Next Article in Journal
Thermoresponsive Ionic Liquid with Different Cation–Anion Pairs as Draw Solutes in Forward Osmosis
Previous Article in Journal
Fluorescent Pyranoindole Congeners: Synthesis and Photophysical Properties of Pyrano[3,2-f], [2,3-g], [2,3-f], and [2,3-e]Indoles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of New Cobalt(III) Meso-Porphyrin Complex, Photochemical, X-ray Diffraction, and Electrical Properties for Photovoltaic Cells

1
Department of Chemistry, College of Science Al-Zulfi, Majmaah University, Majmaah 11952, Saudi Arabia
2
Laboratory of Physical Chemistry of Materials, Faculty of Sciences of Monastir, University of Monastir, Avenue de L’environnement, Monastir 5019, Tunisia
3
Département de Chimie Moléculaire, 301 rue de la Chimie, Université Grenoble Alpes, CS 40700, CEDEX 9, 38058 Grenoble, France
4
Faculty of Chemistry, Wrocław University of Science and Technology, 27 Wybrzeże Wyspiańskiego, 50-370 Wrocław, Poland
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(24), 8866; https://doi.org/10.3390/molecules27248866
Submission received: 2 November 2022 / Revised: 24 November 2022 / Accepted: 1 December 2022 / Published: 13 December 2022

Abstract

:
The present work describes the preparation and characterization of a new cobalt(III) porphyrin coordination compound named (chlorido)(nicotinoylchloride)[meso-tetra(para-chlorophenyl)porphyrinato]cobalt(III) dichloromethane monosolvate with the formula [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4). The single-crystal X-ray molecular structure of 4 shows very important ruffling and waving distortions of the porphyrin macrocycle. The Soret and Q absorption bands of 4 are very red-shifted as a consequence of the very distorted porphyrin core. This coordination compound was also studied by fluorescence and cyclic voltammetry. The efficiency of our four porphyrinic compounds—the H2TClPP (1) free-base porphyrin, the [CoII(TClPP)] (2) and [CoIII(TClPP)Cl] (3) starting materials, and the new Co(III) metalloporphyrin [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4)—as catalysts in the photochemical degradation was tested on malachite green (MG) dye. The current voltage of complexes 3 and 4 was also studied. Electrical parameters, including the saturation current density (Js) and barrier height (ϕb), were measured.

1. Introduction

Cobalt metalloporphyrin complexes have been extensively investigated since the early sixties, firstly as biomimetic models of hemoproteins, especially of vitamin B12 [1,2,3]. These coordination compounds are indeed explored in many domains, such as catalysts [4], gas detection [5], biological activities, antifungal and antibacterial [6,7], building blocks [8,9], photovoltaic cells [10,11,12], and sensors [13,14]. Organic dyes are used in many industrial applications: paper, cosmetics, but especially in the food, textile, pharmaceutical, and medical diagnostic industries [15,16]. Their disposal is one of the main problems in liquid waste treatment.
The majority of organic dyes are toxic causing serious environmental and human health problems because of their mutagenic [17], teratogenic [18] and carcinogenic properties [19]. Among methods used for the decolorization of dyes, heterogeneous photocatalysis is one of the most used [20,21,22,23,24].
This method is based on the irradiation of a catalyst, usually a semiconductor, which can be photoexcited to create electron-donating or -accepting sites and thus cause redox reactions. If the absorbed photons have an energy greater than the energy difference between the valence and the conduction bands, electron–hole pairs are formed in the semiconductor (holes in the BV band and electrons in the BC band). Notably, porphyrin complexes have been successfully tested as catalysts in the chemical and heterogeneous photocatalysis degradation of organic dyes [25,26,27].
In continuation of our general investigation concerning synthesis, the spectroscopic, electrochemical, and structural characterization of new porphyrin coordination compounds, we recently published several papers concerning zinc(II), magnesium(II) and cobalt(II) metalloporphyrins, which were tested as catalysts in the decomposition of several organic dyes [28,29,30].
To get new insights into the para-chloro-substituted meso-tetraphenylporphyrin (H2TClPP) and the sterically hindered nicotinoyl chloride ligand on the electronic and structural properties of a cobalt coordination compound, the reaction of an excess of nicotinoyl chloride with the (chloride)[meso-tetra(para-chlorophenyl)porphyrinato]cobalt(III) complex ([CoIII(TClPP)Cl]) was carried out in dichloromethane under air. The resulting (chlorido)(nicotinoyl chloride)[meso-tetra(para-chlorophenyl)porphyrinato]cobalt(III) dichloromethane monosolvate with the formula [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4) was characterized by UV-vis, IR, 1H NMR and fluorescence spectroscopy. Cyclic voltammetry data as well as the X-ray molecular structure and the Hirshfeld surfaces analysis of 4 were investigated. A photodegradation investigation on malachite green (MG) dye using the free-base H2TClPP (1), the [CoII(TClPP)] (2) and [CoIII(TPP)Cl] (3) starting materials and complex 4 is reported.

2. Experimental

2.1. Cyclic Voltammetry Experiments

These experiments were measured using a CH-660B potentiostat. All analyses were done under an argon atmosphere at room temperature in a standard three-electrode electrochemical cell. Tetrabutyl ammonium hexafluorophosphate (TBAPF6) was used as the supporting electrolyte (0.1 M) in dichloromethane. The Ag/AgNO3 (TBAPF6 0.2 m in CH2Cl2) redox couple was used as the reference electrode. All potentials quoted in the text were transformed into values for the saturated calomel electrode (SCE) by applying the following equation: E(SCE) = E(Ag/AgNO3) + 298 mV.

2.2. The Catalytic Degradation

The photocatalytic degradation of the malachite green (MG) dye in the presence of compounds 14 was realized in air. A 5 mg quantity of H2TClPP (1) or [CoII(TClPP)] (2) or [CoIII(TClPP)Cl] (3) or [CoIII(TClPP)Cl(NTC)]CH2Cl2 (4) (0.0066, 0.0061, 0.0046 and 0.0046 mmol, respectively) was added to 50 mL of an aqueous solution of MG (20 mg·L−1, 0.055 mmol). The pH of the solution was 7. The mixture was first stirred in the darkness for 30 min of illumination to assure that the adsorption–desorption equilibrium was formed. A 300 W Xe lamp was used as the light source with an optical filter (λ > 400 nm) held at about 5 cm from the liquid surface for 30 min. At selected time intervals, an adequate quantity of suspension was centrifuged and filtered using a membrane filter to eliminate solid particles and recover the filtrate for further processing. The degradation efficiency (R%) was measured according to the relation R % = C o C t C o × 100 , where Ct and Co are the concentrations at time t and starting concentration, respectively.

2.3. Synthetic Procedures

All reagents and solvents were obtained commercially and used without further purifications.
In Scheme 1 are illustrated the structures of our four porphyrinic compounds 14:

2.3.1. Synthesis of H2TClPP (1)

One gram of 4-chlorobenzaldheyde (0.007 mol) was dissolved using a magnetic stirrer in 100 mL of propionic acid and heated to 120 °C while maintaining stirring. Pyrrole (0.5 mL, 0.007 mol) was then added dropwise to the yellowish solution and the mixture was kept at 120 °C for an additional hour. The resulting solution was cooled, and the tar mixture was filtered to obtain a blue-black precipitate that was decanted with water (5 × 50 mL) and n-hexane (5 × 50 mL) and finally dried under vacuum. The crude solid was dissolved in 50 mL of chloroform and purified by silica gel column chromatography using the chloroform as eluent to afford the meso-tetra(para-chlorophenyl)porphyrin (H2TClPP) (1) with a yield of 21%.
Anal (%) calcd for C44H26N4Cl4 (750.09): C, 70.46; H, 3.49; N, 7.47; found: C, 70.23; H, 3.65; N, 7.62; UV-vis [CH2Cl2]: λmax (ε.10−3 M−1.cm−1): 421(335), 522(85), 557(56), 599(29), 651(36); 1H NMR (400 MHz, CDCl3) δ(ppm), 8.89 (s, 8 H β-pyrrole), 8.08 (s, 8H ArH) 7.74(s,8H, ArH), −2.75 (NH); FTIR (solid, ν ¯ , cm−1); 3314 ν[(NH) porphyrin], 2928 ν[(CH) porphyrin], 1470 ν[(C=N),(C=C) porphyrin], 963 [δ(CCH) porphyrin].

2.3.2. Synthesis of [CoII(TClPP)] (2)

A 200 mL two-necked flask was charged with 120 mL of dimethylformamide (DMF) and the solvent was degassed with argon for 10 min.H2TClPP porphyrin (1) (200 mg, 1.0 eq) was added to the main reaction vessel and the mixture was heated to 160 °C under magnetic stirring. CoCl2.6H2O (69 mg, 2 eq) was added to the mixture under argon and the mixture was left at 160 °C and magnetic stirring for a further 2 h until the TLC analysis (SiO2, CHCl3 eluent) revealed no trace of the free-base porphyrin. The resulting mixture was cooled to 0 °C. The red-orange color precipitate was collected by filtration, washed with water then with n-hexane and air-dried. The obtained pure [CoII(TClPP)] complex (2) was obtained with a 91% yield (195 mg).
Anal (%) calcd for C44H24N4Cl4Co (809.45): C, 65.29; H, 2.99; N, 6.92; found: C, 65.17; H, 2.79; N, 7.10;UV-vis [CH2Cl2]: λmax (ε.10−3 M−1.cm−1) 414(340), 532(56); 1H NMR (400 MHz, CDCl3) δ(ppm), 15.75 (s, 8 H β-pyrrole), 12.93 (s, 8H ArH), 9.9 (s,8H, ArH); FTIR (solid, ν ¯ , cm−1); 2966 ν[(CH) porphyrin], 1484 ν[(C=N), (C=C) porphyrin], 1007 [δ(CCH) porphyrin].

2.3.3. Synthesis of [CoIII(TClPP)Cl] (3)

[CoII(TClPP)] (2) (100 mg, 0.12 mmol) was dissolved in 10 mL of dichloromethane and 2 mL of HCl (27%) was added and the color changed from red orange to dark green. The mixture was stirred at 60 °C for 3 h. Then, the obtained precipitate was collected by filtration, washed with 25 mL of cold methanol, and dried under vacuum. The resulting cobalt(III) [CoIII(TClPP)Cl] complex (3) was obtained with a yield of 78%.
Anal (%) calcd for C44H24N4Cl5Co (844.90): C, 62.55; H, 2.86; N, 6.63; found: C, 62.72; H, 2.98; N, 6.79; UV-vis [CH2Cl2]: λmax (ε.10−3 M−1.cm−1) λmax: 442(296), 557(50), 596(36); 1H NMR (400 MHz, CDCl3): δ = 8.95 (s, 8 H β-pyrrole); 7.95(s, 8H ArH); 7.78(s, 8H ArH); FTIR (solid, ν ¯ , cm−1): 2900 ν[(CH) porphyrin], 1504 ν[(C=N), (C=C) porphyrin], 1060 [δ(CCH) porphyrin].

2.3.4. Synthesis of [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4)

The coordination compound 3 (20 mg, 0.02 (mmol) was mixed with nicotinoyl chloride hydrochloride C6H4ClNO·HCl (80 mg, 0.76 mmol) in 50 mL of dichloromethane at room temperature for 2 h. Blue-green crystals of [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4) (NTC = nicotinoyl chloride) were obtained by slow diffusion of n-hexane into the dichloromethane solution (Yield ~85%).
Anal (%) calcd for C51H30N5Cl6OCo (1071.39): C, 57.17; H, 2.82; N, 6.54; found: C, 57.51; H, 2.69; N, 6.71; UV-vis [CH2Cl2]: λmax (ε.10−3 M−1.cm−1): 455(335), 560(45), 696(59); 1H NMR (400 MHz, CDCl3): δ = 9.08 (s, 8 H β-pyrrole), 8.09 (s, 8H ArH), 7.75 (s, 8H ArH),7.28 (s, 3H ArH-ligand), 6.65 (s, ArH-ligand); FTIR (solid,   ν   ¯ , cm−1), 2950 ν[(CH) porphyrin], 1708 ν[(C=O) Ligand] 1496 ν[(C=N), (C=C) porphyrin], 1006 [δ(CCH) porphyrin].

3. Results and Discussion

3.1. IR and Proton NMR Spectroscopic Data

The IR spectra of the H2TClPP (1) free-base porphyrin and the [CoII(TClPP)] (2) and [CoIII(TClPP)Cl] (3) starting materials are depicted in Figures S1, S6 and S7, respectively.
Complex 4 exhibits an IR spectrum characteristic of a meso-arylporphyrin coordination compound and confirms the presence of the nicotinoyl chloride (NTC) axial ligand (Figure S8). Indeed, the ν(CH) stretching frequency values of the TClPP porphyrinate and the NTC axial ligand of 4 are in the 3050–2854 cm−1 range while the ν(C=C) and ν(C=N) stretching frequency values are 1496 cm−1. The δ(CCH) bending frequency value of 4 is shown at 1007 cm−1. The coordination of the NTC to the cobalt is confirmed by the absorption bands at 1708 cm−1 corresponding to the ν(C=O) stretching frequency. This value indicates that this axial ligand is coordinated to the central Co(III) metal through the pyridyl group and not the carbonyl group of the chloride acid.
1H NMR spectroscopy is a successful method to check whether a cobalt metalloporphyrin is a paramagnetic cobalt(II) complex or a diamagnetic cobalt(III) coordination complex with 3d7 and 3d6 fundamental state electronic configurations of the Co(II) and Co(III) cations, respectively. Cobalt(II) meso-arylporphyrin derivatives exhibit downfield-shifted 1H NMR spectra with β-pyrrole proton chemical shift values ~16 ppm and phenyl proton values in the range 13–9.8 ppm. On the other hand, for cobalt(III) meso-arylporphyrin complexes the β-pyrrole protons and of the phenyl ring are slightly downfield-shifted with respect to those of the related free-base porphyrins, with chemical shift values ~8.9 ppm and in the range 7.65–8.80 ppm, respectively. The NMR spectrum of complex 4 is depicted in Figure S9 while the chemical shift values of the phenyl and the protons β-pyrrole of 14 and several related cobalt(II) and cobalt(III) metalloporphyrins are listed in Table 1. The β-pyrrole protons of 4 resonate at 9.08 ppm while the chemical shift values of the aryl protons of this complex are in the range 8.09–7.75 ppm, which confirms that our Co-TClPP-NTC derivative is a paramagnetic cobalt(III) metalloporphyrin. The protons of the NTC axial ligand resonate between 7.5 and 6.5 ppm.

3.2. Photophysical Properties

The UV-vis spectra of our four synthetic porphyrinic species (14) are depicted in Figure 1 while in Table 2 is reported the UV-vis data of these compounds and several similar porphyrin species.
The λmax value of the Soret band of the free-base porphyrin H2TClPP is 421 nm while those of the Q bands are 522, 557, 599 and 651 nm which are characteristic for free-base meso-arylporphyrins. The insertion of cobalt leads to a reduction in the number bands Q from four to two and to a shift of the Soret band towards the blue at 414 nm. The absorption spectrum of [CoIII(TClPP)Cl] (3) shows a red-shift Soret band at 442 nm, which is an indication of the oxidation of cobalt(II) to cobalt(III) (Table 2). In addition, the coordination of the nicotinoyl chloride (NTC) to the Co(III) center metal of (3) leads to an important red shift of the Soret band (455 nm) and the Q bands (560 and 696 nm). For this Co(III) pentacoordinated metalloporphyrin (4), the significant redshift of the Soret and Q bands explain the green color of this Co(III) metalloporphyrin, which is due to the important deformations of the porphyrin core (see the crystallographic section) reported by Weiss et al. [36].
The values of the optical gap (Eg) of 14 were calculated by applying the Tauc relationship α h 2 = A [ h E g ], where A is a constant parameter depending on the transition probability, is the incident photon energy and α is the optical absorption coefficient extracted from absorbance data [35]. The intercept from the solid lines and the x-axis allows the determination of Eg values (Figure S10) of 14 which are 1.88, 2.01, 1.96 and 1.73 eV, respectively (Figure S10, Table 2). The free-base porphyrin H2TClPP (1) exhibits a Eg value which is lower than those of [CoII(Telp)] (2) and [CoIII(TClPP)Cl] (3) which could be justify by the fact that the non-metalated porphyrin presents a higher flexibility of the porphyrin core leading to the destabilization of the HOMO-LUMO orbitals and therefore lowering the energy of the HOMO-LUMO orbitals. In the case of [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4), the Eg energy (1.73 eV) is lower than that of the H2TClPP free-base porphyrin (1.88 eV), which is related to the very important distortion of the porphyrin macrocycle of this complex (see crystallographic section).
Porphyrins compounds usually present two emission transition types: (i) the S1 → S0 transition from the S1 first excited state to the S0 ground state corresponding to the Q bands [O(0,0 and Q(0,1)] and (ii) the S2→S0 transition from the second excited state S2 to the ground state S0 corresponding to the Soret band. Given that the S2→ S0 emission transition is very weak compared to the S1→S0 transition for porphyrin species, the former fluorescence is usually not considered for porphyrins and metalloporphyrins.
The fluorescence spectra of our four porphyrinic species (14) are presented in Figure 2 and the fluorescence parameters of these compounds and some free-base meso-arylporphyrins and Co(II) and Co(III) related metalloporphyrins are reported in Table 3.
As shown in Figure 2, the H2TClPP porphyrin (1) presents Q(0,0) and Q(0,1) emission band values at 650 and 714 nm, respectively. For the cobaltous [CoII(TClPP)] (2) complex, the O(0,0) and Q(0,1) bands are blue-shifted with λmax values of 641 and 709 nm, respectively. Our two cobalt(III) complexes (34) show blue-shifted Q(0,0) and Q(0,1) bands at 637 and 699 nm, respectively. Generally, the fluorescence quantum yields (φf) of free-base porphyrins are greater than those of the corresponding metalated, which is explained by the electron transfer from the donor part of the metalloporphyrin to its acceptor part. The donor part of a porphyrin complex is the porphyrin core, and the acceptor part is the central metal. Indeed, the φf value of 1, which is 0.089, is higher than that of 2 with a value of 0.035. The fluorescence lifetime (τf) values have usually the same trend than the fluorescence quantum yields. Thus, the τf of 12 are 7.40 and 6.40 ns, respectively. For the two Co(III) complexes [CoIII(TClPP)Cl] (3) and [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4), we noticed: (i) compounds 34 present very close fluorescence lifetime values (~2.40 ns), (ii) for 3, the φf value is as expected smaller than that of the H2TClPP with a value of 0.051 and (iii) compound 4 exhibits high φf value (0.065). An explanation of this behavior could be the very important deformation of the porphyrin core, which has an important effect on the acceptor–donor characteristics of the [CoIII(TClPP)Cl(NTC)] molecule.

3.3. X-ray Molecular Structure of Complex 4

The molecular structure of our Co(III)-TClPP-Cl-NTC derivative (4) was determined by X-ray diffraction. Single crystals were obtained by slow diffusion of n-hexane through a dichloromethane solution containing complex 4. The crystallographic data and the refinement of the structure of this species are presented in Table S1 while a selection of distances and angles of the same species are given in Table S2. Complex 4 crystallizes in the monoclinic crystal system with P21/c space group. One [CoIII(TMPP)Cl(NTC)] molecule and one disordered dichloromethane solvent molecules are the only constituents of the asymmetric unit of 4. The ORTEP drawing of this Co(III) metalloporphyrin is illustrated in Figure 3. For this Co(III)-NTC porphyrin derivative, the cobalt(III) central ion is coordinated by the four pyrrole N atoms of the porphyrin. The chloride and the nicotinoyl chloride (NTC) axial ligands occupy the two apical sites of the distorted square-bipyramidal coordination polyhedra.
Senge et al. [39] reported that the macrocycle of the porphyrin presents four ideal deformations: (i) the doming deformation (dom) due to the displacement of the metal atom out of the 24-atom mean plane of the porphyrin core and the displacement of the nitrogen atoms to the axial ligand, (ii) the ruffling distortion (ruff) is characterized by the high values of the displacement of the meso-carbon atoms over and below the porphyrin mean plane (Figure 4a,b), (iii) the saddle deformation (sad) originates from the displacement of the pyrrole rings alternatively above and below the mean porphyrin core so that the pyrrole nitrogen atoms are out of the mean plane, and (iv) the waving distortions (wav), for which the four fragments “β-carbon-α-carbon-meso-carbon-α-carbon-β-carbon” are alternatively below and above the mean plane of the porphyrin macrocycle.
The most structural important feature of our new cobalt(III) chloride-nicotinoyl chloride derivative (4) is the very important ruffling and waving deformations of the porphyrin core (Figure 4c). Thus, as shown in Figure 5, the displacements of the meso-carbons from the mean plane of the 24-atom porphyrin core are −67, 63, −63 and 64.10−2 Å, indicating a very important ruffling deformation. The four fragments “β-carbon-α-carbon-meso-carbon-α-carbon-β-carbon” also present very high displacements vis-à-vis the mean plane of the 24-atom porphyrin core with values “−26 −35 −67 −39 −30”, “+22 +37 +63 +36 +27”, “−23 −35 −63 −36 −32” and “+16 +36 +64 +42 +27” 10−2 Å, which confirms the significant waving deformation of the porphyrin core. It is noteworthy that the deformation of the porphyrin core and especially the ruffling distortion and the variation in the equatorial mean distance between the center metal and the nitrogen atoms of the porphyrin core (M__Np, M = center ion) are related. Thus, the greater the ruffling deformation is, the smaller the Co–Np distance is and vice versa [39].
Inspection of Table 4 shows that (i) cobalt(III) metalloporphyrins with significant ruffling and waving deformations present very short Co–Np distances, while those with small deformations of the porphyrin core show longer Co–Np distance values as just indicated above, and (ii) our Co(III)-Cl-NTC derivative (4) presents a very short Co–Np bond distance of 1.947(3) Å, which is in accordance with the fact that this species exhibits very important ruffling and waving deformations. Consequently, the Soret and Q bands of the spectrum of 4 and the Q bands of the fluorescence spectrum are very red-shifted.
The important ruffling and waving deformations of the porphyrin core of 4 are most probably due to coordination of the sterically hindered nicotinoyl chloride axial ligand to the Co(III) center ion. Thus, in order to minimize the interaction of this ligand with the phenyls or the TClPP porphyrinate, the porphyrin core is twisted severally so that the hydrogen atoms of the pyridyl group of the nicotinoyl chloride axial are as far as possible from the hydrogens of the phenyl closest to the porphyrin, leading to the formation of a “ligand binding pocket” in which the nicotinoyl chloride axial is located (Figure 4c).
There are obviously no classical hydrogen bonds in the structure of 4 for which the crystal stability is described as follows (Figures S1–S4): (i) the chlorine Cl2 in para position of a phenyl group of one [CoIII(TClPP)Cl(NTC)] molecule and the C8 atom of one pyrrole ring of a neighboring TClPP porphyrinate of a [CoIII(TClPP)Cl(NTC)] complex are hydrogen bonded with a C8__H8Cl2 distance of 3.398 (4) Å, (ii) the carbon atoms C31 and C32 of one phenyl ring of one [CoIII(TClPP)Cl(NTC)] complex are H bonded to the centroid Cg4 of the N4/C16-C19 pyrrole ring and the Cl5 chloride axial ligand of the same [CoIII(TClPP)Cl(NTC)] closest neighbor, respectively (Table S3). The C31__H31Cg4 and C32__H32Cl5 distance values are 3.503(3) and 3.374 (3) Å, respectively, (iii) the carbon C47 of the phenyl ring of the nicotinoyl chloride (NTC) axial ligand is weakly linked to the centroid Cg3 of the N3/C11-C14 of a nearby [CoIII(TClPP)Cl(NTC)] molecule with a C47__H47Cg3 distance of 3.458(5) Å and (iv) the centroid Cg1 of the N1/C2-C4 of the later molecule is weakly H bonded to the carbons C51A and C51B of the disordered dichloromethane solvent molecule with a C50A__H51BCg1 and C50B__H51CCg1 distance values of 3.675(6) and 3.61(5) Å.
We noticed the presence of small voids 1.2 Å and a grid of 0.7 Å perpendicular to the [010] direction (Figure S4). These voids correspond to 2% of the cell volume with a total volume of 95.12 Å3 per cell.

3.4. Hirshfeld Surface Analysis

In order to further understand the intermolecular interactions in the crystal of compound 4, Hirshfeld surface (HS) analysis was carried out by using Crystal Explorer 17.5 [47].
The white surfaces in the HS plotted over dnorm, indicates contacts with distances equal to the sum of van der Waals radii while the red and blue colors indicate distances shorter or longer than the van der Waals radii, respectively [48]. As shown in Figure 6a, the red spots in the Hirshfeld surface presented by the dnorm surface correspond to the C8__H8Cl2, C32__H32Cl5, C31__H31Cg4, C47__H47Cg3, C51A__H51BCg1 and C51B__H51CCg1. As already indicated by the PLATON calculations [49], the most abundant contributor of the total Hirshfeld surface around complex 4 is the H…Cl/Cl…H interaction which contributed by 33.2%. The other noticeable contributors are: H…H (29.7%), C…Cl/Cl…C (7.4%), Cl-Cl (5.1%) H…O/O…H (3.6%) (Figure 7).
Both the curvature and the shape indices can also be utilized to make identifications of typical stacking modes and the manners in which closely spaced molecules engage with each other. The shape index for 4 shows a red concave area on the surface around the acceptor atom and a blue area around the donor H-atom (Figure 6b) [50,51]. The curvedness is a function of the root mean square curvature of the surface and the maps of curvedness on the HS for complex 4 indicate no plain surface patches, indicating that there is no stacking interaction between the molecules (Figure 6c).

3.5. Cyclic Voltammetry Investigation on [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4)

The electrochemical properties of 4 were assessed by cyclic voltammetry with tetra-n-butylammonium hexafluorophosphate (TBAPF6) used as supporting electrolyte (0.1 M) in dichloromethane under an argon atmosphere. The voltammogram of this cobalt(III) metalloporphyrin is illustrated by Figure 8.
The reduction in [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4) has uncoupled cathodic and anodic peaks that are tagged 1a (reduction) and 1b (oxidation) (Scheme 2) and were spectro-electrochemically analyzed by Kadish et al. [52] for the [CoIII(TPP)Cl] (TPP = meso-tetraphenylporphyrinate) coordination compound as inlaying a Co(III)/Co(II) reduction. In the case of [CoIII(TPP)Cl], the reduction (1a) occurs at Ecp = −0.20 V (Ecp = cathodic potential) and the reoxidation of the generated Co(II) complex (peak 1b) occurs at Eap = 0.57 V (Eap = anodic potential) (Figure 8, Table 5). For complex 4, the Ecp (reduction 1a) and Eap (oxidation 1b) values are −0.37 V and −0.75 V, respectively (Figure 9 Table 5). The large difference in the Ecp and Eap for the Co(III)-Cl-TPP and the Co(III)-Cl-NTC-TClPP could be caused by the important deformation of the porphyrin core in the case of complex 4 leading to the narrowing the energy levels of the HOMO and LUMO orbitals of this species. The uncoupled Co(III)/Co(II) reactions of cobalt(III) porphyrin complexes were explained [52] by the “box mechanism” shown in Scheme 2. In this mechanism, the reduction and reoxidation of the [CoIII(TPP)Cl] occur via two distinct reversible electron-transfer steps, each a chemical reaction:
As reported by Kadish et al. [52], the second reduction of [CoIII(TPP)Cl] occurs at Ecp = −0.90 V [process (3) in Figure 8] involving Co(II)/Co(I) reduction and indicates the characteristics of a couple of chemical reactions (Scheme 2). For our Co(III)-Cl-NTC-TClPP derivative, the Ecp value for the second reduction occurs at −1.29 V, which is shifted toward the negative potential, due probably to the significant deformation of porphyrin core of 4. The third reduction of [CoIII(TPPCl] reported by Kadish et al. involves the ring reduction of the electrogenerated [CoI(TPP)(CH2Cl] with Ecp value of −1.42 V. For complex 4, a similar third reduction occurs at Ecp = −1.65 V. Based on the scheme proposed by Kadish et al. for the second reduction of [CoIII(TPP)Cl] (Scheme 3), is proposed to describe the second reduction of our Co(III) porphyrinic complex 4:
The anodic part of the cyclic voltammogram of 4 contains two reversible one-electron oxidations assigned to the oxidation of the porphyrin ring where the E1/2 values are 1.06 and 1.43 V, respectively. These values are shifted to more positive potential than those of [CoIII(TPP)Cl] (0.90 and 1.15 V, respectively) which also could be attributed to the significant deformation of the porphyrin macrocycle of 4.
The photodegradation reactions of the malachite green (MG) dye using 14 as catalysts were monitored by UV spectroscopy. The λmax of the absorption band of the MG at 618 nm was utilized to make an approximate assessment of the decolorization rate of the organic dye. The variation of λmax of the absorption of MG dye upon radiation time, using 14 is reported in Figure S11. As shown in this figure, the degradation of MG dye did not occur without adding compounds 14.
The effective degradation of MG dye (Ct/Co versus time curve) using 14 as photocatalysts indicates that [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4) shows the higher degradation efficiency after 60 min of irradiation with a yield value of 95% (Figure 9). For compounds 13, the degradation yields are 90%, 80% and 84%, respectively.
The photocatalytic discoloration of a dye is believed to take place according to the following mechanism. When a catalyst is exposed to UV radiation, electrons are promoted from the valence band to the conduction band. As a result, an electron–hole pair is produced [53], and e c b and h v b + are the electrons in the conduction band and the electron vacancy (holes) in the valence band, respectively. Both these entities can migrate to the catalyst surface, where they can enter a redox reaction with other species present on the surface. In most cases, h v b + can react easily with surface bound H2O to produce OH radicals, whereas e c b can react with O2 to produce superoxide radical anions of oxygen [54] (Scheme 4). This reaction prevents the combination of the electron and the hole that are produced in the first step. The OH and O2 produced in the above manner can then react with the dye to form other species and is thus responsible for the discoloration of the dye.
The reusability of the photocatalysts was considered. Thus, our four porphyrinic derivatives (14) used as catalysts were separated by filtration, washed with distilled water after each run, then dried and further subjected to subsequent runs under the same conditions. Figure S12 indicates that the regeneration process could be repeated for 4 cycles, without appreciable activity loss. The regenerated catalysts were also characterized by FTIR analyses after each cycle and no change was observed.
The excellent photocatalytic degradation yields, and the cyclic stable performance testified that 14 could be candidates for other photocatalysis reactions.

4. Photovoltaic Performance of DSSCs

In the present work, the devices consisted of the porphyrin complexes 3 and 4 located between two electrodes, the “ITO-coated glass” and aluminum. The current vs voltage plots of the ITO/Pm/Al systems (Pm = [CoIII(TClPP)Cl] or [CoIII(TClPP)Cl(NTC)]·CH2Cl2]) is depicted in Figure 10. The displayed AFM image clearly reveals the formation of a homogeneous film with a great layer structure (1.89 nm).
This investigation makes it possible to determine the transport properties in the organic materials. Current–voltage plots (J-V) of compounds 3 and 4 were recorded in the dark and at room temperature.
As shown by Figure 10, the (J-V) plots are nonlinear, while the ITO/Pm/Al devices exhibit a clear moderate rectifying performance due to the injection of ITO charges for complexes 3 and 4, leading to an organic device that may be a photovoltaic device [55,56]. The threshold voltages of complexes 3 and 4 are 0.592 and 0.652 V, respectively.
By studying the current-voltage (J-V) curves, we can observe the existence of two different regimes, which depend on the applied voltage.
At low voltages, the first regime presents a symmetry characteristic explained by the localized state with defects causing localized gap states. The second regime exhibits an asymmetric characteristic. This is related to the injection process of the electron and hole barriers due to the difference in the work functions of the two electrodes.
Better analysis of (J-V) characteristics based on semi-logarithmic representation is presented in Figure 11. Two regions have been noticed on these curves, the first region is linear; the current being limited by the resistance via the shunt resistance Rsh [57,58,59]. In the second region, the current starts to saturate due to the series resistance Rs.
The Js values of the developed devices were obtained from the experimental (J-V) data, and Φ b was calculated following the following equation (Equation (1)):
J s = SS *   exp q Φ b KT
where S is the constant area and S* is the Richardson constant, k the Boltzmann constant (1.38 × 10−23 J K−1), T the temperature and q the electronic charge.
Comparing the parameters of the two complexes, we concluded that complex 3 has a minimum barrier height value compared to complex 4 (Table 6). This can be explained by the presence of the ligand, which affects the charge transport mechanism.
Following the insertion of the carrier charges, their transport across the active surface to the opposite electrodes is defined by the conduction characteristics. The J–V characteristics in log-log plot of the compounds 3 and 4 in dark conditions are shown in Figure 12. The investigation of these plots suggests that the dependency of the current on the applied voltage seems to follow the power law: J = Vm. The current density is defined by the space charge of the carriers injected by the electrodes. By observing the plot, we have noticed the existence of two regimes. These regimes depend on the value of the slope m.
For the first phase, the current depends on the voltage (m = 1), which defines an ohmic region, due to the presence of a quantity of interface barrier preventing charge injection.
The current density is calculated by the following equation (Equation (2)):
J Ω = qp 0 μ V d
where μ is the charge carrier mobility, q is the electron charge and d is the film thickness.
In the second phase, the voltage increases, and the current depends on the voltage (m = 2), corresponding to the space charge limited conduction (SCLC) mode [60,61,62,63]. In this phase, the current density is given by the following formula (Equation (3)):
J SCLC = 9 8 ε μ eff V 2 d 3
where μ eff is the effective carrier mobility and ε is the permittivity of the material.

5. Conclusions

In summary, the synthesis and spectroscopic characterization of the free-base porphyrin H2TClPP (1), the [CoII(TClPP)] (2)the [CoIII(TClPP)Cl] (3) starting materials and the [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4) complex (NTC = nicotinyl chloride) were performed. The most important feature of the X-ray molecular structure of 4 is the very important ruffling and waving deformations of the porphyrin core. The Hirshfeld surfaces analysis on this later cobalt(III) metalloporphyrin shows that the crystal lattice is principally supported by C__HCl and C__HCg (Cg is the centroid of a pyrrole ring) involving both [CoIII(TClPP)Cl(NTC)] and the dichloroethane solvent molecules. The UV-vis and fluorescence spectra of 4 are red-shifted due to the significant deformation of the porphyrin macrocycle. The electrochemical property of 4 was investigated using cyclic voltammetry, showing that the reduction potentials are shifted to negative values while the oxidation potentials are shifted to positive values compared to [CoIII(TPP)Cl]. This feature is most probably due to the significant deformation of the porphyrin ring of 4. Furthermore, the photodegradation of the malachite green dye, using 14 as catalysts, gives good yields between 80% for 2 and 95% for compound 4. The photovoltaic characteristics of DSSCs based on complex 3 and 4 were measured.
In summary, we successfully prepared a new hexacoordinated cobalt(III) meso-arylporphyrin complex with the chlorido and the nicotinoylchloride axial ligands with the formula [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (TClPP = the meso-tetra(para-chlorophenyl)porphyrinate and NTC is the nicotinoylchloride axial ligand) (4). This Co3+ metalloporphyrin (4) and the H2TCPP (1), [CoII(TClPP)] (2) and [CoIII(TClPP)Cl] (3) materials were characterized by UV-vis, IR, 1H NMR, fluorescence, mass spectrometry. Complex 4 was also characterized by single-crystal X-ray molecular structure and cyclic voltammetry. The most important feather of 4 is the very important deformation of its porphyrin core. This has consequences on several proprieties of complex 4: (i) a very red shift of the UV-vis spectrum, (ii) a high φf value (0.065) where the very important deformation of the porphyrin core effects the acceptor–donor characteristics of this Co(III) complex, and (iii) the E1/2 values of the two reversible oxidation waves of the porphyrin ring are shifted to the positive potentials. All four porphyrinic compounds (14) were tested as catalysts in the photochemical degradation of malachite green (MG) dye, where complex 4 gave the best degradation yield (95%). Furthermore, the electrical properties of complexes 34 obtained based on their current voltage curves show that the nature of the axial ligand affects the performance of the cells including these two cobalt(III) metalloporphyrins.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27248866/s1, Figure S1: Representation de la structure du complex [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4) showing the disordered NTC axial ligand and the disordered dichloromethane solvent molecule; Figure S2: Schematic representation showing the weak C8–H8··Cl2 intermolecular interaction in the crystal lattice if 4; Figure S3: Schematic representation showing the weak C–H··Cl and C –H…Cg intermolecular interactions in the crystal lattice if 4; Figure S4: Packing diagram of compound 4 showing voids calculated for a ball radius of 1.2 Å and a grid of 0.7Å; Figure S5: IR spectrum of the free base porphyrin (H2TClPP) (1); Figure S6: IR spectra of the complex [CoII(TClPP)] (2); Figure S7: IR spectra of the complex [CoIII(TClPP)Cl] (3); Figure S8: IR spectrum (solid state) of [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4); Figure S9: 1H NMR spectra of 4 recorder in CDCl3 with a concentration ~ 10−3 M; Figure S10: Curves of (αhυ)2 against photon energy E of compounds 14; Figure S11: Variation of the λmax values of the absorption bands of Malachite green (MG) dye in the presence of 14 (5 mg). The concentration of MG is 20 mg·L−1 and pH = 7; Figure S12: Circulatory experiments of photo catalytic degradation of dye using four compounds 14 as photo catalysts; Table S1: Crystal data and structural refinement for [CoIII(TClPP)(NTC)]·CH2Cl2 (4); Table S2: Selected bond lengths (Å) and angles (°) of 4; Table S3: Selected intermolecular interactions for compound 4.

Author Contributions

Conceptualization, M.G., J.B. and Y.O.A.-G.; methodology, F.M. and I.T.-T.; software, I.T.-T.; validation, H.N.; investigation, S.N., M.G. and J.B.; resources, F.L.; data curation, Y.O.A.-G.; writing—original draft preparation, S.N. and J.B.; writing—review and editing, H.N.; supervision, F.L. and H.N.; project administration, S.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to the deputyship for Research and Innovation, Ministry of Education in Saudi Arabia for funding this research work through project IFP-2020-05.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Maki, A.H.; Edelstein, N.; Davison, A.; Holm, R.H. Electron Paramagnetic Resonance Studies of the Electronic Structures of Bis(maleonitriledithiolato)copper(II), -nickel(III), -cobalt(II), and -rhodium(II) Complexes. J. Am. Chem. Soc. 1964, 86, 4580–4587. [Google Scholar] [CrossRef]
  2. Walker, F.A. An Electron Spin Resonance Study of Coordination to the Fifth and Sixth Positions of α,ß,γ,δ-Tetra (p-methoxyphenyl)porphinato cobalt(II). J. Am. Chem. Soc. 1970, 92, 4235–4244. [Google Scholar] [CrossRef]
  3. Lexa, D.; Savéant, J.; Soufflet, J. Chemical catalysis of the electrochemical reduction of alkyl halides Comparison between cobalt-tetraphenyl porphin and vitamin B12 derivatives. J. Electroanal. Chem. 1979, 100, 159–172. [Google Scholar] [CrossRef]
  4. Lyaskovskyy, V.; Suarez, A.I.O.; Lu, H.; Jiang, H.; Zhang, X.P.; de Bruin, B. Mechanism of Cobalt(II) Porphyrin-Catalyzed C-H Amination withOrganic Azides: Radical Nature and H-Atom Abstraction Ability of the Key Cobalt(III)-Nitrene Intermediates. J. Am. Chem. Soc. 2011, 133, 12264–12273. [Google Scholar] [CrossRef] [PubMed]
  5. Kim, J.; Lim, S.-H.; Yoon, Y.; Thangadurai, T.D.; Yoon, S. A fluorescent ammonia sensor based on a porphyrin cobalt(II)–dansyl complex. Tetrahedron Lett. 2011, 52, 2645–2648. [Google Scholar] [CrossRef]
  6. Karimipour, G.; Kowkabi, S.; Naghiha, A. New Amino porphyrins Bearing Urea Derivative Substituents: Synthesis, Characterization, Antibacterial and Antifungal Activity. Braz. Arch. Biol. Technol. 2015, 58, 431–442. [Google Scholar] [CrossRef] [Green Version]
  7. Beyene, B.B.; Wassie, G.A. Antibacterial activity of Cu(II) and Co(II) porphyrins: Role of ligand modification. BMC Chem. 2020, 14, 51. [Google Scholar] [CrossRef]
  8. Crawley, M.R.; Zhang, D.; Oldacre, A.N.; Beavers, C.M.; Friedman, A.E.; Cook, T.R. Tuning the Reactivity of Cofacial Porphyrin Prisms for Oxygen Reduction Using Modular Building Blocks. J. Am. Chem. Soc. 2021, 143, 1098–1106. [Google Scholar] [CrossRef] [PubMed]
  9. Feng, L.; Wang, K.-Y.; Joseph, E.; Zhou, H.-C. Catalytic Porphyrin Framework Compounds. Trends Chem. 2020, 2, 555–568. [Google Scholar] [CrossRef]
  10. Cheng, N.; Kemna, C.; Goubert-Renaudin, S.; Wieckowski, A. Reduction Reaction by Porphyrin-Based Catalysts for Fuel Cells. Electrocatalysis 2012, 3, 238–251. [Google Scholar] [CrossRef]
  11. Zhai, Z.; Liu, Q.; Zheng, R.; Qiu, C.; Qin, J.; Li, J.; Xie, Y.; Wang, A.; Huang, J.; Song, Y. Controlled pyrolysis of ionically self-assembled metalloporphyrins on carbon as cathodic electrocatalysts of polymer electrolyte membrane fuel cells. Int. J. Hydrogen Energy 2021, 46, 11041–11050. [Google Scholar] [CrossRef]
  12. Alvarez, I.B.; Wu, Y.; Sanchez, J.; Ge, Y.; Ramos-Garcés, M.V.; Chu, T.; Jaramillo, T.F.; Colón, J.L.; Villagrán, D. Cobalt porphyrin intercalation into zirconium phosphate layers for electrochemical water oxidation. Sustain. Energy Fuels 2021, 5, 430–437. [Google Scholar] [CrossRef]
  13. Wirojsaengthong, S.; Aryuwananon, D.; Aeungmaitrepirom, W.; Pulpoka, B.; Tuntulani, T. A colorimetric paper-based optode sensor for highly sensitive and selective determination of thiocyanate in urine sample using cobalt porphyrin derivative. Talanta 2021, 231, 122371. [Google Scholar] [CrossRef]
  14. Fan, Z.; Zhao, B.; Wu, S.; Wang, H.; Cao, T.; Zhu, T.; Zhang, X.; Liu, L.; Tong, Z. Construction of cobalt porphyrin/tantalum molybdatenanocomposite for simultaneous electrochemical detection of ascorbic acid and dopamine. J. Mater. Sci. Res. 2021, 36, 916–924. [Google Scholar] [CrossRef]
  15. Benkhaya, S.; El Harfi, S.; El Harfi, A. Classifications, properties and applications of textile dyes: A review. Appl. J. Environ. Eng. Sci. 2017, 3, 311–320. [Google Scholar]
  16. Yamjala, K.; Nainar, M.S.; Ramisetti, N.R. Methods for the analysis of azo dyes employed in food industry—A review. Food Chem. 2016, 192, 813–824. [Google Scholar] [CrossRef]
  17. Ismail, M.; Akhtar, K.; Khan, M.I.; Kamal, T.; Khan, M.A.; MAsiri, A.; Seo, J.; Khan, S.B. Pollution, Toxicity and Carcinogenicity of Organic Dyes and their Catalytic Bio-Remediation. Curr. Pharm. Des. 2019, 25, 3645–3663. [Google Scholar] [CrossRef] [PubMed]
  18. Birhanlı, A.; Ozmen, M. Evaluation of the Toxicity and Teratogenity of Six Commercial Textile Dyes Using the Frog Embryo Teratogenesis Assay–Xenopus. Drug Chem. Toxicol. 2005, 28, 51–65. [Google Scholar] [PubMed]
  19. Feng, M.; Wu, L.; Wang, X.; Wang, J.; Wang, D.; Li, C. A strategy of designed anionic metal–organic framework adsorbent based on reticular chemistry for rapid selective capture of carcinogenic dyes. Appl. Organomet. Chem. 2022, 35, e6546. [Google Scholar] [CrossRef]
  20. Chen, H.; Liu, P.; Liu, J.; Feng, X.; Zhou, S. Mechanochemical in-situ incorporation of Ni on MgO/MgH2 surface for the selective O-/C-terminal catalytic hydrogenation of CO2 to CH4. J. Catal. 2021, 394, 397–405. [Google Scholar] [CrossRef]
  21. Li, W.; Zhang, H.; Zhang, K.; Hu, W.; Cheng, Z.; Chen, H.; Feng, X.; Peng, T.; Kou, Z. Monodispersed ruthenium nanoparticles interfacially bonded with defective nitrogen-and-phosphorus-doped carbon nanosheets enable pH-universal hydrogen evolution reaction. Appl. Catal. B Environ. 2022, 306, 121095. [Google Scholar] [CrossRef]
  22. Zhang, H.; Li, W.; Feng, X.; Zhu, L.; Fang, Q.; Li, S.; Wang, L.; Li, Z.; Kou, Z. A chainmail effect of ultrathin N-doped carbon shell on Ni2P nanorod arrays for efficient hydrogen evolution reaction catalysis. Colloid Interface Sci. 2022, 607, 281–289. [Google Scholar] [CrossRef] [PubMed]
  23. Dong, X.; Li, Y.; Li, D.; Liao, D.; Qin, T.; Prakash, O.; Kumar, A.; Liu, J.-Q. A new 3D 8-connected Cd(ii) MOF as a potent photocatalyst for oxytetracycline antibiotic degradation. CrystEngComm 2022, 24, 6933–6943. [Google Scholar] [CrossRef]
  24. Zheng, M.; Chen, J.; Zhang, L.; Cheng, Y.; Lu, C.; Liu, Y.; Singh, A.; Trivedi, M.; Kumar, A.; Liu, J. Metal Organic Framework as an Efficient Adsorbent for Drugs from Wastewater. Mater. Today Commun. 2022, 31, 103514. [Google Scholar] [CrossRef]
  25. La, D.D.; Nguyen, T.A.; Nguyen, X.S.; Truong, T.N.; Ninh, H.D.; Vo, H.T.; Bhosale, S.V.; Chang, S.W.; Rene, E.R.; Nguyen, T.H.; et al. Self-assembly of porphyrin on the surface of a novel composite high performance photocatalyst for the degradation of organic dye from water: Characterization and performance evaluation. J. Environ. Chem. Eng. 2021, 9, 106034. [Google Scholar] [CrossRef]
  26. Li, M.; Zhao, H.; Lu, Z.-Y. Porphyrin-based porous organic polymer, Py-POP, as a multifunctional platform for efficient selective adsorption and photocatalytic degradation of cationic dyes. Microporous Mesoporous Mater. 2020, 292, 109774. [Google Scholar] [CrossRef]
  27. Silvestri, S.; Fajardo, A.R.; Iglesias, B.A. Supported porphyrins for the photocatalytic degradation of organic contaminants in water: A review. Environ. Chem. Lett. 2022, 20, 731–771. [Google Scholar] [CrossRef]
  28. Kechiche, A.; Fradi, T.; Noureddine, O.; Guergueb, M.; Loiseau, F.; Guerineau, V.; Issoui, N.; Lemeune, A.; Nasri, H. Synthesis, characterization and catalytic studies of chromium(III) porphyrin complex with axial cyanate ligands. J. Mol. Struct. 2022, 1250, 131801. [Google Scholar] [CrossRef]
  29. Amiri, N.; Taheur, F.B.; Chevreux, S.; Rodrigues, C.M.; Dorcet, V.; Lemercier, G.; Nasri, H. Syntheses, crystal structures, photo-physical properties, antioxidant and antifungal activities of Mg(II) 4,4′-bipyridine and Mg(II) pyrazine complexes of the 5,10,15,20 tetrakis(4–bromophenyl) porphyrin. Inorg. Chim. Acta 2021, 525, 120466. [Google Scholar] [CrossRef]
  30. Amiri, N.; Guergueb, M.; Al-Fakeh, M.S.; Bourguiba, M.; Nasri, H. A new cobalt(ii) meso-porphyrin: Synthesis, characterization, electric properties and application in the catalytic degradation of dyes. RSC Adv. 2020, 10, 44920–44932. [Google Scholar] [CrossRef]
  31. Mansour, A.; Belghith, Y.; Belkhiria, M.S.; Bujacz, A.; Guérineau, V.; Nasri, H. Synthesis, crystal structures and spectroscopic characterization of Co(II) bis(4,4′-bipyridine) with meso-porphyrins α,β,α,β-tetrakis(o-pivalamidophenyl) porphyrin (α,β,α,β-TpivPP) and tetraphenylporphyrin (TPP). J. Porphyr. Phthalocyanines 2013, 17, 1094–1103. [Google Scholar] [CrossRef]
  32. Guergueb, M.; Nasri, S.; Brahmi, J.; Loiseau, F.; Molton, F.; Roisnel, T.; Guerineau, V.; Turowska-Tyrk, I.; Aouadi, K.; Nasri, H. Effect of the coordination of p-acceptor 4-cyanopyridine ligand on the structural and electronic properties of meso-tetra(para-methoxy)and meso-tetra(para-chlorophenyl) porphyrincobalt(II) coordination compounds. Application in the catalytic degradation of methylene blue dye. RSC Adv. 2020, 10, 6900–6918. [Google Scholar] [PubMed]
  33. Guergueb, M.; Nasri, S.; Brahmi, J.; Al-Ghamdi, Y.O.; Loiseau, F.; Molton, F.; Roisnel, T.; Guerineau, V.; Nasri, H. Spectroscopic characterization, X-ray molecular structures and cyclic voltammetry study of two (piperazine) cobalt(II) meso-arylporphyincomplexes. Application as a catalyst for the degradation of 4-nitrophenol. Polyhedron 2021, 209, 115468. [Google Scholar] [CrossRef]
  34. Sugimoto, H.; Ueda, N.; Mori, M. Preparation and Physicochemical Properties of Tervalent Cobalt Complexes of Porphyrins. Bull. Chem. Soc. Jpn. 1981, 54, 3425–3432. [Google Scholar] [CrossRef] [Green Version]
  35. Albrecht, M.; Maji, P.; Häusl, C.; Monney, A.; Müller-Bunz, H. N-Heterocyclic carbene bonding to cobalt porphyrin complexes. Inorg. Chim. Acta 2012, 380, 90–95. [Google Scholar] [CrossRef]
  36. Weiss, R.; Fischer, J.; Bulach, V.; Schünemann, V.; Gerdan, M.; Trautwein, A.X.; Shelnutt, J.A.; Gros, C.P.; Tabard, A.; Guilard, R. Structure and mixed spin state of the chloro iron(III) complex of 2,3,7,8,12,13,17,18-octaphenyl-5,10,15,20-tetraphenylporphyrin, Fe(dpp)Cl. Inorg. Chim. Acta 2002, 337, 223–232. [Google Scholar] [CrossRef]
  37. Owens, J.W.; Smith, R.; Robinson, R.; Robins, M. Photophysical properties of porphyrins, phthalocyanines, and benzochlorins. Inorg. Chim. Acta 1998, 279, 226–231. [Google Scholar] [CrossRef]
  38. Amiri, N.; Nouir, S.; Hajji, M.; Roisnel, T.; Guerfel, T.; Simonneaux, G.; Nasri, H. Synthesis, structure, photophysical properties and biological activity of a cobalt(II) coordination complex with 4,4′-bipyridine and porphyrin chelating ligands. J. Saudi Chem. Soc. 2019, 23, 781–794. [Google Scholar] [CrossRef]
  39. Kingsbury, C.J.; Senge, M.O. The shape of porphyrins. Coord. Chem. Rev. 2021, 431, 213760. [Google Scholar] [CrossRef]
  40. Chen, L.; Fox, J.J.B.; Yi, G.-B.; Khan, M.A.; Richter-Addo, G.B. Synthesis and molecular structures of N,N-dialkyl-4-nitrosoaniline adducts of formally d6 metalloporphyrins of ruthenium and cobalt. J. Porphyr. Phthalocyanines 2001, 5, 702–707. [Google Scholar] [CrossRef]
  41. Li, J.; Noll, B.; Oliver, A.; Ferraudi, G.; Lappin, A.G.; Scheidt, W.R. Oxygenation of Cobalt Porphyrinates: Coordination or Oxidation? Inorg. Chem. 2010, 49, 2398–2406. [Google Scholar] [CrossRef] [PubMed]
  42. Belghith, Y.; Daran, J.-C.; Nasri, H. Chlorido(pyridine-jN)(5,10,15,20-tetraphenylporphyrinato-j4 N)cobalt(III)chloroform hemisolvate. Acta Cryst. 2012, 68, m1104–m1105. [Google Scholar]
  43. Kaduk, J.A.; Scheidt, W.R. Stereochemistry of low-spin cobalt porphyrins. V. Molecular stereochemistry of nitro-.alpha.,.beta.,.gamma.,.delta.-tetraphenylporphinato(3,5-lutidine)cobalt(III). Inorg. Chem. 1974, 13, 1875–1880. [Google Scholar] [CrossRef]
  44. Goodwin, J.; Bailey, R.; Pennington, W.; Rasberry, R.; Green, T.; Shasho, S.; Yongsavanh, M.; Echevarria, V.; Tiedeken, J.; Brown, C.; et al. Structural and Oxo-Transfer Reactivity Differences of Hexacoordinate and Pentacoordinate (Nitro)(tetraphenylporphinato)cobalt(III) Derivatives. Inorg. Chem. 2001, 40, 4217–4225. [Google Scholar] [CrossRef] [PubMed]
  45. Doppelt, P.; Fischer, J.; Weiss, R. Synthesis and structure of bis(mercapto)cobalt(III) porphyrins. Models for the active site of cytochromes P 450. J. Am. Chem. Soc. 1984, 106, 5188–5193. [Google Scholar] [CrossRef]
  46. Sakurai, T.; Yamamoto, K.; Naito, H.; Nakamoto, N. The Crystal and Molecular Structure of Chloro-α,β,γ,δ-tetraphenylporphinatocobalt(III). Bull. Chem. Soc. Jpn. 1976, 49, 3042–3046. [Google Scholar] [CrossRef]
  47. Turner, M.J.; McKinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Spackman, P.R.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer17; The University of Western Australia: Perth, Australia, 2017. [Google Scholar]
  48. Venkatesan, P.; Thamotharan, S.; Ilangovan, A.; Liang, H.; Sundius, T. Crystal structure, Hirshfeld surfaces and DFT computation of NLO active (2E)-2-(ethoxycarbonyl)-3-[(1-methoxy-1-oxo-3-phenylpropan-2-yl)amino] prop-2-enoic acid. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2016, 153, 625–636. [Google Scholar] [CrossRef] [Green Version]
  49. Spek, A.L. PLATON SQUEEZE: A tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta. Cryst C. 2015, 71, 9–18. [Google Scholar] [CrossRef] [Green Version]
  50. McKinnon, J.J.; Jayatilaka, D.; Spackman, M.A. Towards quantitative analysis of intermolecular interactions with Hirshfeld surfaces. Chem. Commun. 2007, 7, 3814–3816. [Google Scholar] [CrossRef]
  51. Dennington, R.; Keith, T.; Millam, J. GaussView, Version 5; SemiChem Inc.: Shawnee, KS, USA, 2009. [Google Scholar]
  52. Kadish, K.M.; Lin, X.Q.; Han, B.C. Chloride-binding reactions, and electrochemistry of (tetraphenylporphyrinato)cobalt and chloro(tetraphenylporphyrinato)cobalt in dichloromethane. Inorg. Chem. 1987, 26, 4161–4167. [Google Scholar] [CrossRef]
  53. Rauf, M.A.; Bukallah, S.B.; Hammadi, A.; Soliman, A.; Hammadi, F. The effect of operational parameters on the photoinduced decoloration of dyes using a hybrid catalyst V2O5/TiO2. Chem. Eng. J. 2007, 129, 167–172. [Google Scholar] [CrossRef]
  54. Mahmoodi, N.M.; Arami, M.; Limaee, N.Y.; Tabrizi, N.S. Kinetics of heterogeneous photocatalytic degradation of reactive dyes in an immobilized TiO2 photocatalytic reactor. J. Colloid Interface Sci. 2006, 295, 159–164. [Google Scholar] [CrossRef] [PubMed]
  55. Mhamdi, A.; Sweii, F.B.S.; Bouazizi, A. Effect of Thermal Annealing on the Electrical Properties of Inverted Organic Solar Cells Based on PCDTBT: PC70BM Nanocomposites. J. Electron. Mater. 2019, 48, 352–357. [Google Scholar] [CrossRef]
  56. Li, Y.; Liu, J.-L.; Chao, T.-S.; Sze, S. A new parallel adaptive finite volume method for the numerical simulation of semiconductor devices. Comput. Phys. Commun. 2001, 142, 285–289. [Google Scholar] [CrossRef]
  57. Hamza, S.; Mhamdi, A.; Aloui, W.; Bouazizi, A.; Khirouni, K.; Saidi, H. Effect of illumination on the dielectrical properties of P3HT:PC70BM nanocomposites. Mater. Res. Express 2017, 4, 055003. [Google Scholar] [CrossRef]
  58. Campoy-Quiles, M.; Ferenczi, T.A.M.; Agostinelli, T.; Etchegoin, P.G.; Kim, Y.; Anthopoulos, T.D.; Stavrinou, P.N.; Bradley, D.D.C.; Nelson, J. Morphology evolution via self-organization and lateral and vertical diffusion in polymer:fullerene solar cell blends. Nat. Mater. 2008, 7, 158–164. [Google Scholar] [CrossRef]
  59. Brahmi, J.; Nasri, S.; Saidi, H.; Nasri, H. Aouadi. K. Synthesis of new porphyrin complexes: Evaluations on optical, electrochemical, electronic properties and application as an optical sensor. Chem. Select. 2019, 4, 31–37. [Google Scholar]
  60. Ezhov, A.V.; Aleksandrov, A.E.; Zhdanova, K.A.; Zhdanov, A.P.; Klyukin, I.N.; Zhizhin, K.Y.; Bragina, N.A.; Mironov, A.F.; Tameev, A.R. Synthesis of Zn(II) porphyrin dyes and revealing an influence of their alkyl substituents on performance of dye-sensitized solar cells. Synth. Met. 2020, 269, 116567. [Google Scholar] [CrossRef]
  61. Rashmi; Kappor, A.K.; Annapoorni, S.; Kumar, V. Conduction mechanisms in poly(3-hexylthiophene) thin-film sandwiched structures. Semicond. Sci. Technol. 2008, 23, 035008. [Google Scholar] [CrossRef]
  62. Murgatroyd, P.N. Theory of space-charge-limited current enhanced by Frenkel effect. J. Phys. D: Appl. Phys. 1970, 3, 151–156. [Google Scholar] [CrossRef]
  63. Amorim, C.; Cavallari, M.; Santos, G.; Fonseca, F.; Andrade, A.; Mergulhão, S. Determination of carrier mobility in MEH-PPV thin-films by stationary and transient current techniques. J. Non-Cryst. Solids 2012, 358, 484–491. [Google Scholar] [CrossRef]
Scheme 1. Schematic representations of the structures of compounds 14.
Scheme 1. Schematic representations of the structures of compounds 14.
Molecules 27 08866 sch001
Figure 1. UV-vis spectra of 14 recorderd in dichloromethane with concentrations ~10−6 M.
Figure 1. UV-vis spectra of 14 recorderd in dichloromethane with concentrations ~10−6 M.
Molecules 27 08866 g001
Figure 2. The emission spectra of 14. The spectra have been registered in CH2Cl2 with a concentration ~10−6 M and the λmax of the excitation radiation is 430 nm.
Figure 2. The emission spectra of 14. The spectra have been registered in CH2Cl2 with a concentration ~10−6 M and the λmax of the excitation radiation is 430 nm.
Molecules 27 08866 g002
Figure 3. ORTEP drawing of [CoIII(TClPP)Cl(NTC)] with thermal ellipsoids drawn at 40% probability. The hydrogen atoms are omitted for clarity and only the major position of the disordered nicotinoyl chloride (NTC) axial ligand is shown.
Figure 3. ORTEP drawing of [CoIII(TClPP)Cl(NTC)] with thermal ellipsoids drawn at 40% probability. The hydrogen atoms are omitted for clarity and only the major position of the disordered nicotinoyl chloride (NTC) axial ligand is shown.
Molecules 27 08866 g003
Figure 4. Representation of the porphyrin core-Co(III)-Cl-NTC” moieties showing the important ruffling (a) and waving (b) deformations of the porphyrin core of [CoIII(TClPP)Cl(NTC)]. (c): Drawing of the [CoIII(TClPPCl(NTC)] complex showing the location of the nicotinoyl chloride axial ligand in the ‘‘ligand binding pocket” to minimize the interaction between the nicotinoyl chloride ligand and the phenyl groups of the TClPP porphyrinate.
Figure 4. Representation of the porphyrin core-Co(III)-Cl-NTC” moieties showing the important ruffling (a) and waving (b) deformations of the porphyrin core of [CoIII(TClPP)Cl(NTC)]. (c): Drawing of the [CoIII(TClPPCl(NTC)] complex showing the location of the nicotinoyl chloride axial ligand in the ‘‘ligand binding pocket” to minimize the interaction between the nicotinoyl chloride ligand and the phenyl groups of the TClPP porphyrinate.
Molecules 27 08866 g004
Figure 5. Formal diagram of the porphyrin macrocycle of [CoIII(TClPPCl(NTC)]. The displacement of each atom from the mean plane of the 24-atom porphyrin macrocycle in given in units of 0.01 Å.
Figure 5. Formal diagram of the porphyrin macrocycle of [CoIII(TClPPCl(NTC)]. The displacement of each atom from the mean plane of the 24-atom porphyrin macrocycle in given in units of 0.01 Å.
Molecules 27 08866 g005
Figure 6. (a): Hirshfeld surfaces mapped with dnorm ranging from −0.2017 to 2.1242 Å, (b): the shape index and (c): the curvedness.
Figure 6. (a): Hirshfeld surfaces mapped with dnorm ranging from −0.2017 to 2.1242 Å, (b): the shape index and (c): the curvedness.
Molecules 27 08866 g006
Figure 7. 2D Fingerprint plots highlighting the percentage contribution of each type of intermolecular interaction in total interactions of 4.
Figure 7. 2D Fingerprint plots highlighting the percentage contribution of each type of intermolecular interaction in total interactions of 4.
Molecules 27 08866 g007
Scheme 2. Oxidations and reductions involving the cobalt and the porphyrin core.
Scheme 2. Oxidations and reductions involving the cobalt and the porphyrin core.
Molecules 27 08866 sch002
Figure 8. Cyclic voltammograms of 4. The concentration is ca. 10−3 M in 0.1 M TBAPF6, 100 mV/s, vitreous carbon working electrode (Ø = 2 mm).
Figure 8. Cyclic voltammograms of 4. The concentration is ca. 10−3 M in 0.1 M TBAPF6, 100 mV/s, vitreous carbon working electrode (Ø = 2 mm).
Molecules 27 08866 g008
Scheme 3. Cobalt porphyrin species involved in the reductions in dichloromethane solvent.
Scheme 3. Cobalt porphyrin species involved in the reductions in dichloromethane solvent.
Molecules 27 08866 sch003
Figure 9. The change in the Ct/C0 of the MG dye with time in the photocatalyst degradation using 14.
Figure 9. The change in the Ct/C0 of the MG dye with time in the photocatalyst degradation using 14.
Molecules 27 08866 g009
Scheme 4. Proposed Photocatalytic degradation of the dye involving O2 and water.
Scheme 4. Proposed Photocatalytic degradation of the dye involving O2 and water.
Molecules 27 08866 sch004
Figure 10. (J–V) characteristics of the studied DSSCs.
Figure 10. (J–V) characteristics of the studied DSSCs.
Molecules 27 08866 g010
Figure 11. Logarithmic scale of compounds 3 and 4.
Figure 11. Logarithmic scale of compounds 3 and 4.
Molecules 27 08866 g011
Figure 12. Log–log plots of J–V curve for the ITO/Pm/Al system.
Figure 12. Log–log plots of J–V curve for the ITO/Pm/Al system.
Molecules 27 08866 g012
Table 1. Chemical shift values from 1H NMR spectra of 14 and several other porphyrins compounds.
Table 1. Chemical shift values from 1H NMR spectra of 14 and several other porphyrins compounds.
CompoundHβ-Pyrrolic Protons (ppm)H-Phenyl Protons (ppm)Ref.
Meso-arylporphyrins
H2TPP a8.848.23; 7.91; 7.67; 7.26[31]
H2TMPP b8.868.08; 7.27[32]
H2ClTPP (1)8.898.18; 7.74t.w.
[CoII(TPP)] a15.7513.10; 9.80; 7.95[31]
[CoII(TMPP)] b15.9013.10; 9.43[33]
[CoII(TClPP)] (2)15.7512.93; 9.9t.w.
[CoIII(TPP)(Cl)(py)]9.00 8.80; 7.70[34]
[CoIII(TPP)(DMI)]+,a,c8.957.86; 7.71[35]
[CoIII(TPP)Cl(DMI)] a,c8.837.87; 7.65[35]
[CoIII(TClPP)(Cl)] (3)8.957.95; 7.78t.w.
[CoIII(TClPP)(Cl)(NTC)] (4)9.088.09; 7.75t.w.
a: TPP = meso-tetraphenylporphyrinate, b: TMPP = meso-tetra(para-methoxyphenyl)porphyrinate, c: DMI = N,N′-dimethylimidazolylidene.
Table 2. UV-vis data of 14 and a selection of porphyrin species.
Table 2. UV-vis data of 14 and a selection of porphyrin species.
Compoundλmax (nm) (ε × 10−3 M−1.cm−1)Eg (eV)Ref.
Soret BandQ Bands
Free-base meso-arylporphyrins
H2TClPP (1)421(335)522(85) 557(56) 599(29) 651(36)1.88t.w.
H2TMPP a423(344)521(24) 558(20) 597(16) 650(15)1.76[21]
Cobalt meso-arylporphyrins
[CoII(TClPP)] (2)414 (340)532(56)2.01t.w.
[CoII(TPP)] b412528 [21]
[CoIII(TClPP)(Cl)] (3)442(296)557(50) 596(36)1.96t.w.
[CoIII(TClPP)Cl(NTC)] (4)455(335)560 (sh) 696(59)1.73t.w.
a: H2TMP = meso-tetratolylporphyrin, b: TPP = meso-tetraphenylporphyrinate.
Table 3. Emission spectra data, quantum yields (φf) and lifetime (τf) of complexes 14 and a selection of related porphyrin compounds.
Table 3. Emission spectra data, quantum yields (φf) and lifetime (τf) of complexes 14 and a selection of related porphyrin compounds.
Compoundλmax (nm)φf aτf b (in ns)SolventRef.
O(0,0)Q(0,1)
                  Free-base meso-arylporphyrins
H2TPP c6537220.1209.60DMF[37]
H2TMPP d6567190.0807.16CH2Cl2[32]
H2TClPP (1)6507140.0897.40CH2Cl2this work
                  Cobalt(II) meso-arylporphyrins
[CoII(TMPP)] d6557190.0356.02CH2Cl2[32]
[CoII(TClPP)] (2)6417090.046.10CH2Cl2this work
[CoII(TClPP)(4-CNpy)]6537140.062.00CH2Cl2[32]
[CoII(TPBP)(4,4′-bipy)2] e6527180.036 CH2Cl2[38]
                  Cobalt(III) meso-arylporphyrins
[CoIII(TClPP)Cl] (3)6417050.0512.30CH2Cl2this work
[CoIII(TClPP)Cl(NTC)] (4)6376990.0652.50CH2Cl2this work
a: φf = fluorescence quantum yield, b: fluorescence lifetime, c: TPP = meso-tetraphenylporphyrin, d: TMPP = meso-tetra(para-methoxyphenyl)porphyrinate, e: meso-tetrakis-[4-(benzoyloxy)phenyl]porphyrinate.
Table 4. Selection bond lengths (Å) and angles (°) for [CoIII(TClPP)Cl(NTC)]·CH2CH2 (4) and a selection of related porphyrin compounds.
Table 4. Selection bond lengths (Å) and angles (°) for [CoIII(TClPP)Cl(NTC)]·CH2CH2 (4) and a selection of related porphyrin compounds.
CompoundPorphyrin Core Deformation Type aCp–Np bCo–XL cRef.
[CoIII(TPP)(ONC6H4NMe2)2] d,eRuf ++,Wav ++1.957(1)1.94(2)/1.98(2)[40]
[CoIII(TpivPP)(2-MeIm)(2-MeHIm)] f,g,iRuf +,Wav +++1.938(3)1.972(4)/1.953(3)[41]
[CoIII(TPP)Cl(py)] d,jRuf +,Wav +1.938(3)2.234(1) 1/1.999(2) 2[42]
[CoIII(TPP)(NO2)(2,5-Lut)] d,kRuf ++,Wav ++1.9531.948 3/2.037 4[43]
[CoIII(TPP)(NO2)(Cl2py)] d,lRuf ++,Wav ++1.955(3)1.912(3) 3/2.044(3) 4[44]
[CoIII(TPP)(DMIC)(MeOH)]BF4 mRuf +++,Wav +++1.927(3)2.059(2) 5/1.929(3) 6[40]
[CoIII(TPP)(SPhF4)2] d,nPlanar1.9782.347[45]
[CoIII(TPP)Cl] dPlanar1.9842.150[46]
[CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4)Ruf +++,Wav +++1.939(4)2.230(1) 1/1.991(4) 7this work
a: See the description of various forms of the porphyrin core deformation in the text; planar refers a planar porphyrin core. +: moderate, ++, important and +++: very important, b: Co–Np =the mean equatorial distance between the center metal and the nitrogen atoms of the pyrroles, c: M–XL = distance between the center metal and the coordinated atoms of the axial ligands, d: TPP = meso-tetraphenylporphyrin, e: ONC6H4NMe2 = N,N-dimethy-l-4-nitrosoaniline, f: TpivPP = α,α,α,α-tetrakis(o-pivalamidophenyl)porphinate, g: 2-MeIm = 2-methylimidazolate, i: 2-MeHIm) = 2-methylimiodazole, j: py = pyridine, k: 2,5-Lut = 2,5 lutidine, l: Cl2py = 3,5-dichloropyridine, m: DMIC = N,N-dimethylimidazolium-2-carboxylate, n: SPhF4 = 2,3,5,6-Tetrafluorobenzenethiolato 1: chloride ligand, 2: pyridine ligand, 3: nitrito-N ligand, 4:3,5-dichloropyridine ligand, 5: N,N-dimethylimidazolium-2-carboxylate ligand, 6: MeOH ligand, 7: NTC ligand.
Table 5. Potentials (in V vs SCE) of investigated [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4) and [CoIII(TPP)Cl] [52]. The voltammograms are recorded at room temperature in dichloromethane.
Table 5. Potentials (in V vs SCE) of investigated [CoIII(TClPP)Cl(NTC)]·CH2Cl2 (4) and [CoIII(TPP)Cl] [52]. The voltammograms are recorded at room temperature in dichloromethane.
OxidationReductionRef.
Complex1st Metal Oxd2nd Metal Oxd1st Metal Red2nd Metal Red1st Metal Red
(O1, R1)(O2, R2)(1a)(1b)(3)(4) (R4, O4)
E1/2 aE1/2 aEap bEcp cEap bEcp cEap bEcp cEap bEcp c
[CoIII(TPP)Cl] d0.901.15-0.10-0.57-−0.90-−1.42[53]
[CoIII(TClPP)Cl(NTC)]1.061.43-0.37-−0.71-−1.29-−1.65this work
a: E1/2 = half-wave potential, b: Eap = anodic potential, c: Ecp = cathodic potential, d: TPP = meso-tetraphenylporphyrinate.
Table 6. Electrical parameters compounds 3 and 4.
Table 6. Electrical parameters compounds 3 and 4.
Complex J S A Φ b V
(3) 6.03   ×   10 9 0.52
(4) 5.42   ×   10 6 0.48
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nasri, S.; Guergueb, M.; Brahmi, J.; Al-Ghamdi, Y.O.; Molton, F.; Loiseau, F.; Turowska-Tyrk, I.; Nasri, H. Synthesis of New Cobalt(III) Meso-Porphyrin Complex, Photochemical, X-ray Diffraction, and Electrical Properties for Photovoltaic Cells. Molecules 2022, 27, 8866. https://doi.org/10.3390/molecules27248866

AMA Style

Nasri S, Guergueb M, Brahmi J, Al-Ghamdi YO, Molton F, Loiseau F, Turowska-Tyrk I, Nasri H. Synthesis of New Cobalt(III) Meso-Porphyrin Complex, Photochemical, X-ray Diffraction, and Electrical Properties for Photovoltaic Cells. Molecules. 2022; 27(24):8866. https://doi.org/10.3390/molecules27248866

Chicago/Turabian Style

Nasri, Soumaya, Mouhieddinne Guergueb, Jihed Brahmi, Youssef O. Al-Ghamdi, Florian Molton, Frédérique Loiseau, Ilona Turowska-Tyrk, and Habib Nasri. 2022. "Synthesis of New Cobalt(III) Meso-Porphyrin Complex, Photochemical, X-ray Diffraction, and Electrical Properties for Photovoltaic Cells" Molecules 27, no. 24: 8866. https://doi.org/10.3390/molecules27248866

Article Metrics

Back to TopTop