kcal mol−1 in the excited state. Similarly, [Co(SALPHEN)(H2O)] yields 101 (HOMO) and 102 (LUMO) at 0.81071 eV with an orbital energy of 18.65 kcal mol−1 in the excited state. [Cu(SALPHEN)], in the excited state (f = 0.6320, λ = 294.64 nm), the transition gives 97 (HOMO) and 98 (LUMO) at 0.59439 eV with an orbital energy of 13.67 kcal mol−1. The decreased HOMO–LUMO energy gap observed for the copper complex suggests its increased charge-transfer (CT) excitations and that it can be used in photovoltaic applications to achieve high solar power generation.
Alternatively, to understand the light harvesting properties of the synthesized metal complexes, different excited-state levels were considered. After analyzing approximately 40 excited-state levels for [Fe(SALPHEN)Cl(H2O)], excited energy level 25 was chosen as the excited state given that it describes a relatively greater oscillator strength (f = 0.0651 and λ = 413.17 nm). Similarly, for [Co(SALPHEN)(H2O)] and [Cu(SALPHEN)], the NTOs resulted in the particle and hole at excited state level 20 for the former and level 35 for the latter complex (Fig. 10 and 11), respectively.
The hole contribution with the delocalization from LUMO over HOMO was identified for each complex, and it shows the hole NTO localization on the metal center and the electron NTO delocalization over the σ* orbital for Cl and π* for the H2O moiety and the ligand σ and π orbitals. The absorption maximum (413.17 nm) obtained for [Fe(SALPHEN)Cl(H2O)] gives NTO pairs related to the MLCT states (charge-transfer), indicating that the excitation electron moves from the occupied orbitals (hole) to the unoccupied orbitals. Thus, the transition would be 109 (HOMO) and 110 (LUMO) for [Fe(SALPHEN)Cl(H2O)] at 0.57251 eV, and its orbital energy is 13.17 kcal mol−1 in the excited state. Similarly, [Co(SALPHEN)(H2O)] yields 101 (HOMO) and 102 (LUMO) at 0.81071 eV with an orbital energy of 18.65 kcal mol−1 in the excited state. [Cu(SALPHEN)], in the excited state (f = 0.6320 and λ = 294.64 nm), the transition gives 97 (HOMO) and 98 (LUMO) at 0.59439 eV with an orbital energy of 13.67 kcal mol−1.
According to the CT excitation in the gas phase, the Co(II)-SALPHEN chromophore can be used in photovoltaic applications to achieve high solar power generation, which determines the efficiency of the photovoltaic device in the open-circuit voltage (Voc) related to the charge transfer excitation energy.
![]() | ||
| Fig. 12 Optimized molecular structures by B3LYP functional: SALPHEN and its frontier molecular orbitals (HOMO and LUMOs) in the gaseous state (a) B3LYP/DGDZVP and (b) B3LYP/6-31G**. | ||
In addition, the observed lower energy unoccupied orbitals (LUMOs) show the overlapping of π-type [SALPHEN nitrogen and O− atoms] with d (metal ion) orbital, explaining the overlapping of dx2−y2 (metal ion) with the orbitals from O−. Therefore the excitations can be from HOMO−X to the unoccupied antibonding orbitals, i.e., π* (LUMO and LUMO+X).
According to the MOs, the decrease in the dx2−y2 energy was observed to follow the order of [Fe(SALPHEN)Cl(H2O)] (−2.63 eV) > [Cu(SALPHEN)] (−2.44 eV) > [Co(SALPHEN)(H2O)] (−2.37 eV) in the ground state (Fig. 13). Similarly, the dz2 energy decreases in the order of [Fe(SALPHEN)Cl(H2O)]2+ (−2.76 eV) > [Cu(SALPHEN)]2+ (−2.22 eV) > [Co(SALPHEN)(H2O)]2+ (−2.06 eV) (Fig. 13). Alternatively, in the excited state, the dx2−y2 dz2 energy decreased considerably for all the metal complexes, which was observed to follow the order of [Cu(SALPHEN)] (−4.50 eV) > [Fe(SALPHEN)Cl(H2O)] (−3.37 eV) > [Co(SALPHEN)(H2O)] (−2.11 eV) and [Cu(SALPHEN)] (−3.06 eV) > [Fe(SALPHEN)Cl(H2O)] (−2.64 eV) > [Co(SALPHEN)(H2O)] (−1.79 eV). In both the ground and the excited state, a decrease in the d orbital energy was observed in the cobalt complexes, suggesting that there is increased repulsion on the d orbitals, especially with the dx2−y2 energy. This allows the easier charge transfer (CT), which is consistent with the Jahn–Teller distortion because of the electrostatic axial interaction (O−) with the Co(II) ions (see ESI,† Fig. S7–S9).
![]() | ||
| Fig. 13 Frontier molecular orbital energy level diagram: (a) in the ground state and (b) in the excited state, low spin. (I) [Fe(SALPHEN)Cl(H2O)]; (II) [Co(SALPHEN)(H2O)]; and (III) [Cu(SALPHEN)]. | ||
In addition, the electron density isosurface of the HOMO and LUMO for all the complexes was analyzed at the B3LYP/DGDZVP level to understand the polarization and charge transfer contributions with the connectivity of atoms through the electron density movements, as shown in Fig. 14. According to the molecular graph for [Fe(III)(SALPHEN)Cl(H2O)], the contour electronic density over the molecular orbital surface shows the connectivity of the atoms (N1, N2, O1, O2, and axial Cl) with the metal ions forming a square pyramid base. A negative electron density for oxygen (−0.501), nitrogen (−0.345), and chloride (−0.282), and a positive electron density (0.623) for Fe(III) in Fe(III)–O in the excited state were detected. Similarly, for [Co(II)(SALPHEN)(H2O)], the electron density was (−0.818) for oxygen, (−0.400) for nitrogen, and (0.720) for Co(II) in Co(II)–O. In the case of [Cu(II)(SALPHEN)], the electron density of (−0.488) for oxygen, (−0.352) for nitrogen and (0.583) Cu(II) was observed for Cu(II)–O. The bonding of oxygen with the metal ions follows the order of Co–O > Cu–O > Fe–O and the negative electron density observed for Co–O suggests that the Co–O bond is so strong that it forms a coordination bond with the ligand. The maximum and minimum of the electron density indicate that there is an interaction between the metal (acceptor) and O or N (donor). According to the results, it is observed that the iron and copper complexes can participate in the higher charge transfer in the gaseous state.
Therefore, to determine the energy conversion efficiency of each solar cell device fabricated using SALPHEN and its metal complex sensitizers, the following equations were applied from the literature.89 For example, to determine theoretical open circuit voltage, the energy of each sensitizer molecule (ELUMO), which requires electron injection from the LUMO to the conduction band energy of TiO2, was used, as shown in the following equation:
| Voc = ELUMO − ECB | (3) |
Similarly, the short circuit current (Isc) for the DSSCs was determined theoretically using the following equation:90
![]() | (4) |
| LHE = 1 − 10−f | (5) |
![]() | (6) |
Thus, the theoretical power conversion efficiency of each of the synthesized complex-based DSSCs was calculated using the eqn (3)–(6) and summarized in Table 6.
| Compound | Spin | E HOMO (eV) | E LUMO (eV) | E 00 | E Sensitizer (eV) | (eV) | ΔGinject (eV) | ΔGreg (eV) | LHE | V oc (eV) | Oscillator strength (f) | Half-life (ns) | |
|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
| Functional/double-zeta basis set | |||||||||||||
| SALPHEN | B3LYP/DGDZVP | GS | −6.154 | −2.134 | 4.02 | 6.154 | 2.134 | 1.866 | 1.354 | 0.95984 | 1.866 | 0.4016 | 0.09855824 |
| B3LYP/6-31G** | GS | −5.826 | −1.83 | 3.996 | 5.826 | 1.83 | 2.17 | 1.026 | 0.97155 | 2.17 | 0.2845 | 0.15523098 | |
| Fe-SALPHEN | B3LYP/DGDZVP | GS | −6.02 | −2.764 | 3.256 | 6.02 | 2.764 | 1.236 | 1.22 | 0.99349 | 1.236 | 0.0651 | 0.6353714 |
| ES | −6.217 | −3.372 | 2.844 | 6.217 | 3.373 | 0.627 | 1.417 | 0.99435 | 0.628 | 0.0565 | 0.6864224 | ||
| Co-SALPHEN | GS | −5.476 | −2.375 | 3.101 | 5.476 | 2.375 | 1.625 | 0.676 | 0.9805 | 1.625 | 0.195 | 0.25635403 | |
| ES | −5.498 | −2.29 | 3.208 | 5.498 | 2.29 | 1.71 | 0.698 | 0.98259 | 1.71 | 0.1741 | 0.28483502 | ||
| Cu-SALPHEN | GS | −5.742 | −2.98 | 2.762 | 5.742 | 2.98 | 1.02 | 0.942 | 0.9368 | 1.02 | 0.632 | 0.07193797 | |
| ES | −5.683 | −4.499 | 1.184 | 5.683 | 4.499 | −0.499 | 0.883 | 0.97066 | −0.499 | 0.2934 | 0.15819262 | ||
| Functional/triple-zeta basis set | |||||||||||||
| SALPHEN | B3PW91/6-311++G(d,p) | GS | −6.305 | −2.307 | 3.998 | 6.305 | 2.307 | 1.693 | 1.505 | 0.97059 | 1.693 | 0.2941 | 0.12821429 |
| Fe-SALPHEN | GS | −6.066 | −2.739 | 3.327 | 6.066 | 2.739 | 1.261 | 1.266 | 0.99078 | 1.261 | 0.0922 | 0.44184088 | |
| ES | −6.314 | −3.198 | 3.116 | 6.314 | 3.198 | 0.802 | 1.514 | 0.99393 | 0.802 | 0.0607 | 0.61944648 | ||
| Co-SALPHEN | GS | −5.482 | −2.441 | 3.041 | 5.482 | 2.441 | 1.559 | 0.682 | 0.98548 | 1.559 | 0.1452 | 0.3435239 | |
| ES | −5.606 | −2.339 | 3.267 | 5.606 | 2.339 | 1.661 | 0.806 | 0.98635 | 1.661 | 0.1365 | 0.3494321 | ||
| [Cu(SALPHEN)]2+ | GS | −5.787 | −2.529 | 3.258 | 5.787 | 2.529 | 1.471 | 0.987 | 0.96657 | 1.471 | 0.3343 | 0.13389313 | |
| ES | −5.558 | −4.872 | 0.686 | 5.558 | 4.872 | −0.872 | 0.758 | 0.96887 | −0.872 | 0.3113 | 0.15587822 | ||
The calculated photovoltaic parameters of the metal complex photosensitizers show that their metal ions and corresponding electronic properties have a significant influence on the power conversion efficiency of the DSSCs. For example, ΔGinject, ΔGreg, LHE, and the excited-state lifetime (τ), which were used to determine the open circuit voltage (Voc) and short circuit current (Isc), have a direct influence on the effectiveness of the DSSC, showing a significant change upon metal coordination. As shown in Table 6, the metal ion generally increased the LHE and half-life time compared to SALPHEN alone, revealing the higher absorption ability of the photons by the metal complexes in the UV-visible region, which can assist in the higher injection of photo-excited electrons into the CB of TiO2. This observation is consistent with the experimental results (Table 5), where the metal complexes showed a higher power conversion efficiency than SALPHEN alone. Furthermore, the decrease in the ΔGinject value after the metal ion coordination suggests the easier injection of electrons into the CB of TiO2. Among the studied complexes, Cu(II)-SALPHEN showed the lowest ΔGinject due to its multi-valence response behavior, which supports its strongest metal to ligand electron transfer. In addition, the greater value of ΔGreg for Co(II)-SALPHEN and Fe(III)-SALPHEN indicates the longer life time of excited electrons in their CB, which can enhance the overall luminescence quantum yield of the DSSC.
In addition, the reactivity parameters such as chemical potential (μ), chemical hardness (η), softness (S), and electronegativity (χ) of the sensitizers (metal complexes) were calculated using EHOMO and ELUMO, which were used to explain the influence of the physicochemical properties on the electronic structure and PCE. The following equations (eqn (7)–(9)) were used and the data are shown in Table 7.
![]() | (7) |
| Electronegativity (χ) = −μ | (8) |
![]() | (9) |
| Compounds | E HF | HOMO | LUMO | ΔE | Hardness (η) | Softness (σ) | Chemical potential (μ) | Electronegativity (χ) |
|---|---|---|---|---|---|---|---|---|
| B3LYP/dgdzvp (double-zeta basis set) | ||||||||
| SALPHEN | ||||||||
| −1031.81 | −6.154 | −2.134 | 4.020 | 2.010 | 0.467 | 4.14 | ||
| [Fe(SALPHEN)Cl(H2O)] | ||||||||
| GS | −2830.85 | −6.020 | −3.256 | 3.256 | 1.628 | 0.614 | −4.390 | 4.390 |
| ES | −2830.87 | −6.217 | −3.372 | 2.844 | 1.422 | 0.703 | −4.795 | 4.795 |
| [Co(SALPHEN)(H2O)] | ||||||||
| GS | −2489.72 | −5.476 | −2.375 | 3.101 | 1.551 | 0.645 | −4.360 | 4.360 |
| ES | −2489.61 | −5.498 | −2.290 | 3.208 | 1.604 | 0.624 | −3.894 | 3.894 |
| [Cu(SALPHEN)] | ||||||||
| GS | −2670.90 | −5.742 | −2.890 | 2.762 | 1.381 | 0.724 | −3.930 | 3.930 |
| ES | −2670.54 | −5.683 | −4.499 | 1.184 | 0.592 | 1.689 | −5.091 | 5.091 |
| B3LYP/6-311++G(d,p) (triple-zeta basis set) | ||||||||
| SALPHEN | ||||||||
| −1031.28 | −6.305 | −2.307 | 3.998 | 1.999 | 0.500 | −4.306 | 4.306 | |
| [Fe(SALPHEN)Cl(H2O)] | ||||||||
| GS | −2830.44 | −6.066 | −2.739 | 3.327 | 1.664 | 0.601 | −4.403 | 4.403 |
| ES | −2830.46 | −6.314 | −3.198 | 3.116 | 1.558 | 0.642 | −4.756 | 4.756 |
| [Co(SALPHEN)(H2O)] | ||||||||
| GS | −2489.62 | −5.482 | −2.441 | 3.041 | 1.521 | 0.658 | −3.962 | 3.962 |
| ES | −2489.61 | −5.606 | −2.339 | 3.267 | 1.633 | 0.612 | −3.973 | 3.973 |
| [CU(SALPHEN)] | ||||||||
| GS | −2670.62 | −5.787 | −2.529 | 3.258 | 1.629 | 0.614 | −4.158 | 4.158 |
| ES | −2670.54 | −5.558 | −4.872 | 0.686 | 0.343 | 2.914 | −5.215 | 5.215 |
The chemical reactivity parameters (Table 7) such as the chemical potential (μ) of the metal complex sensitizers have higher values than that of the TiO2 semiconductor (−4.66 eV),93 suggesting that the electron transfer from the sensitizer to the CB of TiO2 would be relatively straightforward. The chemical hardness (η) of all the compounds was lower than that of TiO2 (3.78 eV),94 indicating that all the sensitizers can release electrons compared to TiO2. Also, the electronegativity (χ) of the metal complexes is lower than that of TiO2 (−4.66 eV), illustrating that TiO2 may easily accept electrons from these sensitizers.
The influence of the excited-state electron lifetime for each sensitizer was determined using eqn (10), as follows:95
![]() | (10) |
| SALPHEN | N,N-Bis(salicylimine)-o-phenyldiammine |
| DSSC | Dye-sensitized solar cell |
| η | Power conversion efficiency |
| MLCT | Metal-to-ligand charge transfer |
| LMCT | Ligand-to-metal charge transfer |
| MC | Metal-centered |
| TGA | Thermogravimetric analysis |
| FF | Fill factor |
| DFT | Density functional theory |
| TD-DFT | Time-dependent density functional theory |
| D | Electron donor |
| A | Electron acceptor |
| NIR | Near-infrared region |
| PCE | Power conversion efficiency |
| ICT | Intramolecular charge transfer |
| LHE | Light-harvesting efficiency |
| V oc | Open circuit voltage |
| ΔGinject | Driving force |
| HOMO | Highly occupied molecular orbital |
| LUMO | Low unoccupied molecular orbital |
| μ eff | Effective magnetic moments |
| DSC | Differential scanning calorimetry |
| FTO | Fluorine-doped tin oxide |
| CB | Conduction band |
| B3LYP | Hybrid functionals Becke, 3-parameter, Lee–Yang–Parr. |
| CAM-B3LYP | Coulomb attenuating method-hybrid functionals Becke, 3-parameter, Lee–Yang–Parr |
| B3PW91 | Exchange component of Perdew and Wang's 1991 functional |
| B3P86 | Hybrid functionals Becke, 3-parameter Perdew 86. |
| FF | Fill factor |
| Φ inject | Electron injection efficiency |
| η collect | Charge collection efficiency |
| ΔGinject | Driving force |
| RMSD | Root-mean square of deviation |
| NBO | Natural bond orbital |
| NTO | Natural transition orbitals |
| f | Oscillator strength |
| τ | Electron lifetime |
| GS | Ground state |
| ES | Excited state |
| μ | Dipole moment |
| Δα | Polarizability anisotropy |
| PCE | Power conversion efficiency |
| NIR | Near-infrared region |
| ICT | Intramolecular charge transfer |
| TGA | Thermogravimetric analysis |
| DSC | Differential scanning calorimetry |
| ε | Molar extinction coefficient |
| λ | Wavelength (nm) |
| ns | Half life |
| μ | Chemical potential |
| χ | Electronegativity |
| η | Chemical hardness |
| Cz-D2 dyes | 4,4′-(9-Octyl-9H-carbazole-3,6-diyl)bis(2,3-difluoro benzaldehyde) [Cz-2PhF2-CHO] |
Footnote |
| † Electronic supplementary information (ESI) available: Optical properties of SALPHEN by DFT, Fukui indexes, dipole moment, polarizability and hyperpolarizability for SALPHEN at gaseous state, bond lengths and bond angles resulted in ground state and excited state for SALPHEN and its complexes, using functional B3LYP, CAM-B3LYP/DGDZVP double-zeta basis set and functionals B3LYP, B3PW91, B3P86/6-311++(d,p) triple-zeta basis set at level of theory, the optimized geometrical data of SALPHEN and its complexes at gaseous state, condensed Mulliken charges of SALPHEN with its metal complexes at gaseous state, TD-DFT spectral data of electronic transitions SALPHEN at gaseous state and it is complexes, optimized structure of metal complexes, theoretical FTIR spectra of metal complexes, TD-DFT spectra of SALPHEN using different functionals at gaseous state, TD-DFT spectra of SALPHEN and its complexes using functional B3LYP/DGDZVP and B3PW91/6-311++G(d,p) basis set at gaseous state, in ground state and in excited state. See DOI: https://doi.org/10.1039/d3ma00982c |
| This journal is © The Royal Society of Chemistry 2024 |