Mondher Dhifet*ab,
Nabil Benzerrougc,
Tahani M. Almutairi*d,
Khadra B. Alomarie,
Nikolay Tumanov
f and
Noureddine Issaoui
gh
aLaboratory of Physical Chemistry of Materials (LR01ES19), Faculty of Sciences of Monastir, Avenue of the Environment, 5019 Monastir, Tunisia. E-mail: mondherdhifet_2005@yahoo.fr
bFaculty of Sciences of Gafsa, University of Gafsa, Sidi Ahmed Zarrouk, 2112 Gafsa, Tunisia
cLaboratory of Micro-Optoelectronics and Nanostructures (LR99ES29), Faculty of Sciences of Monastir, University of Monastir, Avenue of the Environment, Monastir 5019, Tunisia
dDepartment of Chemistry, College of Science King Saud University, Riyadh 11451, Saudi Arabia. E-mail: Talmutari1@ksu.edu.sa
eJazan University, Department of Physical Sciences, Chemistry Division, P. O. Box 114, 45142, Jazan, Saudi Arabia
fUniversity of Namur, Rue de Bruxelles 61, 5000 Namur, Belgium
gUniversity of Monastir, Laboratory of Quantum and Statistical Physics LR18ES18, Faculty of Sciences of Monastir, Monastir, 5079, Tunisia
hHigher Institute of Computer Sciences and Mathematics of Monastir, University of Monastir, Monastir 5000, Tunisia
First published on 12th September 2025
In this work, we report the synthesis and spectroscopic properties of a hexa-coordinated chromium(III) porphyrin complex, namely, bis(cyanato-N)[meso-tetraphenylporphyrinato]chromate(III) (cryptand-222)sodium chloroform monosolvate with the formula [Na(2,2,2-crypt)][CrIII(TPP)(NCO)2]·0.406CHCl3 (I). Complex I was characterized in solution by UV/vis and IR spectroscopies. The structural determination of compound I was performed by single-crystal X-ray diffraction and Hirshfeld surface area calculations. This compound crystallized in the triclinic system with the non-centrosymmetric space group P1. The average distance between the central Cr(III) ion and the nitrogen atoms in the equatorial position is 2.039(3) Å, while the CrIII–N (axial ligand) distances from the bis-cyanate ligand are 2.012(3) Å and 2.016(4) Å. Crystal packing cohesion was stabilized by unconventional intramolecular C–H⋯O and C–H⋯Cl hydrogen bonds. In addition, we conducted a theoretical investigation of several key physical properties to provide a comprehensive understanding of the electron charge transfer mechanisms of the chromium(III) porphyrin complex using density functional theory (DFT) at the B3LYP-D3/LanL2DZ level. This includes the analysis of frontier molecular orbitals (FMOs) and associated reactivity descriptors; molecular electrostatic potential (MEP) assessment; non-covalent interaction (NCI) analysis through reduced density gradient (RDG) surfaces and bond critical points (BCPs); as well as electron localization function (ELF), localized orbital locator (LOL), and Hirshfeld surface analyses.
Chromium(III) porphyrin complexes are currently present in many applications such as chemistry, biology, medicine, electronics, and physics. Today, porphyrins are employed in a wide array of fields, including catalysis,6 photovoltaics,7 sensing,8 photodynamic therapy (PDT),9 supramolecular chemistry,10 photocatalysis,11 semiconductors12,13 and optoelectronic systems.14 Years of study of these compounds have shown that structural modification of the porphyrin macrocycle allows for broad variations in the physicochemical properties of porphyrins and metalloporphyrins and, in many cases, determines their effectiveness. Among synthetic porphyrins, meso-substituted derivatives are the most studied as they possess tremendous potential for structural transformation. It has been shown that their properties can be easily modulated by the nature of the substituents attached to the tetrapyrrolic macrocycle and by the choice of the metallo-complex, which has the ability to complex almost any metal ion in the periodic table, for example, zinc(II) and magnesium(II) complexes with different axial ligands.15 Furthermore, the central metal of these metalloporphyrins can coordinate axial ligands of various types. Axial coordination is a process that profoundly alters the spectroscopic, photophysical, and structural properties of metalloporphyrins.
In order to better understand the electrochemical and spectroscopic properties of chromium porphyrins, we synthesized a new hexacoordinated coordination compound, titled bis(cyanato-N)[meso-tetraphenylporphyrinato]chromate(III) (cryptand-222) sodium chloroform monosolvate. This compound and its precursors were characterized by various spectral methods (UV-visible and IR). Structural determination was performed by single-crystal X-ray diffraction, complemented by a Hirshfeld surface study of complex I. To strengthen the overall coherence of this study, we performed a comprehensive theoretical analysis of the investigated complex through advanced computational methods, aiming to gain deeper insights into its structural and electronic behavior.
Fourier transform infrared (FT-IR) spectra were recorded using a PerkinElmer Spectrum Two spectrometer.
Electronic absorption spectra were recorded at room temperature using a SHIMADZU UV-2401 spectrometer.
Elemental analyses were performed using a Flash EA 1112 Series analyzer (Thermo Electron) fitted with a Prepack column 2 m (PTFE) and an MX5 microbalance (Mettler Toledo). Hirshfeld surfaces (HSs) and 2D fingerprint plots (FPs) were generated using CrystalExplorer 17.5,17,18 based on the data obtained from X-ray diffraction analysis. The dnorm function represents a normalized ratio that incorporates the distances from a designated surface point to the closest internal (di) and external (de) atoms, relative to the van der Waals radii associated with the respective atoms.17,19
(i) The oxygen O2 and the carbon C46 of the one axial ligand NCO− of one [CrIII(TPP)(NCO)2]− ion complex are disordered in two positions: C46A-O2A/C46B-O2B with major position occupancies of 0.56873(1).
(ii) The two counterions [Na1(2,2,2-crypt)]+ and [Na2(2,2,2-crypt)]+ are disordered over two orientations with refined occupancy coefficients converged to 0.56873(1) (for the first [Na1A(2,2,2-crypt)]+) and 0.71542(1) (for the second [Na2A(2,2,2-crypt)]+).
(iii) The chloride and the carbon atoms of chloroform solvent are disordered in two positions: C129-Cl1A-Cl2A-Cl3A/C130-Cl1B-Cl2B-Cl3B with major position occupancies of 0.75574(2). Hydrogen atoms were positioned using calculated coordinates and refined as riding on their respective parent atoms. Geometry-related analyses were performed using the PLATON program,23 and molecular and packing representations were created using MERCURY.24 The crystallographic and refinement characteristics of complex I are summarized in Table 1.
a R1 = ∑‖Fo| − |Fc‖/∑|Fo|.b wR2 = {∑[w(|Fo|2 − |Fc|2)2]/[w(|Fo|2)2]}1/2. | |
---|---|
Formula | C64.41H64.41Cl1.22CrN8NaO8 |
Formula weight, M | 1196.69 |
Crystal system | Triclinic |
Space group | P1 |
a (Å) | 12.9467(12) |
b (Å) | 16.0849(15) |
c (Å) | 17.0355(16) |
α (°) | 63.553(3) |
β (°) | 69.614(3) |
γ (°) | 71.150(3) |
V (Å3) | 2917.1(5) |
Z | 2 |
ρcalc. (g cm−3) | 1.362 |
μ (mm−1) | 0.176 |
F (000) | 1253 |
Crystal size (mm3) | 0.22 × 0.12 × 0.11 |
Crystal color | Purple |
Crystal shape | Block |
T (K) | 150 |
θmin − θmax (°) | 2.039–21.552 |
Limiting indices | −16 ≤ h ≤ 16, −20 ≤ k ≤ 20, −22 ≤ l ≤ 22 |
R (int) | 0.0775 |
R (sigma) | 0.0454 |
Measured reflections | 23![]() |
Reflections independent | 27![]() |
Parameters/restraints | 2053/2760 |
S [goodness of fit] | 1.034 |
R1a, wR2b [Fo > 4σ(Fo)] | R1 = 0.0439, wR2 = 0.1125 |
wR2 [all data] | R1 = 0.0530, wR2 = 0.1187 |
Min./max. res. (eÅ−3) | 0.369/−0.388 |
CCDC | 2414564 |
Elemental analysis calcd (%) for C64.41H64.41Cl1.22CrN8NaO8 (1196.694 g mol−1): C 64.64, H 5.42, N 9.36; found: C 64.93, H 5.46, N 9.46;
UV-visible: [λmax (nm) in CHCl3, (logε)]: 403 (4.25), 454 (5.13), 534 (2.56), 583 (3.41), 618 (3.72);
FT-IR [solid, ν (cm−1)]: 2964–2820: [ν(CH) Porph]; 2165: [ν(CN) NCO-ligand]; 1093: [ν(–CH2–O–CH2–) crypt-222]; 1007: [δ(CCH) Porph].
![]() | ||
Fig. 1 UV-Vis spectrum of H2TPP, recorded in CHCl3 at room temperature with a concentration of ∼10−6 M. |
Fig. 2 shows the UV-visible absorption spectra, recorded in chloroform, for the starting compound [CrIII(TPP)Cl] and our derivative [Na(2,2,2-crypt)][CrIII(TPP)(NCO)2]·0.406CHCl3 (I). The maximum wavelengths (λmax) of these compounds were compared with those of other arylporphyrins and Cr(III)-arylporphyrin complexes, as shown in Table 2.
![]() | ||
Fig. 2 UV-Vis spectra of [CrIII(TPP)Cl] (blue) and complex I (red), recorded in CHCl3 at room temperature with a concentration of ∼10−6 M. |
Compounds | λmax (nm) (ε × 10−3 M−1 cm−1) | Eg-opt (eV) | References | |||||
---|---|---|---|---|---|---|---|---|
Soret band | Q bands | |||||||
a TMPP: meso-tetra-methoxyphenylporphyrinato.b TClPP: meso-tetra-chlorophenylporphyrinato.c TTP: meso-tetratolylporphyrinato.d OAc: Acetato.e 1-MeIm: 1-methylimidazole. | ||||||||
Free meso-arylporphyrins in CHCl3 | ||||||||
[H2(TPP)] | 419(6.08) | 515(5.73) | 550(4.33) | 590(4.20) | 650(4.21) | 1.86 | This work | |
[H2(TMPP)]a | 420(5.95) | 518(5.81) | 556(4.37) | 595(4.24) | 653(4.17) | 1.87 | 28 | |
[H2(TClPP)]b | 417(5.95) | 515(5.82) | 553(4.28) | 598(4.17) | 646(4.10) | 1.87 | 29 | |
[H2(TTP)]c | 422(6.11) | 519(5.03) | 556(4.94) | 594(4.84) | 651(4.85) | 1.86 | 29 | |
![]() |
||||||||
Cr(III) meso-arylporphyrin complexes in CHCl3 | ||||||||
[CrIII(TTP)Cl] | 400(4.44) | 452(5.15) | 525(3.75) | 566(3.39) | 605(3.90) | 1.94 | 30 | |
[CrIII(TPP)(NCO)2]− | 404(4.24) | 453(5.14) | 532(2.59) | 584(3.39) | 619(3.71) | 1.97 | 30 | |
[CrIII(TPP)(OAc)]d | 396(3.34) | 450(5.20) | 523(3.65) | 565(3.99) | 605(3.90) | — | 31 | |
[CrIII(TPP)Cl] | 401(4.26) | 452(5.15) | 525(2.75) | 556(3.62) | 598(3.82) | 1.93 | This work | |
[CrIII(TPP)(NCO)2]− | 403(4.25) | 454(5.13) | 534(2.56) | 583(3.41) | 618(3.72) | 1.96 | This work | |
![]() |
||||||||
Cr(II) meso-arylporphyrin complexes in THF | ||||||||
[CrII(TPP)] | 402 | 421 | 460 | 516 | 601 | 655 | — | 32 |
[CrII(TPP)(1-MeIm)2]e | 404 | 423 | 453 | 532 | 620 | 682 | — | 32 |
The electronic spectrum of the free base H2TPP shows a strong absorption band, known as the Soret (or B) band, positioned at 419 nm. This band corresponds to a permitted transition from the initial state to the second excited state (S0 ← S2). Additionally, four less intense Q bands are observed at 515, 550, 590, and 650 nm. These are assigned to the Qy (1, 0), Qy (0, 0), Qx (1, 0), and Qx (0, 0) absorption bands, respectively, and correspond to forbidden transitions from the ground state to the first excited state (S0 ← S1).
In the UV-visible study of porphyrin complexes, it is noted that the grouping in the meta-substitution of the porphyrin core has no effect. However, the insertion of the metal and the coordination of a ligand have a large effect.33 This can be explained by an increase in π-conjugation, resulting from the addition of the donor axial ligand.
As shown in Fig. 2, the electronic spectra of [CrIII(TTP)Cl] and complex I in a CHCl3 solution exhibit the low-intensity band (LMCT band) in the Soret region, which is slightly red-shifted compared to that of the chloride chromium(III) derivative. The λmax values of the Soret bands of the chloride and bis(cyanato-N) complexes are not the same, and the absorption bands in the Q region are red-shifted. This probably can be explained by the negative charge of this complex.34 Comparing our synthesized compound with pentacoordinated Cr(III) complexes, we deduced that the Soret band is deflected little towards the red, while it is deflected a lot towards the red compared with the Cr(II) complexes. The shift in the Soret and Q bands in the absorption spectra of [CrIII(TPP)Cl], [CrIII(TPP)(CN)2]−,35 and [CrIII(TPP)(OCN)2]− has been discussed. This hypsochromic shift was found to be axial ligand dependent, and from the spectrochemical series of ligands, it is concluded that the anionic ligand CN− is a stronger π-acceptor ligand than NCO−, leading to the transfer of electrons from the Cr-porphyrins, which may be an effect on the Soret band.
The optical band gap is a key property describing semiconductor physics. The optical band gap energy (Eg-opt) values of H2TPP and [CrIII(TPP)Cl] (see Fig. S1), calculated using the Tauc method,36,37 are 1.86 and 1.93 eV, respectively. However, in complex I, this value is 1.96 eV, which was determined using the tangent method according to the formula:
Compound | ν(NH) | ν(CH) | δ(CCH) | ν(CN)Laxb | Reference |
---|---|---|---|---|---|
a Absorption bands IR (cm−1).b Lax: axial ligand.c TTP: meso-tetratolylporphyrinato.d TMPP: meso-tetra-methoxyphenylporphyrinato.e NCS: thiocyanato. | |||||
H2(TPP) | 3321 | 2930–2829 | 969 | — | This work |
H2(TTP)c | 3317 | 3020–2930 | 952 | — | 29 |
H2(TMPP)d | 3313 | 2996–2814 | 969 | — | 38 |
[CrIII(TPP)Cl] | — | 3025–2847 | 1000 | — | This work |
[CrIII(TTP)Cl] | — | 3025–2847 | 1002 | — | 30 |
[CrIII(TMPP)Cl] | — | 3059–2849 | 1001 | — | 39 |
[Na(222)][CrIII(TPP)(NCO)2] | — | 2973–2819 | 1007 | 2165 | This work |
[K(222)][CrIII(TTP)(NCO)2] | — | 2973–2819 | 1006 | 2196–2149 (sh) | 30 |
[K(222)][CrIII(TMPP)(NCS)2]e | — | 2957–2810 | 1005 | 2050 | 39 |
H2TPP exhibits a characteristic IR spectrum of a meso-arylporphyrin with ν(NH) and ν(CH) stretching frequencies at 3420 cm−1 and in the range of 2930–2829 cm−1, respectively.
The metalation of H2TPP with chromium(II) chloride dihydrate (CrCl2·2H2O) in THF to form the complex [CrIII(TPP)Cl] leads to the disappearance of the absorption band ν(NH) and the shift toward the high fields of the δ(CCH) bending vibration from 969 cm−1 to 1000 cm−1.
For our compound [Na(2,2,2-crypt)][CrIII(TPP)(NCO)2], there is a strong band in the IR spectrum at 1093 cm−1, which is attributed to the counterion (cryptand-222) sodium(+), and the δ(CCH) bending frequency value is 1007 cm−1. The absorption bands assigned to νa(CN) and νs(CN) of the cyanate group are 2165 cm−1 and 1300 cm−1. The presence of the counterion and δ(CCH)porph at 1007 cm−1 confirmed the presence of hexacoordinate chromium(III) porphyrin species with the bis cyanate axial ligands.
The UV-visible and IR spectroscopies cannot give us a decisive answer or confirm the ground-state electronic configuration of the central-ion.
Fig. 4 depicts an ORTEP diagram40 of the [CrIII(TPP)(NCO)2]− ion complex.
![]() | ||
Fig. 4 ORTEP view of the [CrIII(TPP)(NCO)2]− ion complex; thermal ellipsoids are drawn at the 30% probability level. Hydrogen atoms are omitted for clarity. |
The Cr(III) center metal is coordinated to the four nitrogen atoms of the TPP porphyrinato and the bis-cyanate axial ligands with the nitrogen atoms.
Fig. S3 and 5 illustrate the ORTEP diagrams of the [Na(2,2,2-crypt)][CrIII(TPP)(NCO)2]·0.406CHCl3 complex and the [Na(2,2,2-crypt)]+ counterion, respectively (only the major position of the disordered is shown).
The charge of the anionic complex [CrIII(TTP)(NCO)2]− is balanced by the counterion [Na(2,2,2-crypt)]+. The sodium atom is eight-coordinated, where it is coordinated to two nitrogen atoms and six oxygen atoms of the cryptand-222. The average Na–O(crypt-222) distance is 2.569(7) Å (K–O (2,2,2-crypt), 2.774 Å30), and the average Na–N(crypt-222) bond length is 2.812(8) Å ((K–N (2,2,2-crypt), 3.099 (6) Å)30). It is concluded that there is a small difference between both complexes, indicating the good atomic organization and stability of our Na counterion.
Several bond distances and angles of complex I, as well as those of several related Cr(III) meso-tetra-arylporphyrins are listed in Table 4.
Complex | Cr–Npa | Cr–Laxb | Reference |
---|---|---|---|
a Average equatorial distance between the chromium center metal and the nitrogen atoms of the pyrrole rings.b Chromium–axial ligand distance.c TTP refers to tetratolylporphyrin. | |||
[CrIII(TPP)(NCO)2]− | 2.039(3) | 2.012(3)/2.016(3) | This work |
[CrIII(TTP)(NCO)2]−c | 2.042(4) | 2.019(4)/2.019(4) | 30 |
[CrIII(TPP)(N3)(py)] | 2.031(5) | 2.004(5) (azido)/2.135(5) (py) | 41 |
[CrIII(TPP)(Cl)(py)] | 2.038(8) | 2.294(4) (Cl)/2.12(1) (py) | 42 |
[CrIII(TPP)(Cl)(H2O)] | 2.043(2) | 2.311(4) (Cl)/2.057(2) (H2O) | 43 |
The bis Cr2(III)–N(NCO) bond lengths are 2.012(3) and 2.016(3) Å, respectively, which are within the range of the two non-porphyrinic (cyanato-N)-chromium(III) complexes reported in the literature: [CrIII(NCO)2(L1)2]− (L1 = N,N-bis(3,5-di-tert-butylsalicylidene)-1,2-ethylenediamino)44 and [CrIII(NCO)(NO)2(Cp)] (Cp = cyclopentadienyl)45 with values of 2.025 and 1.980 Å, respectively.
Fig. 6 represents the coordination polyhedron of the bis-cyanato chromium(III) tetraphenylporphyrin derivative (I). Thus, the Cr(III) cation is located at the center of the porphyrin core and defines a distorted square pyramidal environment of four nitrogen atoms of the porphyrin macrocycle and the bis-cyanate axial ligand of two nitrogen atoms.
The bis-axial NCO− ligand of complex I was found to be practically linear with N11–C91–O3 and N12–C92–O4 angles of 178.37(7)° and 179.67(7)°, respectively, as has been observed for all known complexes containing the cyanate-N ligand. The value of the Cr2–N11–C91 angle is equal to 174.2(4)° and is within the typical range (137–180°) for various complexes with the cyanato-N ligand.
The N11–C91 and C91–O3 distances in the one axial NCO− ligand are 1.146(7) and 1.203(9) Å, respectively, which are very close to those of the (cyanato-N)–porphyrinic and non-porphyrinic Cr(III) complexes reported in the literature (Table 5). These bond lengths are also similar to those of the NaOCN salt (N–C = 1.16 Å and C–O = 1.27 Å), which probably indicates an ionic character of the Cr2–N(NCO) bond.
Complexes | M–N(NCO) | N–C | C–O | N–C–O | M–N–C | Reference |
---|---|---|---|---|---|---|
a TPP = meso-tetratolylporphyrinato.b TpivPP = α,α,α,α-tetrakis(o-pivalamidophenyl)porphyrinate.c L1 = N,N-bis(3,5-di-tert-butylsalicylidene)-1,2-ethylenediamine.d Cp = cyclopentadienyl. | ||||||
Cyanato-N metalloporphyrins | ||||||
[CrIII(TPP)(NCO)2]− | 2.012(3) | 1.146(7) | 1.203(9) | 178.37(7) | 174.2(4) | This work |
2.016(3) | 1.151(7) | 1.210(9) | 179.67(7) | 152.03(4) | ||
[CrIII(TTP)(NCO)2]− | 2.019(4) | 1.148(7) | 1.234(9) | 174.6(13) | 155.2(5) | 30 |
[Mg(TPP)(NCO)]− | 2.0471(17) | 1.165(2) | 1.210(2) | 178,34(2) | 157.34(15) | 46 |
[Zn(TTP)(NCO)]−a | 2.0293(3) | 1.200(5) | 1.163(4) | 179.66 | 137.00(3) | 47 |
[MnIII(TPP)(NCO)] | 2.029(5) | 1.124(9) | 1.200(9) | 179.25 | 150.0(5) | 48 |
[FeII(TpivPP)(NCO)]−b | 2.005(3) | 1.150(5) | 1.200(5) | 179.9(5) | 176.6(3) | 49 |
[CoIII(TPP)(NCO)2]− | 1.905 | 1.151 | 1.202 | 178.6 | 159.8 | 50 |
1.919 | 1.155 | 1.193 | 176.9 | 144.6 | ||
![]() |
||||||
Non-porphyrinic cyanato-N chromium(III) complexes | ||||||
[CrIII(NCO)2(L1)2]−c | 2.025 | 1.173 | 1.209 | 177.01 | 162.08 | 44 |
[CrIII(NCO)(NO)2(Cp)]d | 1.980 | 1.126 | 1.180 | 178.59 | 180 | 45 |
Fig. 7 is a formal diagram of the porphyrin macrocycle of [CrIII(TPP)(NCO)2]−, which shows the displacements of each atom of the C20N4Cr plane from the 24-atom core plane of the porphyrin in units of 0.01 Å. As shown in this figure, the distance between iron and the 24-atom mean plane of the porphyrin ring (Cr-PC = 1.08(1) Å) is very small; this explains that the porphyrin ring for [CrIII(TPP)(NCO)2]− exhibits a planar conformation.
![]() | ||
Fig. 7 Formal diagrams of the porphyrin core of [CrIII(TPP)(NCO)2]− illustrating the displacement of each atom from the 24-atom core plane in units of 0.01 Å. |
The content of the unit cell is depicted in Fig. 8, which is made by two [CrIII(TPP)(NCO)2]− ion complexes, two [Na(2,2,2-crypt)]+ counterions and 0.406 molecules of chloroform solvent.
![]() | ||
Fig. 8 Cell content of complex I. Only the major positions of the disordered regions are shown. Hydrogen atoms have been omitted for clarity. |
The intermolecular interactions within the crystal lattice have been generated using the PLATON program23 are: classic hydrogen bands such as non-conventional hydrogen bonds such as C–H⋯O and C–H⋯Cl interactions.23,51
For complex I, the intermolecular interactions responsible for the stability of the crystal lattices are of type C–H⋯O and C–H⋯Cl.
The visualization of the intermolecular interactions within the crystal lattice of complex I was made using the MERCURY program.24 These intermolecular contacts are depicted in Fig. 9 and 10, while the values of these distances are given in Table S1.
![]() | ||
Fig. 9 Drawing illustrating the C–H⋯Cl intermolecular interactions in the crystal lattice of complex I. |
![]() | ||
Fig. 10 Drawing illustrating the C–H⋯O and C–H⋯Cl intermolecular interactions in the crystal lattice of complex I. |
As shown in Fig. 9, the chloride Cl1A of the chloroform solvent is H-bonded to the carbon C88 of one pyrrole ring of a nearby [CrIII(TTP)(NCO)2]− ion complex with a C88–H88⋯CL1A distance of 3.668(5) Å. This Cl1A atom is also linked to the carbon C128 of a close [Na(2,2,2-cryp)]+ counterion with a C128–H12R⋯Cl1A distance of 3.672(5) Å. Fig. 10 illustrates the C–H⋯O and C–H⋯Cl intermolecular interactions. The oxygen O3 of the one axial ligand NCO− of one [CrIII(TPP)(NCO)2]− ion complex and the two carbon atoms C95A and C96A of a neighboring [Na(2,2,2-cryp)]+ counterion are linked by a weak H bond with C95A–H95B⋯O3 and C96A–H96B⋯O3 distances of 3.209(3) Å and 3.048(3) Å, respectively. The second chloride Cl2A of the chloroform solvent is H bonded to the carbon C9 of one pyrrole ring of a nearby [Cr(TTP)(NCO)2]− ion complex with a C9–H9⋯CL2A distance of 3.714(5) Å.
![]() | (1) |
To make this work more consistent, we have explored the Hirshfeld surface analysis to deeply understand the intermolecular interactions within the complex I that are exhibited by X-ray diffraction. Fig. 11 demonstrates the 3D Hirshfeld surfaces mapped with different surfaces for chromium(III) porphyrin complexes.
![]() | ||
Fig. 11 3D Hirshfeld surfaces mapped with (a) the shape index surface, (b) curvedness surface, (c) dnorm, (d) di and (e) de for chromium(III) porphyrin complexes under consideration. |
Firstly, Fig. 11a demonstrates the shape index surface, which describes the topology of the electron density distribution. Generally, it is characterized by these two types of mapping: concave and convex regions depicted in red and blue, respectively, typically indicative of hydrogen bond acceptors and donors. Additionally, the next illustration of Fig. 11b displays the mapped curvedness of our chromium(III) porphyrin complex. It can be observed that there exist several green flat areas separated by blue edges related to high values of curvedness, which contribute to get information about interactions between the closest molecules, which are commonly associated in our case with van der Waals contacts as the white regions in the surface mapped with dnorm.
Moreover, in Fig. 11c of dnorm is illustrated using a red-to-blue color gradient, ranging from −0.2858 to 1.6287 Å, respectively. The red color visualized with circular depressions or spots in the HS indicates O–H⋯O, C–H⋯Cl and C–H⋯N hydrogen bonds. However, weaker interactions compared to hydrogen bonds are determined with blue- and white-colored regions in HS. As well, the dnorm surface is taken as a combination of the di and de of Fig. 11d and e by eqn (1).
In the same context, we carried out the two-dimensional fingerprint plots for the chromium(III) porphyrin complex in Fig. 12.
Accordingly, the most important contributions of interactions responsible for the structure stability are considered as H⋯H, C⋯H/H⋯C and O⋯H/H⋯O contacts for 49.8%, 14% and 6.1%, respectively, of the total HS area. Moreover, the interatomic bonds of Cl⋯H/H⋯Cl, C⋯C and N⋯H/H⋯N contribute less than the first ones, with 2.2%, 2.1% and 2% from the total area, respectively.
Commonly, all the 3D Hirshfeld surfaces with the 2D-fingerprint data conform considerably with experimental results, confirming the strong stability and molecular order with atomic arrangement of this synthetized bis(cyanato-N)[meso-tetraphenylporphyrinato]chromate(III) (cryptand-222) sodium chloroform monosolvate.
Obviously, the theoretical values of bond lengths and angles presented in Table 6 are highly close to those extracted from X-ray diffraction analysis data, which display good agreement with the experimental ones, while there is some difference only for angles related to the sodium-cryptand-222, which is due to the difference in matter phase as gas in the theoretical calculation compared to the solid phase in realistic experiment, which confirms that this optimized structure is excellent for deeply understanding more different physical properties of such a substance.
Distance (Å) | Experimental | Theoretical | Angle (°) | Experimental | Theoretical |
---|---|---|---|---|---|
Chrome(III) coordination polyhedron of I | |||||
Cr–N4 | 2.038 | 2.0529 | N4–Cr–N5 | 90.00 | 89.1089 |
Cr–N5 | 2.032 | 2.0582 | N4–Cr–N7 | 90.06 | 91.1891 |
Cr–N6 | 2.047 | 2.064 | N4–Cr–N8 | 91.68 | 90.6889 |
Cr–N7 | 2.056 | 2.054 | N4–Cr–N9 | 88.94 | 88.3412 |
Cr–N8 | 2.012 | 1.9161 | N5–Cr–N6 | 89.64 | 90.706 |
Cr–N9 | 2.016 | 2.0035 | N5–Cr–N8 | 90.03 | 91.279 |
N5–Cr–N9 | 91.05 | 89.2877 | |||
N6–Cr–N7 | 90.30 | 88.9621 | |||
N6–Cr–N8 | 88.41 | 90.1833 | |||
N6–Cr–N9 | 90.98 | 90.7883 | |||
N7–Cr–N8 | 89.71 | 90.9484 | |||
N7–Cr–N9 | 89.21 | 88.4908 | |||
![]() |
|||||
Sodium-cryptand-222 | |||||
Na84–O86 | 2.543 | 2.5269 | O86–Na84–O87 | 102.29 | 85.8633 |
Na84–O87 | 2.614 | 2.597 | O86–Na84–O88 | 94.56 | 96.9821 |
Na84–O88 | 2.577 | 2.5109 | O86–Na84–O90 | 92.08 | 113.2318 |
Na84–O89 | 2.698 | 2.6702 | O87–Na84–O88 | 64.11 | 66.3752 |
Na84–O90 | 2.462 | 2.8026 | O87–Na84–O89 | 99.33 | 92.3405 |
O87–Na84–O90 | 114.79 | 99.1895 | |||
O88–Na84–O89 | 109.63 | 87.4619 | |||
O88–Na84–O90 | 173.34 | 145.8867 | |||
O89–Na84–O90 | 63.79 | 61.4342 |
ELUMO | −2.067 |
EHOMO | −4.203 |
Eg = ELUMO − EHOMO | 2.136 |
Ionization potential (I ≈ −EHOMO) | 4.203 |
Electron affinity (A ≈ −ELUMO) | 2.067 |
Chemical potential (μ = 1/2 (ELUMO + EHOMO)) | −3.135 |
Global hardness (η = 1/2 (ELUMO − EHOMO)) | 1.068 |
Global softness (S = 1/2η) | 0.468 |
Mulliken electronegativity (χ = −μ) | 3.135 |
Global electrophilicity index (w = μ2/2η) | 4.601 |
Crucial molecular orbitals that display valuable critical parameters in quantum chemistry are very well known by HOMO and LUMO, representing the highest occupied molecular orbital of the valence band and the lowest unoccupied molecular orbital of the conduction band, respectively.54 The advantages related to these orbitals lie in the robust estimation of the chemical stability and the electron localization in reactivity sites for further realizing the process of the electronic charge transfer.55 Moreover, the FMOs contribute greatly to the consideration of the energy separation value between HOMO and LUMO energy levels, known as the gap energy Eg to better understand information about electronic properties with the identification of the competitor groups of donors and acceptors constituting the complex under study, thereby improving its application in different areas as optoelectronic devices and sensor technologies.
According to the results, one may clearly observe that HOMO and LUMO electron densities spread similarly, and their distributions are predominantly localized in the porphyrin ring and the axial ligand around the central anionic chrome(III) ion. It means that donor and acceptor electrons prefer these regions in the structural molecule and the charge transfer can be achieved in such localization between the central ion and its surrounding porphyrin core. As well, our chemical compound exhibited semi-conductor properties as a result of the calculated gap energy, which was found to be approximately 2.136 eV, providing our material suitable application characteristics for next-generation technologies and several electronic devices. The small difference between the theoretical energy gap compared to the experimental one, 1.968 eV, can be attributed to the matter's phase. More precisely, the realistic molecular system is synthesized as a solid material, while the calculated chemical system in the B3LYP/LanL2DZ level of DFT is considered to be more conditioned, as with the gas phase.
According to Table 7, the chemical potential of our complex compound is estimated as a negative value, leading to it being biologically active56 and having better molecular stability compared to [K(2,2,2-crypt)][CrIII(TTP)(NCO)2]·2H2O.30 In other words, it is difficult to spontaneously disintegrate into separated elements. Also, the global hardness can identify the resistance to different deformations of the electron cloud after chemical treatment of compounds.57 This Cr(III) complex has a little greater hardness than [CrIII(TPP)(Cl)(H2O)]43 and [K(2,2,2-crypt)][CrIII(TTP)(NCO)2]·2H2O.30
Notably, the most pronounced electron-accepting regions – colored in red – are located around the chloride axial ligands and the adjacent cationic porphyrin ring, suggesting that these sites possess significant electrophilic reactivity. Conversely, the blue-colored electron-donating region is primarily observed in the cationic species [Na(2,2,2-crypt)]+. The spatial coexistence of these donor and acceptor regions enhances the potential for efficient intramolecular electron charge transfer within the chloride-coordinated complex.
From these results, the color map in this figure was determined from blue to red, passing through green, related to the strong attraction of hydrogen bonds, strong repulsion of steric effects, and van der Waals interactions. Clearly, the strong repulsions can be seen around the central anionic chromium(III) ion with N atoms in the porphyrin ring. Furthermore, the existences of van der Waals bonds are much depicted in many regions of the chemical compound.
Besides, the definition of BCP was determined as the point into the bond pathway between two nuclei (or atoms), from which their gradient vector of ρ(r) can be found to be equal to zero (∇ρ(r) = 0). The bond nature in such BCP is classified by the Laplacian ∇2ρ(r). Meanwhile, the bond order defining the strength is relative to the charge density ρ(r). Furthermore, the BCPs of the complex compound are presented in Fig. 17, and their topological parameters are summarized in Table 8.
![]() | ||
Fig. 17 Different views of the molecular representation of the complex compound with several bond critical points (BCPs). |
BCP | ρ(r) | ∇2ρ(r) | H(r) | G(r) | V(r) |
---|---|---|---|---|---|
1 | 0.4028 | −0.6547 | −0.6320 | 0.4683 | −1.1003 |
2 | 0.3977 | −0.6815 | −0.6161 | 0.4457 | −1.0619 |
3 | 0.3728 | −0.5050 | −0.5597 | 0.4334 | −0.9931 |
4 | 0.3620 | −0.5670 | −0.5267 | 0.3849 | −0.9116 |
5 | 0.3045 | −0.8377 | −0.3212 | 0.1118 | −0.4330 |
6 | 0.3043 | −0.8347 | −0.3209 | 0.1123 | −0.4332 |
7 | 0.2888 | −0.6114 | −0.3016 | 0.1488 | −0.4504 |
8 | 0.2887 | −0.6121 | −0.3033 | 0.1502 | −0.4535 |
9 | 0.0034 | 0.0130 | 0.9712 | 0.2269 | −0.1298 |
10 | 0.0033 | 0.0106 | 0.7757 | 0.1896 | −0.1120 |
First, we see that the upper BCP(1) to BCP(4) have the highest ρ(r) values, which are 0.4028, 0.3977, 0.3728, and 0.3620 a.u., respectively. Such bonds represent the strongest interactions in the entire chemical complex I related to the chloride axial ligand, between the nitrogen, carbon, and oxygen atoms are bound by N9C83, N8
C82, C82
O2 and C83
O3. Secondly, the inverse phenomenon is figured straightly in the lower bonds in the table, and the weakest bonds correspond to van der Waals interactions at BCP(9) and BCP(10), which connect the cationic part to the anionic one with H145⋯H104 and C25⋯H104, indicating the interaction colored in green in the representation of NCI and RDG iso-surface in Fig. 16.
However, besides the bonds first discussed, the next four bonds (BCP(5) → BCP(8)) relate carbon atoms with carbons and also nitrogen only around the axial ligand in the porphyrin ring. Moreover, these findings are confirmed with both of the RDG function iso-surfaces shown in Fig. 16 and the MEP results shown in Fig. 15.
Furthermore, the localized orbital locator (LOL) is related to the orbital gradient and presents a prominent descriptor of chemical bonding,62 providing an explanation of the adapted regions of localized orbitals to maximum overlapping behavior.63 As qualitative manner, the lowest and highest LOL values characterize the fastest and slowest electrons; such slow particles are associated with localized electrons that could be found in bonds or lone pairs. Both the ELF and LOL iso-surfaces are complementary to each other and display valuable computational methods used in quantum chemistry depending on the kinetic energy density.
The findings of Fig. 18 illustrate the ELF and LOL iso-surfaces of the studied compound with XY and XZ projections determined using the Multiwfn program.64
Clearly, the shaded color maps of the ELF and LOL confirm the strong electronic localization around hydrogen (H) atoms, with a maximum value colored with red indicating the presence of bonding and non-bonding electrons like a covalent bond or a lone pair of electrons in such a region. Meanwhile, the delocalization of the electronic cloud illustrated with blue areas is observed surrounding carbon (C) atoms in the studied compound. Additionally, it should be mentioned that a little white color tint is figured in central regions of some H atoms, which indicates that the electronic density exceeded the maximum color scale limit in such regions. Hence, these differentiations in regions strengthen the electronic charge transfer; otherwise, the more different regions there are in the compound, the better the complex is advantageously stable with the formation of electrostatic interactions between atoms.65,66 These meaningful conclusions are further supported by MEP and FMO analysis.
CCDC 2414564 contains the supplementary crystallographic data for this paper.67
Supplementary information: UV/Vis spectra and X-ray molecular structures are provided in Fig. S1–S3 and Table S1. See DOI: https://doi.org/10.1039/d5ra04378f.
This journal is © The Royal Society of Chemistry 2025 |