Next Article in Journal
Mechanistic Insight of Sensing Hydrogen Phosphate in Aqueous Medium by Using Lanthanide(III)-Based Luminescent Probes
Next Article in Special Issue
Heterofullerene MC59 (M = B, Si, Al) as Potential Carriers for Hydroxyurea Drug Delivery
Previous Article in Journal
A Novel Thermochemical Metal Halide Treatment for High-Performance Sb2Se3 Photocathodes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Theoretical Study of the Occupied and Unoccupied Electronic Structure of High- and Intermediate-Spin Transition Metal Phthalocyaninato (Pc) Complexes: VPc, CrPc, MnPc, and FePc

1
Dipartimento di Scienze Chimiche, Università degli Studi di Padova, via F. Marzolo 1, 35131 Padova, Italy
2
Istituto di Chimica della Materia Condensata e di Tecnologie per l’Energia (ICMATE-CNR), via F. Marzolo 1, 35131 Padova, Italy
*
Authors to whom correspondence should be addressed.
Submission received: 10 November 2020 / Revised: 18 December 2020 / Accepted: 22 December 2020 / Published: 28 December 2020
(This article belongs to the Special Issue Theoretical Calculation and Molecular Modeling of Nanomaterials)

Abstract

:
The structural, electronic, and spectroscopic properties of high- and intermediate-spin transition metal phthalocyaninato complexes (MPc; M = V, Cr, Mn and Fe) have been theoretically investigated to look into the origin, symmetry and strength of the M–Pc bonding. DFT calculations coupled to the Ziegler’s extended transition state method and to an advanced charge density and bond order analysis allowed us to assess that the M–Pc bonding is dominated by σ interactions, with FePc having the strongest and most covalent M–Pc bond. According to experimental evidence, the lightest MPcs (VPc and CrPc) have a high-spin ground state (GS), while the MnPc and FePc GS spin is intermediate. Insights into the MPc unoccupied electronic structure have been gained by modelling M L2,3-edges X-ray absorption spectroscopy data from the literature through the exploitation of the current Density Functional Theory variant of the Restricted Open-Shell Configuration Interaction Singles (DFT/ROCIS) method. Besides the overall agreement between theory and experiment, the DFT/ROCIS results indicate that spectral features lying at the lowest excitation energies (EEs) are systematically generated by electronic states having the same GS spin multiplicity and involving M-based single electronic excitations; just as systematically, the L3-edge higher EE region of all the MPcs herein considered includes electronic states generated by metal-to-ligand-charge-transfer transitions involving the lowest-lying π* orbital (7eg) of the phthalocyaninato ligand.

Graphical Abstract

1. Introduction

Phthalocyanines (H2Pc) share with porphyrins (H2P), everywhere present “as far as the living world is concerned’’ [1], the same four nitrogen-based coordinative pockets. Even though H2Pc and its metal complexes (MPc) are not present in Nature, they have been attracting great interdisciplinary interest because their technological potential spans over a wide range of applications [2,3]. Besides traditional appliances, such as dyestuffs for textiles and inks [3], MPcs are currently used as intrinsic semiconductors, chemical sensors, organic light-emitting diodes, organic photovoltaic cells, thin-film transistors, materials for nonlinear optics, spintronics and laser recording [4,5,6,7,8]. Moreover, bio-inspired oxygen-binding MPcs have been shown as viable substitutes for precious metals in catalysts for the oxygen reduction reaction (ORR) in low-temperature fuel cells [9,10,11,12]. As intimate an understanding as possible of the origin, symmetry, and strength of the M–Pc interaction is then mandatory to enhance the efficiency of new MPc-based devices. As such, X-ray absorption spectroscopy (XAS) is unanimously recognized as a valuable tool to probe, element-selectively, the empty electronic structure of M complexes, the M coordinative environment, as well as the nature and the strength of the M–ligand interaction [13,14,15,16].
Metal L2,3-edges’ spectral features are related to the electronic states generated by the electric dipole-allowed 2p → nd excitations [17], thus providing information about the contribution of the M-based nd atomic orbitals (AOs) to the frontier virtual molecular orbitals (VMOs). Metal-based 2p6ndk → 2p5ndk+1 excitations create a hole in the M 2p core AOs with an angular momentum quantum number = 1, and spin-orbit coupling (SOC) allows to couple with s , whose quantum number s = 1/2. Two distinct states are then produced (j = 3/2 and j = 1/2), with the former (the L3-edge) lying at lower excitation energy (EE) and having an intensity approximately twice that of the L2-edge associated with j = 1/2. Besides the M L2,3-edges, the ligand (L) donor atom K-edge XA spectra of ML complexes with partly filled nd AOs are usually characterized by rather intense pre-edge features. These are associated with the electronic states generated by the electric dipole-allowed L-based 1s → mp transitions [17], whose intensity gauges the L mp character of frontier VMOs [18,19,20,21]. XAS at the L K-edge thus directly probes the so-called M−L symmetry-restricted covalency [22], affording information complementary to that gatherable by XAS at the M L2,3-edges. Despite the fact that the L2,3-edges spectra of ML complexes contain a huge amount of chemical information, their first-principle modelling is theoretically demanding because, besides the ligand field and covalency effects, SOC between the possible many final-state multiplets has to be considered [23,24,25,26,27,28,29,30,31,32].
At the very beginning of this century, Koshino et al. [33] recorded the L2,3 excitation spectra of MPc (24 ≤ Z ≤ 29; Z corresponds to the atomic number of the metals they considered) by exploiting the inner-shell electron energy-loss spectroscopy (ISEELS); a little over ten years later, Kroll et al. [34] investigated the electronic structure of MPc (25 ≤ Z ≤ 30) by combining soft L-edge XAS and 2p photoemission spectroscopy. In addition, few years ago, Eguchi et al. [35] succeeded in the ultra-high vacuum synthesis and XAS characterization of VPc on Ag(111), while neither XAS nor ISEELS data have been so far reported for TiPc to our knowledge.
As a part of a systematic investigation of the electronic properties of energy-targeted materials, some of us have recently investigated their structure/reactivity relationships by exploiting XAS at the N K-edge and at the M L2,3-edges of diverse MPc (M = V [36,37], Fe [12,38] and Cu [37,39,40]) and CuTPP/CuTPP(F) [41] (H2TPP = tetraphenylporphyrin; H2TPP(F) = tetrakispentafluorophenyl porphyrin) surface-supported films coupled to quantum mechanical calculations. Two different methodologies were adopted to model the M L2,3-edges’ features: the current Density Functional Theory variant of the Restricted Open-Shell Configuration Interaction Singles (DFT/ROCIS) method [42] (VPc, FePc and CuPc) and the Relativistic Time-Dependent DFT (RTD-DFT), including SOC with full use of symmetry and correlation effects [43] within the Tamm−Dancoff approximation [44] (CuPc, CuTPP, CuTPP(F)). As such, it has to be noted that, besides L donor atom K-edge XA spectra of ML complexes [39,41,45,46,47], the RTD-DFT approach may be employed to satisfactorily models the M L2,3-edges XAS features of closed shell [46,48] and CuII complexes. Indeed, the modelling of the CuII L2,3-edges features corresponds, among open shell complexes, to the simplest possible case because the electric dipole allowed 2p6… 3d9 → 2p5…3d10 transitions generate a final configuration, which has only two term symbols. The corresponding spectral splitting is dominated by the 2p SOC contribution and the overall energetics and intensities are strongly influenced by ligand-field and covalency, respectively [40]. At variance to that, the RTD-DFT approach is unable to suitably describe SOC in open-shell molecules as explicitly reported in the ADF manual. All the MPcs herein considered (M = V, Cr, Mn and Fe) are open-shell systems with a quite complex electronic structure. A RTD-DFT modelling of their XAS features would then be utterly inadequate, while they can be properly handled by exploiting the module ROCIS of the ORCA program package [42].
In this contribution, a homogeneous modelling of the VPc, CrPc, MnPc, and FePc L2,3-edges’ features is presented and discussed with the ultimate goal of providing a first-principle rationale of differences characterizing the M–Pc interaction along the investigated series. As such, it can be useful to mention that, with the exception of VPc, recently synthesized in extreme conditions [35], the remaining MPcs herein considered all have relevant catalytic applications, including the: (i) CO and NO oxidation as well as ORR (CrPc) [49,50,51,52]; (ii) oxidation reactions in homogeneous and heterogeneous phase (MnPc) [9]; (iii) N-alkylation [53], C–H amination [54], C–C bond formation [55], synthesis of esters [56] and oximes [57] as well as reduction [58], oxidation [59,60], and radical reactions (FePc) [61]. Moreover, bio-inspired oxygen-binding Fe-macrocycles are particularly appealing as a consequence of their ability to reversibly bind O2, a crucial step in processes such as respiration, photosynthesis or the ORR catalysis [62]. Thus, there is no doubt that a comprehension as intimate as possible of the M–Pc interaction is of paramount importance to design new MPc-like species devoted to specific purposes.

2. Computational Details

The ground state (GS) electronic and structural properties of title molecules have been herein investigated by exploiting the Amsterdam Density Functional (ADF) package [63] within the assumption of an idealized D4h symmetry [17] (see Figure 1), by running spin-unrestricted, nonrelativistic DFT calculations, with generalized gradient corrections self-consistently included through the Becke−Perdew formula [64,65], by adopting a triple-ζ with a polarization function Slater-type basis set for all the atoms, and by freezing the M 1s–2p AOs and the 1s AO of N and C atoms throughout the calculations. MPc optimized Cartesian coordinates are reported in Tables S1–S4 of the Supplementary Materials.
Insights into the origin, symmetry, and strength of the M–Pc interaction have been gained by combining the Nalewajski−Mrozek approach [66,67,68,69,70,71], well suited to estimating bond multiplicity indices (NMI) in M complexes [66,67,68,69,70,71,72], with the Ziegler’s extended transition state (ETS) method [73]. According to the ETS scheme, the MPc bonding energy (BE) may be written as:
B E = ( Δ E e s + Δ E P a u l i + Δ E o r b )
where ΔEes accounts for the pure electrostatic interaction, ΔEPauli represents the destabilizing two-orbital−four-electrons interaction between the occupied orbitals of the interacting fragments (only atomic fragments have been herein considered), and ΔEorb corresponds to the stabilizing interaction between the occupied and empty orbitals of the atomic fragments. In passing, ΔEorb may be further decomposed into contributions due to the different irreducible representations (IRs) of the D4h point group, according to
Δ E o r b = χ Δ E o r b χ ,
The MPc L2,3-edges’ XA spectra [33,34,35] have been modelled by evaluating EEs and corresponding oscillator strength distributions (f(EE)) for transitions with the M 2p-based MOs as initial spin orbitals (isos), by means of the DFT/ROCIS method [30], which includes SOC in a molecular Russell−Saunders fashion [25], by adopting the B3LYP exchange–correlation (XC) functional [74] for VPc, CrPc and FePc and the M06 meta-GGA XC [75] for MnPc (vide infra), and by using the def2-TZVP(-f) basis set [76,77]. The combined use of DFT and configuration interaction requires a set of three semi-empirical parameters (c1 = 0.18, c2 = 0.20, and c3 = 0.40), which have been calibrated by Roemelt and Neese [25] for a test set of M L2,3-edges. Throughout the M L2,3-edges modelling, the resolution of identity approximation has been used with the def-TZVP/J basis set [76,77]. Moreover, the zero-th order regular approximation has been adopted to treat the scalar relativistic effects [78]. Numerical integrations for DFT/ROCIS calculations have been carried out on a dense Lebedev grid (302 points) [79]. In addition, MPc-modelled spectra have been shifted by 10.9 (VPc), 9.8 (CrPc), 8.3 (MnPc) and 13.1 (FePc) eV to superimpose the highest intensity features of the simulated and experimental L3-edge, which does not suffer from the extra broadening and the distortion due to the Coster–Kronig Auger decay process [32,80]. This was needed because absolute theoretical EEs carry errors arising from DF deficiencies in the core region, one-particle-basis set restrictions and inadequacies in the modelling of spin-free relativistic effects [24].

3. Results and Discussion

3.1. MPc Occupied Electronic Structure

The aim of obtaining an understanding as intimate as possible of the origin, symmetry and strength of the M–Pc interaction may benefit from a preliminary, qualitative description of the MPc frontier orbitals simply based on symmetry arguments and overlap considerations. MPcs are united by the presence of the Pc2− ligand whose electronic properties have been thoroughly described elsewhere [81]. Pc2− frontier MOs may be split into σ and π sets. MPc symmetry adapted linear combinations (SALC) of C and N 2pσ (C and N 2pπ) are bases for the following D4h IRs: a1g, a2g, b1g, b2g, eu (eg, a1u, a2u, b1u, b2u); moreover, among π MOs, no a1u SALC of N 2pπ AOs, no b1u SALC of NPy 2pπ AOs and no b2u SALC of Nm 2pπ AOs (see Figure 1) is present. In addition, the four NPy lone pairs pointing toward the centre of the coordinative pocket are bases for the a1g, b1g and eu IRs. In more detail, the two Pc2− highest occupied MOs (HOMOs) correspond to the 15b1g MO, σ in character and strongly localized on the NPy lone pairs pointing toward the centre of the coordinative pocket, and the 2a1u π MO. In this regard, it is noteworthy that the D4h a1u IR is anti-symmetric with respect to the reflections through the σh, σv and σd symmetry planes of the D4h point group; the a1u π MOs have then a node on symmetry planes passing through NPy and Nm atoms (see Figure 1). As far as the D4h Pc2− lowest unoccupied MO (LUMO) is concerned, the 6eg VMO has a π character too.
The presence of the Pc2− square planar ligand field lifts the five-fold degeneracy of the M 3d AOs, generating a 3dσ and a 3dπ set. The former set includes the 21a1g (z2-based) and the 16b1g (x2–y2-based) MOs, while the latter takes in the 14b2g (xy-based) and the 6eg (xz- and yz-based) ones [17]. Among them, the 21a1g and the 14b2g MOs are substantially M–NPy non-bonding, while the 6eg and 16b1g MOs are M–NPy π and σ antibonding, respectively. Relative energy positions of the 3dσ/3dπ spin up (↑)/spin down (↓) sets in VPc, CrPc, MnPc and FePc are displayed in Figure 2, together with those of selected Pc-based MOs.
VPc ground state. Experimental [35] and theoretical [36,37,82] results disagree about the VPc ground state (GS) spin multiplicity. Carlotto et al. [36,37] recently proposed a 4Eg high-spin (HS) GS ( a 1 g 1   b 2 g 1   e g 1   b 1 g 0 ; see Figure 2 and Table 1; possible Jahn–Teller distortions [83] associated with orbitally degenerate GS or excited states have not been taken into account), while Eguchi et al. [35] presumed a 2Eg low-spin (LS) GS ( a 1 g 0   b 2 g   2 e g 1     b 1 g 0 ). The 2Eg state generated by the a 1 g 0 b 2 g 2 e g 1 b 1 g 0 configuration is 52.1 (35.5) kcal/mol less stable than the HS 4Eg (LS 2B1g) one; moreover, the 4B1g/4A2g states generated by the constrained a 1 g 0 b 2 g 1 e g 2 b 1 g 0 / a 1 g 1 b 2 g 0 e g 2 b 1 g 0 configurations are 1.8/1.7 kcal/mol less stable than the 4Eg GS. Consistently with the VPc a 1 g 1   b 2 g 1   e g 1   b 1 g 0 GS configuration, the NMIV-NPy is quite large (0.64) [36]; as such, since NMI includes both covalent and ionic contributions, it is of some relevance to mention that the V Hirshfeld [84] charge (QV) amounts to 0.34.
A closer look at the frontier VPc GS electronic structure indicates that all the V 3d-based singly occupied MOs (SOMOs) lie well above the ring-based, V-free, 2a1u π doubly occupied MO (DOMO). The ionization energies (IEs) of VPc frontier MOs are not available in the literature; nevertheless, their values may be estimated by exploiting the Slater transition state (TS) method [85], which allows the evaluation of excitation/ionization energies “… by means of an artificial state that is halfway between the ground state of an atom or molecule and an excited state” [86]. Interestingly, the lowest VPc TSIE (6.17 eV) is associated with the ionization from the V 3dπ-based 6eg SOMO rather than with the photoemission from the ring-based, V-free, 2a1u π DOMO (6.59 eV). In this regard, it is of value to highlight that in his seminal paper devoted to the investigation of gas-phase photoelectron (PE) spectra of H2Pc and MPc (M = Mg, Fe, Co, Ni, Cu, and Zn), J. Berkowitz pointed out that ‘‘… the first ionization potential occurs at ~6.4 eV, and it varies almost imperceptibly from sample to sample, including metal free and MgPc. Therefore, the conclusion seems inescapable that the first ionization potential corresponds to electron ejection from a ring orbital, and not a metal orbital’’ [87].
The comparison of the VPc bond lengths and bond angles with those optimized for the other MPc (see Table 2) [88] indicates that the structural perturbations induced by the presence of different MII ions in the Pc2- coordinative pocket are rather minute.
CrPc ground state. As already mentioned, CrPc has been attracting great interest as a catalyst for the CO and NO oxidation, as well as for the ORR [49,50,51,52], thus making particularly interesting the study of its electronic structure. CrII has a 3d4 configuration, which may generate three spin states with S = 0 (LS), S = 1 (intermediate spin; IS) and S = 2 (HS). Any attempt to optimize the LS state failed (NC, non-converged in Table 1), while the 3Eu IS state, associated with the a 1 g 1 b 2 g 1 e g 2 b 1 g 0 2 a 1 u 0 / a 1 g 0 b 2 g 0 e g 1 b 1 g 0 2 a 1 u 1 configuration, has been found less stable than the 5B1g HS one, generated by the a 1 g 1 b 2 g 1 e g 2 b 1 g 0 configuration, by 25.6 kcal/mol (IS and HS CrPc optimized structures are perfectly superimposable). Incidentally, the 3Eu IS state implied a non-Aufbau occupation accompanied by a pseudo reduction (oxidation) of the CrII ion (macrocycle). As such, even though the 5B1g HS GS has been experimentally revealed [89,90] and theoretically predicted [36,82], the localization of VMOs is still controversial. Indeed, SIESTA [91] numerical experiments carried out by Arillo-Flores et al. [82] are consistent with the absence of “… metal contributions to HOMO and LUMO, they principally localize upon the inner ring.”, which is certainly correct for the M-free 2a1u HOMO, but wrong for the 6eg LUMO. In addition, differently from VPc, the ring-based, Cr-free, the 2a1u π DOMO corresponds to the CrPc HOMO (see Figure 2). Analogously to VPc, the IEs of the CrPc frontier MOs are not available in the literature, but, differently from VPc, the lowest TSIE value (6.60 eV) is estimated for the ionization from the ring-based, Cr-free, 2a1u π HOMO.
Upon moving from VPc to CrPc, the GS frontier electronic configuration evolves from a 1 g 1 b 2 g 1 e g 1 b 1 g 0 to a 1 g 1 b 2 g 1 e g 2 b 1 g 0 . The addition of an electron to the 3dπ-based 6eg MO, M–NPy anti-bonding could then be invoked to rationalize the NMIM-NPy reduction from 0.64 (see above) to 0.43. Nevertheless, three things need to be kept in mind before drawing conclusions: (i) as already mentioned, NMI includes both covalent and ionic contributions; (ii) QCr (0.49) is larger than QV (0.34); (iii) the 3dπ-based 6eg MO is more anti-bonding in CrPc than in VPc (the localization % of the VPc 6eg MO on NPy is negligible; see Figure 3).
Analogously to VPc, no crystallographic data are available in the literature for CrPc; nevertheless, the tiny differences between the CrPc- and VPc-optimized structural parameters seem to indicate that a subtle balance between ionic and anti-bonding covalent contributions to the M–Pc interaction, both of them larger in CrPc than in VPc, takes place.
MnPc ground state. Likewise CrPc, LS, IS and HS states are possible; moreover, even though the optimized structural parameters corresponding to different spin states are very similar, the 4Eg IS state associated to the a 1 g 1 b 2 g 1 e g 3 b 1 g 0 configuration (see Figure 2) has been found more stable than the 6Eg HS and the 2B2u LS ones by 16.5 and 14.9 kcal/mol, respectively. As such, it has to be noted that: (i) the MnPc IS GS has been experimentally [94,95,96,97,98,99] and theoretically [36,100,101,102,103,104,105] assessed; (ii) the 6Eg HS state is generated by the a 1 g 1 b 2 g 1 e g 2 b 1 g 0 7 e g 1 configuration; (iii) the 2B2u LS state has the following occupation numbers: a 1 g 1 b 2 g 1 e g 2 b 1 g 0 2 a 1 u 0 / a 1 g 0 b 2 g 0 e g 2 b 1 g 0 2 a 1 u 1 . In this contest, it is noteworthy that: (i) both the 7eg and the 2a1u MOs are ring-based π orbitals, so that the HS (LS) state would imply a pseudo MnII oxidation (reduction) with the Mn 3d-based orbitals occupied by four (six) electrons (HSQMn, ISQMn and LSQMn amount to 0.41, 0.33 and 0.28, respectively); (ii) to our knowledge, only two contributions [33,106] suggested a MnPc HS state. As regards the MnPc IS GS, a further controversy concerns its symmetry, or, equivalently, the occupation numbers of the Mn 3d-based 21a1g, 14b2g and 6eg MOs. Three different configurations may be considered: a 1 g 1 b 2 g 1 e g 3 b 1 g 0 (4Eg), a 1 g 1 b 2 g 2 e g 2 b 1 g 0 (4A2g) and a 1 g 2 b 2 g 1 e g 2 b 1 g 0 (4B1g). In agreement with the theoretical results of Brumboiu et al. [105], the ADF outcomes rule out the 4B1g state because of its high energy; moreover, it is noteworthy that different experimental studies support either a 4Eg or a 4A2g GS. Specifically, XAS evidence [98], magnetic circular dichroism (MCD)/UV-Vis results [95], and XAS/MCD outcomes [99] favour a 4Eg GS [107], while magnetic susceptibility measurements have been rationalized within the assumption of a 4A2g GS [95,97], which has been attributed to intermolecular interactions in the MnPc crystal. In agreement with the literature [101], the ADF results herein reported estimate the 4Eg state to be more stable than the 4A2g by 7.3 kcal/mol.
Before going on, it is of some relevance to point out that the MnPc a 1 g 1 b 2 g 1 e g 3 b 1 g 0 GS configuration implies the presence of a high-lying Mn 3dπ-based SOMO well above the ring-based, Mn-free, 2a1u π DOMO (see Figure 2). As such, no gas-phase photoemission results are available in the literature for MnPc; however, Grobosch et al. [108] were able to record the He(I) photoemission spectrum of an MnPc thin film deposited on polycrystalline Au. Interestingly, they assigned the lowest lying peak of the MnPc valence band photoemission spectrum to the ionization from the Mn 3dπ-based 6eg MO, observing also that the ΔIE between the ring-based, Mn-free, π 2a1u DOMO and the 3dπ-based 6eg orbital is ~0.5 eV [108,109]. In perfect agreement with these findings [108], the TSIEs [85] of the highest-lying 6eg and 2a1u orbitals are 5.99 and 6.51 eV, respectively.
Among the investigated molecules, MnPc is the lightest one for which structural parameters are available [88,92]. The data reported in Table 2 reveal that optimized bond lengths and bond angles fairly reproduce experimental evidence. In this context, the NMIMn-NPy value (0.52), just in between the NMICr-NPy and the NMIV-NPy ones (0.43 and 0.64, respectively), seems to indicate that the M–NPy bond-weakening associated with the addition of a further electron to the 3dπ-based 6eg MO is negligible (see Figure S1 of the Supplementary Materials). In addition, it has to be underlined that QMn (0.33) and QV (0.34) are almost identical.
FePc ground state. Similarly to MnPc, experimental [34,110,111,112,113,114] and theoretical [38,104,115] evidence indicates a FePc IS GS whose symmetry is, however, still debated. On the computational side, the IS GS symmetry, inextricably linked to the occupation numbers of frontier MOs, has been found to be extremely sensitive to the adopted XC functional and basis set. Carlotto et al. [38] have proposed a 3Eg IS GS, generated by a a 1 g 1 b 1 g 2 e g 3 b 2 g 0 configuration, on the basis of numerical experiments carried out by employing the ORCA program package [42], by using the hybrid B3LYP XC–functional [74], by adopting the def2-TZVP(-f) basis set [76,77] and the c1, c2, and c3 semi-empirical parameters 0.21, 0.49, and 0.29 (hereafter, old set), respectively. In passing, the a 1 g 1 b 1 g 2 e g 3 b 2 g 0 GS configuration is a consequence of the orientation of the molecule in the xy plane. The x and y axes of the framework they adopted point toward the Nm atoms rather than toward the NPy ones (see Figure 1). The four NPy lone pairs were then bases for the a1g, b2g and eu IRs rather than for the a1g, b1g and eu ones, and the Fe 3d-based VMO accounting for the Fe-NPy σ anti-bonding interaction corresponded to the 14b2g level rather than to the 16b1g one. All the possible configurations compatible with a FePc IS GS have been herein considered. As such, even though the ADF 3B2g ( a 1 g 1 b 2 g 1 e g 4 b 1 g 0 ) and the 3Eg ( a 1 g 1 b 2 g 2 e g 3 b 1 g 0 ) IS states are less stable than the 3A2g one ( a 1 g 2 b 2 g 2 e g 2 b 1 g 0 ) by minute amounts (1.4 and 1.1 kcal/mol, respectively), it is of some relevance to underline that the Δ E o r b χ (χ = a1g, b2g and eg) of the IS states differ by up to ~230 kcal/mol ( Δ E o r b e g = −2584.4, −2810.4 and −2690.4 kcal/mol in the 3A2g, 3B2g and 3Eg states, respectively). A further ADF 3Eg state may be generated by the a 1 g 2 b 2 g 1 e g 3 b 1 g 0 configuration. Besides the non-Aufbau filling of the corresponding electronic levels, the latter 3Eg state is significantly less stable than the 3A2g GS (12.3 kcal/mol). Additional numerical experiments have been carried out to estimate the FePc BE of the LS and HS states. As far as the former is concerned, it may imply either the a 1 g 0 b 2 g 2 e g 4 b 1 g 0 configuration or the a 1 g 2 b 2 g 0 e g 4 b 1 g 0 one. The LS a 1 g 2 b 2 g 0 e g 4 b 1 g 0 frozen configuration generates a non-Aufbau energy level filling, and a BE lower (~32 kcal/mol) than that of the 3A2g GS; any attempt to get a converged BE value for the LS a 1 g 2 b 2 g 0 e g 4 b 1 g 0 frozen configuration failed. Analogous considerations hold for the HS state.
FePc is the lightest MPc for which gas-phase photoemission spectroscopy data have been recorded [87]. According to experimental evidence, the TSIE of the ring-based, Fe-free, π 2a1u DOMO is the lowest one. Incidentally, its value (6.51 eV) is very close to the one (6.49 eV) estimated by Liao and Scheiner [101], even though a non-relativistic approach has been herein adopted.
According to the experiment in Ref. [93], the shortening of the M–NPy distance upon moving from MnPc to FePc is correctly reproduced (see Table 2). In this context, it is of relevance to mention that QFe = 0.13; i.e., the smallest value along the investigated series. Such a result, coupled to the NMIFe-NPy = 0.64, ultimately indicates that the Fe–NPy interaction is the most covalent among the molecules we took into account (look at the MPc ΔEorb values reported in Table 3). In addition, the presence of a strong Fe–NPy covalent σ interaction is consistent with the high-energy position of the 16b1g VMO (see Figure 2), which accounts for the σ anti-bonding interaction between the Fe 3dx2-y2 AO and the b1g SALC of the NPy lone pairs.

3.2. MPc Unoccupied Electronic Structure

The MPc L2,3-edges’ XA spectra [33,34,35,38] herein modelled have been collected from deposits of different thicknesses, often consisting of randomly oriented and weakly interacting molecules. Thus, according to a procedure successfully tested in the past [36,37,38,39,40,41,116,117,118,119,120], XAS outcomes have been rationalized by completely neglecting the adsorbate/substrate interactions. Now, before moving to the modelling of the MPc XAS features, a few words about the 2p → 3d excitations simply based on symmetry arguments may be useful to facilitate the forthcoming discussion. In a simplified picture of the 2p → 3d one-electron excitations, the MII electronic configuration moves from the starting …2p6…3dk to the ending …2p5…3dk+1 with 3 ≤ k ≤ 6 for 23 ≤ Z ≤ 26. The electronic states associated to the …2p5…3dk+1 configurations, straightforwardly obtained by evaluating the 2P ⊗ Dk+1 direct products (the whole set of multiplets arising from a particular 3dk+1 configuration is herein collectively labelled Dk+1) [121], are collected in Tables S5–S7 of the Supplementary Materials. The ligand-field, covalent interactions and SOC admixture further split them, to generate totals of 6 × 210 = 1260 (k + 1 = 4), 6 × 252 = 1512 (k + 1 = 5), 6 × 210 = 1260 (k + 1 = 6), and 6 × 120 = 720 (k + 1 = 7) molecular magnetic spin sublevels, respectively. The MII electronic configurations of the MPc herein considered (HS VII a 1 g 1 b 2 g 1 e g 1 b 1 g 0 , HS CrII a 1 g 1 b 2 g 1 e g 2 b 1 g 0 , IS MnII a 1 g 1 b 2 g 1 e g 3 b 1 g 0 , IS FeII a 1 g 2 b 2 g 2 e g 2 b 1 g 0 ) allow us to foresee that the one-electron excitation patterns describing the MII final states in the D4h symmetry should in principle include states having a spin multiplicity either equal to (ΔS = 0) or lower/higher than (ΔS = ±1) the GS. The ΔS = 0 spin-selection rule is slightly released when SOC is considered; more specifically, SOC connects the terms with resultant spins S and S′, where |S − S′| = 0, 1 [122].
Electric dipole-allowed transitions imply that [17]
ΓGS ⊗ Γμ ⊗ ΓES ⊃ ΓSym,
where ΓGS, Γμ, ΓES and ΓSym correspond to the IRs of the MPc electronic GS, the dipole moment operator (a2u + eu) [17], the electronic excited state (ΓES = Γiso ⊗ ΓGS ⊗ Γfso; iso and fso stand for initial and final spin orbitals, respectively), and the totally symmetric representation of the D4h point group (a1g), respectively. Equation (3) may then evolve to
Γiso ⊗ (a2u + eu) ⊗ Γfso ⊃ a1g,
which implies that, within the adopted approximation, which reduces the complete one-electron excited configuration space (1h–1p space) to the subspace where only the M 2p core electrons (transforming as a2u + eu) [17] are excited, the allowed electric dipole transitions are
(a2u → a1g)
(a2u → eg)||
(eu → eg)
(eu → a1g/a2g/b1g/b2g)||
where the ||/⊥ symbols stand for parallel/perpendicular to the molecular σh plane (see Figure 1).
VPc L3-edge. To date, only Eguchi et al. have succeeded in synthesizing, even though in extreme conditions, surface-supported mono- and multi-layers of VPc, whose angle-dependent linearly polarized XA spectra at the V L2,3-edges are reported in Ref. [35]. As such, a thorough analysis of the || and ⊥ V L2,3-edges components of the VPc and OVPc XA spectra has been recently reported by Carlotto et al. [36,37]. With specific reference to the VPc complex, the authors were able to conclusively assess its spin state by modelling corresponding XAS features for both LS (S = 1/2) and HS (S = 3/2) states. A brief description of their HS results [36] is herein included to favour the comparison with the modelled spectra of diverse MPcs.
It has been already mentioned that the one-electron excitation pattern describing the VII final states in D4h symmetry should be dominated by states which may have (see Table S5 of the Supporting Materials) either a spin multiplicity equal to (ΔS = 0) or lower/higher (ΔS = ±1) than the GS one [122]. As such, it has to be underlined that the ORCA B3LYP HS GS corresponds to the 4A2g state generated by the a 1 g 1 b 2 g 0 e g 2 b 1 g 0 electronic configuration. B3LYP/ROCIS outcomes [36,37] indicate that both the ||f(EE) and the f(EE) distributions of the L3-edge mainly arise from states having ΔS = 0, while ΔS = ± 1 contributions are negligible. Moreover, the lowest-lying ||/L31 features (see Figure 4a) are due to states generated by V 2p → SOMOs single electronic excitations. Incidentally, SOMOs correspond to the V 3d-based 6eg and 21a1g orbitals, while states associated to coupled-single electronic excitations [25,32], mainly involving V 2p → SOMOs and SOMOs → VMOs excitations, contribute to ||/L32. SOMOs naturally correspond to the 21a1g and 6eg MOs, while the VMOs are the V 3d-based 14b2g (~90%) and the π* Pc-based 7eg (~10%) orbitals. In passing, the MPc 7eg VMO corresponds to the lowest-lying Pc-based π* orbital and metal-to-ligand-charge-transfer (MLCT) transitions in diverse MPc L3-edge spectra (vide infra) involve this VMO. No contribution from V 2p → 16b1g excitations is provided to states associated with the VPc L3-edge.
The agreement between the experimental evidence and B3LYP/ROCIS outcomes has been documented in detail elsewhere [36,37]; here, it is sufficient to underline that both the relative positions and relative intensities of spectral features are well reproduced by ||/f(EE). Major disagreements between theory and experiment usually affect the L2 region [31]; nonetheless, it deserves mentioning that the B3LYP/ROCIS ||L21||L32 HSΔEE (6.4 eV, see Figure 4a) fairly reproduces the experimental value (6.9 eV). Any further comment about the VPc L2-edge is herein avoided.
CrPc L3-edge. To date, the only CrPc f(EE) distribution for the 2p excitations is the one recorded by Koshino et al. by exploiting ISEELS [33]. Their spectrum includes both the L3- and the L2-edge, but the coarse EE scale they adopted prevents the possibility of revealing the presence of possible structures associated to them; moreover, no information is provided by the authors about the L2–L3 ΔEE. Similarly to VPc, the one-electron excitation pattern describing the Cr final states in the CrPc D4h symmetry is dominated by states which may have (see Table S6 of the Supporting Material) a spin multiplicity either equal to (ΔS = 0) or lower/higher (ΔS = ±1) than the GS one. The HS CrPc ||f(EE) and f(EE) B3LYP/ROCIS distributions are superimposed upon the CrPc ISEEL spectrum in Figure 4b. In the L3-edge region, ||/f(EE) consists of an intense peak (L31) with an evident shoulder on its higher EEs (L32, ΔEE = 1.40 eV). Moreover, the B3LYP/ROCIS outcomes also indicate that both ΔS = 0 and ΔS = −1 states, both of them associated with single electronic excitations, contribute to the ||/f(EE) distributions. In more detail, ||/L31 features are caused by ΔS = 0 states generated by Cr 2p → SOMOs transitions, with the whole set of the half-occupied Cr 3d-based orbitals (6eg, 21a1g and 14b2g SOMOs) as fsos. At variance with that, the ||/L32 shoulders include contributions from both ΔS = 0 (64%) and ΔS = −1 (30%) states. Quintet states have the same origin as those associated with ||/L31, while the triplet ones are related to Cr-based 2p → 16b1g and MLCT 2p → π* Pc-based transitions. As such, it is noteworthy that states associated with the Cr-based 2p → 16b1g transition only contribute to ||L32. Analogously to VPc, any comment about the CrPc L2-edge is herein avoided as it cannot be unambiguously determined by experiment [32]; nevertheless, we underline that the B3LYP/ROCIS L2–L31 HSΔEE (8.6 eV, see Figure 4b) fairly reproduces the experimental value (7.9 eV).
MnPc L3-edge. The MnPc L2,3-edges’ XA spectrum recorded by Koshino et al. [33] suffers from the same issues already mentioned for CrPc. Otherwise, the experimental evidence reported by Kroll et al. [34] for MPc (25 ≤ Z ≤ 30) is much more informative as a consequence of the EE range they showed for each single MPc, thus allowing the detection of the fine structure eventually contributing to spectral features; moreover, both || and ⊥ polarized XA spectra are reported for each MPc. In the forthcoming discussion, ISEELS outcomes of Koshino et al. will no longer be considered as a reference for the L2,3-edges XAS modelling herein presented. At least three well-evident and closely spaced peaks (herein labelled ||L31, ||L32 and ||L33, and lying at ~640.0, ~641.0 and ~643.0 eV, respectively) contribute to the MnPc ||L3-edge spectrum (see Figure 4c). In addition, the ||L33’s higher EE side is characterized by the presence of an evident shoulder at ~644.0 eV. The || → ⊥ light polarization switching is accompanied by a significant relative intensity reduction in spectral features having EEs in between 642 and 644 eV, thus reducing the L3-edge spectrum to the L31 and L32 peaks with comparable intensity and lying at ~640.5 and ~641.5 eV, respectively (see Figure 4c).
The MnPc B3LYP/ROCIS ||/f(EE) distributions, not herein included, poorly reproduces the L3-edge XA spectrum in terms of number of peaks, relative energy positions and relative intensities. A few years ago, Carlotto et al. [123] tested the efficiencies of several diverse XC functionals (non-hybrid, hybrid and hybrid meta-GGA) in reproducing the L2,3-edges’ absorption spectra of Mn complexes. The use of the hybrid M06 meta-GGA XC functional [75] was found to be decisive for a detailed assignment of the Mn(acac)3 (acac = acetylacetonato) L2,3-edges XAS features. Now, despite the inability even of the M06 XC functional to reproduce in detail the complex structure of the MnPc XA spectrum, in particular the presence of the closely spaced ||L31 and ||L32 peaks, the IS-corresponding modelling (see Figure 4c) provides a satisfactory agreement between experiment and theory. In more detail, theoretical results indicate that the lowest-lying feature of the ||f(EE) distribution (||L31, representative of the ||L31 and ||L32 experimental peaks; see Figure 4c) has to be associated to the quartet electronic states (ΔS = 0) generated by single electronic excitations having the whole Mn 2p set as isos and the low-lying Mn 3d-based 21a1g, 14b2g and 6eg VMOs as fsos. At variance with that, electronic states with ΔS = 0 (58%) and ΔS = 1 (24%) contribute to the intense ||L33 feature, while electronic states with ΔS = -1 negligibly contribute to the ||/L3 patterns. Former states (ΔS = 0) imply Mn 2p → SOMOs (21a1g, 14b2g and 6eg) and SOMOs → 16b1g/7eg coupled-single excitations [25,32], while the latter (ΔS = 1) are mainly generated by Mn 2p → π* Pc-based single electronic excitations. The ΔS = 1 Mn 2p → π* Pc-based 4b2u single electronic excitation violates selection rules stated by Eqns 5–8. As such, it has to be mentioned that i) ORCA calculations have to be run in C1 symmetry and ii) the f value of the ΔS = 1 Mn 2p → π* Pc-based 4b2u single electronic excitation is very low.
The satisfactory agreement between ||/f(EE) distributions and experimental evidence [34] prompts us to further detail the proposed assignment. Despite the fact that electronic states generating the ||/L31 features of Figure 4c are associated with Mn-based ΔS = 0 2p → 3d single electronic excitations involving the whole 2p set and the low-lying Mn 3d-based ↓ VMOs, it sounds reasonable that the (1a2u → 21a1g) and (1eu → 6eg) transitions contribute to the L3 lower EE region more than the (1a2u → 6eg)||, (1eu → 21a1g)|| and (1eu → 14b2g)|| ones, while the opposite is true when L33 is considered. In fact, once again in agreement with the experimental results of Kroll et al. [34], the electronic states determined by ΔS = 0, 1 coupled-single and single ⊥ polarized electronic transitions provide to the L3 higher EE region a contribution significantly lower than those || polarized.
DFT/ROCIS calculations fairly reproduce both the L2–L3 ΔEE (~10 eV) and the corresponding relative intensities; nevertheless, any detailed assignment of the L2 feature is herein avoided as it is not unambiguously determined by experimentation [32,80].
FePc L3-edge. FePc has been the object of a huge number of L2,3-edges XAS studies [33,34,38,120,124,125,126,127]. Among them, Bartolomé et al. [125] investigated the Fe magnetic moment switching in the catalytic ORR of FePc adsorbed on Ag(110) by combining the results of X-ray linear polarized absorption spectroscopy with those of X-ray magnetic circular dichroism at the Fe L2,3-edges. A detailed analysis of the || and ⊥ Fe L2,3-edges components of the FePc and FePc(η2-O2) XA spectra has been recently reported by Carlotto et al. [38] by adopting a computational set-up slightly different from that herein employed. The adopted set of c1, c2, and c3 semi-empirical parameters corresponded to that herein labelled old set; moreover, the saturation of the final-state manifold was obtained by considering forty nonrelativistic roots per multiplicity.
The comparison of the FePc ||/f(EE) distributions (see Figure S2 of the Supplementary Materials) with the literature’s theoretical results [38] proves an evident disagreement. To disentangle the effects induced by the adoption of a particular set of semi-empirical parameters from those generated by the number of nonrelativistic roots per multiplicity, the ||/f(EE) distributions have been again evaluated by using the c1, c2, and c3 old set (see Figure 4d). Besides minor differences, most likely due to the higher number of states and expansion vectors herein adopted in the iterative solution of the CI equations, the ||/f(EE) distributions obtained by adopting c1 = 0.21, c2 = 0.49, and c3 = 0.29 substantially mirror the literature’s ones [38].
According to Carlotto et al. [38], weighty contributions to the ||/f(EE) distributions arise from states having a spin multiplicity either equal to (ΔS = 0) or higher/lower (ΔS = ±1) than the GS one. More specifically, the ||/L31 features lying at the lowest excitation energy (see Figure 4d) are both associated to triplet states (ΔS = 0), and are generated by single electronic excitations with the whole Fe 2p set as isos and the low-lying Fe 3d-based 21a1g VMO and 6eg SOMO as fsos. Perfectly in agreement with the well-evidenced experimental dichroism [34,38], L31 is much more intense than ||L31, thus indicating that electronic states associated with (a2u → a1g)/(eu → eg) excitations contribute much more than the (a2u → eg)||/(eu → a1g)|| ones to the lower EE region of the L3-edge. Excitations with ΔS = 0, −1 comparably participate in states determining L32 (the theoretical set-up herein adopted makes negligible ΔS = 1 contributions). As such, ΔS = 0 electronic excitations involve the Fe-based 6eg and 21a1g VMO, while the ΔS = −1 ones have an MLCT character and imply Pc-based π* VMOs. Moving to the analysis of the ||L32 and ||L33 features, the comparison with the XAS evidence indicates that their excitation energies are slightly overestimated with respect to the ||L31 one; moreover, electronic states with ΔS = 0 (71%), 1 (19%) and −1 (7%) contribute to them. In even more detail, both single (44%) and coupled-single (56%) excitations contribute to the prevailing ΔS = 0 states. The former involves Fe-based (2p → 21a1g/6eg, 2p → 16b1g) and MLCT (2p → 7eg) transitions, while the latter are Fe-based (2p → SOMOs and SOMOs → 16b1g) excitations, determining the states contributing to the ||L32 and ||L33 lower excitation energy sides.
The DFT/ROCIS calculations [38] fairly reproduce both the L2–L3 ΔEE (~13 eV) and the corresponding relative intensities; nevertheless, any detailed assignment of the L2 feature is herein avoided as it is not unambiguously determined by experimentation [32,80].

4. Conclusions

The occupied and empty states of HS VPc, CrPc, IS MnPc and FePc have been thoroughly investigated by exploiting the original/homogeneous theoretical results and experimental evidence form the literature. The use of the Hirshfeld charges [84] coupled with the Nalewajski−Mrozek [66,67,68,69,70,71] approach ultimately indicates that, among the investigated molecules, FePc is characterized by the strongest and most covalent M–Pc σ interaction. Even though Slater’s transition state calculations ultimately confirm the Berkowitz hypothesis that, for FePc, “…the first ionization potential corresponds to electron ejection from a ring orbital, and not a metal orbital’’, the extension of the method to lighter MPcs reveals significant differences, the rationale of which lies with the relative energy position of the Pc2−-based and MII 3d-based occupied/half-occupied MOs. Insights into the MPcs’ virtual electronic structure have been gained by revisiting XAS data from the literature in light of DFT/ROCIS calculations. The higher EE side of the MPc L3-edge XA spectra systematically includes states associated with MLCT transitions, most of them involving the metal 2p → Pc2-based 7eg π* VMO excitations; moreover, the same EE region of all but one of (VPc) the L3-edge XA spectra is characterized by the presence of electronic states associated with M-based 2p → 16b1g excitations. The agreement between theory and experiment is satisfactory, but it required a “tuning” of the modelling set-up in terms of XC functionals and/or c1, c2, and c3 semiempirical parameters. As a final consideration, we underline that the theoretical outcomes obtained for the HS MPc (VPc and CrPc) are a true challenge for the experimental community called upon to confirm or deny them.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2079-4991/11/1/54/s1, Figure S1: 3D plot of the MnPc 6eg MO. Displayed isosurfaces correspond to ±0.015 e1/2 × A−3/2; Figure S2: ISFePc ||/⊥f(EE) distributions estimated by adopting c1 = 0.18, c2 = 0.20, and c3 = 0.40. Blue and red lines correspond to || and ⊥ components, respectively. Simulated spectra have a Gaussian broadening of 1.5 eV. Table S1: Optimized BP86 Cartesian coordinates for HS VPc; Table S2: Optimized BP86 Cartesian coordinates for HS CrPc; Table S3: Optimized BP86 Cartesian coordinates for IS MnPc; Table S4: Optimized BP86 Cartesian coordinates for IS FePc; Table S5: Symmetries of a 2p53d4/6 system; Table S6: Symmetries of a 2p53d5 system; Table S7: Symmetries of a 2p53d7 system.

Author Contributions

Methodology, analysis, investigation, and writing—original draft preparation: S.C., M.S., F.S. and A.V.; review and editing, supervision: M.C.; funding acquisition: S.C. and M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the University of Padova (Grant P-DISC #CARL-SID17 BIRD2017-UNIPD, Project CHIRoN).

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

All data have been illustrated in the manuscript and in the supplementary materials.

Acknowledgments

The Computational Chemistry Community (C3P) of the University of Padova is kindly acknowledged.

Conflicts of Interest

All authors of this manuscript directly participated in the planning, execution, and analysis of this study. The contents of this manuscript have not been copyrighted or published previously and are not currently under consideration for publication elsewhere. The contents of this manuscript will not be copyrighted, submitted, or published elsewhere after acceptance by NANOMATERIALS.

References

  1. Ghosh, A. Letters to a Young Chemist; John Wiley & Sons: Hoboken, NJ, USA, 2011; p. 34. [Google Scholar]
  2. Functional Phthalocyanine Molecular Materials; Springer: Berlin/Heidelberg, Germany, 2010; Volume 135.
  3. Wöhrle, D.; Schnurpfeil, G.; Makarov, S.G.; Kazarin, A.; Suvorova, O.N. Practical Applications of Phthalocyanines—From Dyes and Pigments to Materials for Optical, Electronic and Photo-electronic Devices. Macroheterocycles 2012, 5, 191–202. [Google Scholar] [CrossRef]
  4. Angione, M.D.; Pilolli, R.; Cotrone, S.; Magliulo, M.; Mallardi, A.; Palazzo, G.; Sabbatini, L.; Fine, D.; Dodabalapur, A.; Cioffi, N.; et al. Carbon based materials for electronic bio-sensing. Mater. Today 2011, 14, 424–433. [Google Scholar] [CrossRef]
  5. Jang, S.H.; Jen, A.K.Y. Structured Organic Non-Linear Optics. Compr. Nanosci. Technol. 2011, 1, 143–187. [Google Scholar]
  6. Nguyen, T. Polymer-based nanocomposites for organic optoelectronic devices. Surf. Coat. Technol. 2011, 206, 742–752. [Google Scholar] [CrossRef]
  7. Hains, A.W.; Liang, Z.; Woodhouse, M.A.; Gregg, B.A. Molecular Semiconductors in Organic Photovoltaic Cells. Chem. Rev. 2010, 110, 6689–6735. [Google Scholar] [CrossRef] [PubMed]
  8. Cao, W.; Xue, J. Recent progress in organic photovoltaics: Device architecture and optical design. Energy Environ. Sci. 2014, 7, 2123–2144. [Google Scholar] [CrossRef]
  9. Sorokin, A.B. Phthalocyanine Metal Complexes in Catalysis. Chem. Rev. 2013, 113, 8152–8191. [Google Scholar] [CrossRef]
  10. Tolman, W.B.; Solomon, E.I. Preface: Forum on Dioxygen Activation and Reduction. Inorg. Chem. 2010, 49, 3555–3556. [Google Scholar] [CrossRef]
  11. Scherson, D.A.; Palencsár, A.; Tolmachev, Y.; Stefan, I. Transition Metal Macrocycles as Electrocatalysts for Dioxygen Reduction. In Electrochemical Surface Modification: Thin Films, Functionalization and Characterization; Alkire, R.C., Kolb, D.M., Lipkowski, J., Ross, P.N., Eds.; Wiley-VCH: Weinheim, Germany, 2008; pp. 191–288. [Google Scholar]
  12. Sedona, F.; Di Marino, M.; Forrer, D.; Vittadini, A.; Casarin, M.; Cossaro, A.; Floreano, L.; Verdini, A.; Sambi, M. Tuning the Catalytic Activity of Ag(110)-supported Fe Phthalocyanine in the Oxygen Reduction Reaction. Nat. Mater. 2012, 11, 970–977. [Google Scholar] [CrossRef]
  13. Bianconi, A. XANES Spectroscopy. In X-ray Absorption: Principles, Applications, Techniques of EXAFS, SEXAFS and XANES; Koningsberger, D.C., Prins, R., Eds.; John Wiley & Sons: New York, NY, USA, 1988; pp. 573–662. [Google Scholar]
  14. Stöhr, J. NEXAFS Spectroscopy; Springer: Berlin, Germany, 1996. [Google Scholar]
  15. Zhang, H.H.; Hedman, B.; Hodgson, K.O. X-ray Absorption Spectroscopy and EXAFS Analysis. The Multiple-Scattering Method and Applications in Inorganic and Bioinorganic Chemistry. In Inorganic Electronic Structure and Spectroscopy; Solomon, E.I., Lever, A.B.P., Eds.; John Wiley & Sons: New York, NY, USA, 1999; Volume 1, pp. 513–554. [Google Scholar]
  16. Solomon, E.I.; Bell, C.B., III. Inorganic and Bioinorganic Spectroscopy. In Physical Inorganic Chemistry, Principle, Methods and Models; Bakac, A., Ed.; John Wiley & Sons: New York, NY, USA, 2010; pp. 1–37. [Google Scholar]
  17. Douglas, B.E.; Hollingsworth, C.A. Symmetry in Bonding and Spectra, an Introduction; Academic Press: Orlando, FL, USA, 1985. [Google Scholar]
  18. Neese, F.; Hedman, B.; Hodgson, K.O.; Solomon, E.I. Relationship between the Dipole Strength of Ligand Pre-Edge Transitions and Metal−Ligand Covalency. Inorg. Chem. 1999, 38, 4854–4860. [Google Scholar] [CrossRef]
  19. Glaser, T.; Hedman, B.; Hodgson, L.O.; Solomon, E.I. Ligand K-Edge X-ray Absorption Spectroscopy:  A Direct Probe of Ligand−Metal Covalency. Acc. Chem. Res. 2000, 33, 859–868. [Google Scholar] [CrossRef] [PubMed]
  20. Solomon, E.I.; Hedman, B.; Hodgson, K.O.; Dey, A.; Szilagyi, R.K. Ligand K-edge x-ray absorption spectroscopy: Covalency of ligand-metal bonds. Coord. Chem. Rev. 2005, 249, 97–129. [Google Scholar] [CrossRef]
  21. Baker, M.-L.; Mara, M.W.; Yan, J.J.; Hodgson, K.O.; Hedman, B.; Solomon, E.I. K- And L-edge X-ray Absorption Spectroscopy (XAS) and Resonant Inelastic X-ray Scattering (RIXS) Determination of Differential Orbital Covalency (DOC) of Transition Metal Sites. Coord. Chem. Rev. 2017, 345, 182–208. [Google Scholar] [CrossRef] [PubMed]
  22. Jørgensen, C.K. Absorption Spectra and Chemical Bonding in Complexes; Pergamon Press: Oxford, UK, 1962; p. 77. [Google Scholar]
  23. de Groot, F. Multiplet effects in X-ray spectroscopy. Coord. Chem. Rev. 2005, 249, 31–63. [Google Scholar] [CrossRef]
  24. de Groot, F.; Kotani, A. Core Level Spectroscopy of Solids; CRC Press: Boca Raton, FL, USA, 2008. [Google Scholar]
  25. Roemelt, M.; Neese, F. Excited States of Large Open-Shell Molecules: An Efficient, General, and Spin-Adapted Approach Based on a Restricted Open-Shell Ground State Wave function. J. Phys. Chem. A 2013, 117, 3069–3083. [Google Scholar] [CrossRef]
  26. Bagus, P.S.; Freund, H.; Kuhlenbeck, H.; Ilton, E.S. A new analysis of X-ray adsorption branching ratios: Use of Russell−Saunders coupling. Chem. Phys. Lett. 2008, 455, 331–334. [Google Scholar] [CrossRef]
  27. Ikeno, H.; Mizoguchi, T.; Tanaka, I. Ab initio charge transfer multiplet calculations on the L2,3 XANES and ELNES of 3d transition metal oxides. Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 83, 155107. [Google Scholar] [CrossRef] [Green Version]
  28. Josefsson, I.; Kunnus, K.; Schreck, S.; Föhlisch, A.; de Groot, F.; Wernet, P.; Odelius, M. Ab Initio calculations of x-ray spectra: Atomic multiplet and molecular orbital effects in a multiconfigurational SCF approach to the L-edge spectra of transition metal complexes. J. Phys. Chem. Lett. 2012, 3, 3565–3570. [Google Scholar] [CrossRef] [Green Version]
  29. Maganas, D.; Roemelt, M.; Hävecker, M.; Trunschke, A.; Knop-Gericke, A.; Schlögl, R.; Neese, F. First principles calculations of the structure and V L-edge X-ray absorption spectra of V2O5 using local pair natural orbital coupled cluster theory and spin−orbit coupled configuration interaction approaches. Phys. Chem. Chem. Phys. 2013, 15, 7260–7276. [Google Scholar] [CrossRef] [Green Version]
  30. Roemelt, M.; Maganas, D.; DeBeer, S.; Neese, F. A combined DFT and restricted open-shell configuration interaction method including spin-orbit coupling: Application to transition metal L-edge X-ray absorption spectroscopy. J. Chem. Phys. 2013, 138, 204101 (1:22). [Google Scholar] [CrossRef]
  31. Maganas, D.; DeBeer, S.; Neese, F. Restricted open-shell configuration interaction cluster calculations of the L-edge x-ray absorption study of TiO2 and CaF2 solids. Inorg. Chem. 2014, 53, 6374–6385. [Google Scholar] [CrossRef] [PubMed]
  32. Maganas, D.; Roemelt, M.; Weyhermüller, T.; Blume, R.; Hävecker, M.; Knop-Gericke, A.; DeBeer, S.; Schlögl, R.; Neese, F. L-edge X-ray absorption study of mononuclear vanadium complexes and spectral predictions using a restricted open shell configuration interaction ansatz. Phys. Chem. Chem. Phys. 2014, 16, 264–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Koshino, M.; Kurata, H.; Isoda, S.; Kobayashi, T. Branching ratio and L2 + L3 intensities of 3d-transition metals in phthalocyanines and the amine complexes. Micron 2000, 31, 373–380. [Google Scholar] [CrossRef]
  34. Kroll, T.; Kraus, R.; Schönfelder, R.; Aristov, V.Y.; Molodtsova, O.V.; Hoffmann, P.; Knupfer, M. Transition metal phthalocyanines: Insight into the electronic structure from soft x-ray spectroscopy. J. Chem. Phys. 2012, 137, 054306. [Google Scholar] [CrossRef]
  35. Eguchi, K.; Nakagawa, T.; Takagi, Y.; Yokoyama, T. Direct Synthesis of Vanadium Phthalocyanine and Its Electronic and Magnetic States in Monolayers and Multilayers on Ag(111). J. Phys. Chem. C 2015, 119, 9805–9815. [Google Scholar] [CrossRef]
  36. Carlotto, S.; Sambi, M.; Rancan, M.; Casarin, M. Theoretical Investigation of the Electronic Properties of Three Vanadium Phthalocyaninato (Pc) Based Complexes: PcV, PcVO, and PcVI. Inorg. Chem. 2018, 57, 1859–1869. [Google Scholar] [CrossRef]
  37. Casarin, M.; Carlotto, S. Pigments of Life”, Molecules Well Suited to Investigate Metal–Ligand Symmetry-Restricted Covalency. Eur. J. Inorg. Chem. 2018, 3145–3155. [Google Scholar] [CrossRef]
  38. Carlotto, S.; Sambi, M.; Sedona, F.; Vittadini, A.; Bartolomé, J.; Bartolomé, F.; Casarin, M. L2,3-edges absorption spectra of a 2D complex system: A theoretical modelling. Phys. Chem. Chem. Phys. 2016, 18, 28110–28116. [Google Scholar] [CrossRef]
  39. Nardi, M.V.; Detto, F.; Aversa, L.; Verucchi, R.; Salviati, G.; Iannotta, S.; Casarin, M. Electronic properties of CuPc and H2Pc: An experimental and theoretical study. Phys. Chem. Chem. Phys. 2013, 15, 12864–12881. [Google Scholar] [CrossRef]
  40. Mangione, G.; Sambi, M.; Nardi, M.V.; Casarin, M. A theoretical study of the L3 pre-edge XAS in Cu(II) complexes. Phys. Chem. Chem. Phys. 2014, 16, 19852–19855. [Google Scholar] [CrossRef]
  41. Mangione, G.; Sambi, M.; Carlotto, S.; Vittadini, A.; Ligorio, G.; Timpel, M.; Pasquali, L.; Giglia, A.; Nardi, M.V.; Casarin, M. Electronic structure of CuTPP and CuTPP(F) complexes: A combined experimental and theoretical study II. Phys. Chem. Chem. Phys. 2016, 18, 24890–24904. [Google Scholar] [CrossRef] [PubMed]
  42. Neese, F.; Wiley. The ORCA program system. Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
  43. Wang, F.; Ziegler, T. The calculation of excitation energies based on the relativistic two-component zeroth-order regular approximation and time-dependent density-functional with full use of symmetry. J. Chem. Phys. 2005, 122, 204103. [Google Scholar] [CrossRef] [PubMed]
  44. Hirata, S.; Head-Gordon, M. Time-dependent density functional theory within the Tamm–Dancoff approximation. Chem. Phys. Lett. 1999, 314, 291–299. [Google Scholar] [CrossRef]
  45. Carlotto, S.; Floreano, L.; Cossaro, A.; Dominguez, M.; Rancan, M.; Sambi, M.; Casarin, M. The electronic properties of three popular high spin complexes [TM(acac)3, TM = Cr, Mn, and Fe] revisited: An experimental and theoretical study. Phys. Chem. Chem. Phys. 2017, 19, 24840–24854. [Google Scholar] [CrossRef]
  46. Casarin, M.; Finetti, P.; Vittadini, A.; Wang, F.; Ziegler, T. Spin-Orbit Relativistic Time-Dependent Density Functional Calculations of the Metal and Ligand Pre-Edge XAS Intensities of Organotitanium Complexes: TiCl4, Ti(η5-C5H5)Cl3, and Ti(η5-C5H5)2Cl2. J. Phys. Chem. A 2007, 111, 5270–5279. [Google Scholar] [CrossRef]
  47. Carlotto, S.; Finetti, P.; de Simone, M.; Coreno, M.; Casella, G.; Sambi, M.; Casarin, M. Comparative Experimental and Theoretical Study of the C and O K-Edge X-ray Absorption Spectroscopy in Three Highly Popular, Low Spin Organoiron Complexes: [Fe(CO)5], [(η5-C5H5)Fe(CO)(μ-CO)]2, and [(η5-C5H5)2Fe]. Inorg. Chem. 2019, 58, 16411–16423. [Google Scholar] [CrossRef]
  48. Carlotto, S.; Finetti, P.; de Simone, M.; Coreno, M.; Casella, G.; Sambi, M.; Casarin, M. Comparative Experimental and Theoretical Study of the Fe L2,3-Edges X-ray Absorption Spectroscopy in Three Highly Popular, Low Spin Organoiron Complexes: [Fe(CO)5], [(η5-C5H5)Fe(CO)(μ-CO)]2, and [(η5-C5H5)2Fe]. Inorg. Chem. 2019, 58, 5844–5857. [Google Scholar] [CrossRef]
  49. Wu, W.; Harrison, N.M.; Fisher, A.J. Suitability of chromium phthalocyanines to test Haldane’s conjecture: First-principles calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 88, 224417. [Google Scholar] [CrossRef]
  50. Li, Y.; Sun, Q. The superior catalytic CO oxidation capacity of a Cr-phthalocyanine porous sheet. Sci. Rep. 2014, 4, 4098. [Google Scholar] [CrossRef] [Green Version]
  51. Junkaew, A.; Meeprasert, J.; Jansang, B.; Kungwan, N.; Namuangruk, S. Mechanistic study of NO oxidation on Cr–phthalocyanine: Theoretical insight. RSC Adv. 2017, 7, 8858–8865. [Google Scholar] [CrossRef] [Green Version]
  52. Liu, K.; Lei, Y.; Wang, G. Correlation between oxygen adsorption energy and electronic structure of transition metal macrocyclic complexes. J. Chem. Phys. 2013, 139, 204306. [Google Scholar] [CrossRef] [PubMed]
  53. Bala, M.; Verma, P.K.; Sharma, U.; Kumar, N.; Singh, B. Iron phthalocyanine as an efficient and versatile catalyst for N-alkylation of heterocyclic amines with alcohols: One-pot synthesis of 2-substituted benzimidazoles, benzothiazoles and benzoxazoles. Green Chem. 2013, 15, 1687–1693. [Google Scholar] [CrossRef]
  54. Paradine, S.M.; White, M.C. Iron-Catalyzed Intramolecular Allylic C–H Amination. J. Am. Chem. Soc. 2012, 134, 2036–2039. [Google Scholar] [CrossRef] [PubMed]
  55. Lamar, A.A.; Nicholas, K.M. Direct synthesis of 3-arylindoles via annulation of aryl hydroxylamines with alkynes. Tetrahedron 2009, 65, 3829–3833. [Google Scholar] [CrossRef]
  56. Taniguchi, T.; Hirose, D.; Ishibashi, H. Esterification via Iron-Catalyzed Activation of Triphenylphosphine with Air. ACS Catal. 2011, 1, 1469–1474. [Google Scholar] [CrossRef]
  57. Prateeptongkum, S.; Jovel, I.; Jackstell, R.; Vogl, N.; Weckbecker, C.; Beller, M. First iron-catalyzed synthesis of oximes from styrenes. Chem. Commun. 2009, 1990–1992. [Google Scholar] [CrossRef]
  58. Kudrik, E.; Makarov, S.V.; Zahl, A.; van Eldik, R. Mechanism of the Iron Phthalocyanine Catalyzed Reduction of Nitrite by Dithionite and Sulfoxylate in Aqueous Solution. Inorg. Chem. 2005, 44, 6470–6475. [Google Scholar] [CrossRef]
  59. Sugimori, T.; Horike, S.I.; Tsumura, S.; Hande, M.; Kasuga, K. Catalytic oxygenation of olefin with dioxygen and tetra-t-butylphthalocyanine complexes in the presence of sodium borohydride. Inorg. Chim. Acta 1998, 283, 275–278. [Google Scholar] [CrossRef]
  60. Leggans, E.K.; Barker, T.J.; Duncan, K.K.; Boger, D.L. Iron(III)/NaBH4-Mediated Additions to Unactivated Alkenes: Synthesis of Novel 20′-Vinblastine Analogues. Org. Lett. 2012, 14, 1428–1431. [Google Scholar] [CrossRef] [Green Version]
  61. Taniguchi, T.; Goto, N.; Nishibata, A.; Ishibashi, H. Iron-Catalyzed Redox Radical Cyclizations of 1,6-Dienes and Enynes. Org. Lett. 2010, 12, 112–115. [Google Scholar] [CrossRef] [PubMed]
  62. Auwärter, W.; Écija, D.; Klappenberger, E.; Barth, J.V. Porphyrins at interfaces. Nat. Chem. 2015, 7, 105–120. [Google Scholar] [CrossRef] [PubMed]
  63. ADF2014, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands. Available online: http://www.scm.com (accessed on 28 December 2020).
  64. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behaviour. Phys. Rev. A: At. Mol. Opt. Phys. 1988, 38, 3098–3100. [Google Scholar] [CrossRef] [PubMed]
  65. Perdew, J.P. Density functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B Condens. Matter Mater. Phys. 1986, 33, 8822–8824. [Google Scholar] [CrossRef]
  66. Nalewajski, R.F.; Mrozek, J. Modified valence indices from the two-particle density matrix. Int. J. Quantum Chem. 1994, 51, 187–200. [Google Scholar] [CrossRef]
  67. Nalewajski, R.F.; Mrozek, J.; Formosinho, S.J.; Varandas, A.J.C. Quantum mechanical valence study of a bond-breaking−bond forming process in triatomic systems. Int. J. Quantum Chem. 1994, 52, 1153–1176. [Google Scholar] [CrossRef]
  68. Nalewajski, R.F.; Mrozek, J. Hartree-Fock difference approach to chemical valence: Three-electron indices in UHF approximation. Int. J. Quantum Chem. 1996, 57, 377–389. [Google Scholar] [CrossRef]
  69. Nalewajski, R.F.; Mrozek, J.; Mazur, G. Quantum chemical valence indices from the one-determinantal difference approach. Can. J. Chem. 1996, 74, 1121–1130. [Google Scholar] [CrossRef]
  70. Nalewajski, R.F.; Mrozek, J.; Michalak, A. Two-electron valence indices from the Kohn-Sham orbitals. Int. J. Quantum Chem. 1997, 61, 589–601. [Google Scholar] [CrossRef]
  71. Nalewajski, R.F.; Mrozek, J.; Michalak, A. Exploring bonding patterns of molecular systems using quantum mechanical bond multiplicities. Polym. J. Chem. 1998, 72, 1779–1791. [Google Scholar]
  72. Michalak, A.; DeKock, R.L.; Ziegler, T. Bond Multiplicity in Transition-Metal Complexes: Applications of Two-Electron Valence Indices. J. Phys. Chem. A 2008, 112, 7256–7263. [Google Scholar] [CrossRef]
  73. Ziegler, T.; Rauk, A. On the calculation of bonding energies by the Hartree Fock Slater method. Theor. Chim. Acta 1977, 46, 1–10. [Google Scholar] [CrossRef]
  74. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  75. Zhao, Y.; Truhlar, D.J. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  76. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef] [PubMed]
  77. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [Google Scholar] [CrossRef]
  78. Pantazis, D.A.; Chen, X.Y.; Landis, C.R.; Neese, F. All-electron scalar relativistic basis sets for third-row transition metal atoms. J. Chem. Theory Comput. 2008, 4, 908–919. [Google Scholar] [CrossRef]
  79. Lebedev, V.I. Values of the nodes and weights of quadrature formulas of Gauss−Markov type for a sphere from the ninth to seventeenth order of accuracy that are invariant with respect to an octahedron group with inversion. Zhurnal Vychislitel’noi Mat. Mat. Fiz. 1975, 15, 48–54. [Google Scholar]
  80. Coster, D.; Kronig, R.D.L. A new type of Auger effect and its influence on the X-ray spectrum. Physica 1935, 2, 13–24. [Google Scholar] [CrossRef]
  81. Mangione, G.; Carlotto, S.; Sambi, M.; Ligorio, G.; Timpel, M.; Vittadini, A.; Nardi, M.V.; Casarin, M. Electronic structures of CuTPP and CuTPP(F) complexes. A combined experimental and theoretical study I. Phys. Chem. Chem. Phys. 2016, 18, 18727–18738. [Google Scholar] [CrossRef]
  82. Arillo-Flores, O.I.; Fadlallah, M.M.; Schuster, C.; Eckern, U.; Romero, A.H. Magnetic, electronic and vibrational properties of metal and fluorinated metal phthalocyanines. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 87, 165115. [Google Scholar] [CrossRef] [Green Version]
  83. Jahn, H.; Teller, E. Stability of polyatomic molecules in degenerate electronic states-I—Orbital degeneracy. Proc. R. Soc. London, Ser. A 1937, 161, 220–235. [Google Scholar]
  84. Hirshfeld, F.L. Bonded-atom fragments for describing molecular charge densities. Theor. Chim. Acta 1977, 44, 129–138. [Google Scholar] [CrossRef]
  85. Slater, J.C. Quantum Theory of Molecules and Solids. The Self-Consistent-Field for Molecules and Solids; McGraw-Hill: New York, NY, USA, 1974. [Google Scholar]
  86. Liberman, D.A. Slater transition-state band-structure calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 2000, 62, 6851–6853. [Google Scholar] [CrossRef]
  87. Berkowitz, J. Photoelectron spectroscopy of phthalocyanine vapors. J. Chem. Phys. 1979, 70, 2819–2828. [Google Scholar] [CrossRef]
  88. A comprehensive collection of single-crystal structures of phthalocyanine complexes and related macrocycles is reported in M.K. Engel. In The Porphyrin Handbook; Phthalocyanines: Structural Characterization; Academic Press: New York, NY, USA, 2003; Volume 20, pp. 1–242.
  89. Elvidge, J.A.; Lever, B.P. Metal chelates. Part II. Phthalocyanine–chromium complexes and perpendicular conjugation. J. Chem. Soc. 1961, 1257–1265. [Google Scholar] [CrossRef]
  90. Lever, A.B.P. The magnetic behaviour of transition-metal phathalocyanines. J. Chem. Soc. 1965, 1821–1829. [Google Scholar] [CrossRef]
  91. Soler, J.M.; Artacho, E.; Gale, J.D.; García, A.; Junquera, J.; Ordejón, P.; Sánchez-Portal, D. The SIESTA method for ab initio order-N materials simulation. J. Phys. Condens. Matter 2002, 14, 2745–2779. [Google Scholar] [CrossRef] [Green Version]
  92. Figgis, B.N.; Mason, R.; Williams, G.A. Structure of Phthalocyaninatomanganese(II) at 5.8 K Determined by Neutron Diffraction. Acta Cryst. 1980, B36, 2963–2970. [Google Scholar] [CrossRef]
  93. Kirner, J.F.; Dow, D.; Scheidt, W.R. Molecular Stereochemistry of Two Intermediate-Spin Complexes. Iron(II) Phthalocyanine and Manganese(II) Phthalocyanine. Inorg. Chem. 1976, 15, 1685–1690. [Google Scholar] [CrossRef]
  94. Barraclough, C.G.; Martin, R.L.; Mitra, S.; Sherwood, R.C. Paramagnetic Anisotropy, Electronic Structure, and Ferromagnetism in Spin S = 3/2 Manganese(II) Phthalocyanine. J. Chem. Phys. 1970, 53, 1638–1642. [Google Scholar] [CrossRef]
  95. Mitra, S.; Greson, A.K.; Hatfield, W.E.; Weller, R.R. Single-Crystal Magnetic Study on Ferromagnetic Manganese(II) Phthalocyaninate. Inorg. Chem. 1983, 22, 1729–1732. [Google Scholar] [CrossRef]
  96. Williamson, B.E.; van Cott, T.C.; Boyle, M.E.; Misener, C.G.; Stillman, M.J.; Schatz, P.N. Determination of the Ground State of Manganese Phthalocyanine in an Argon Matrix Using Magnetic Circular Dichroism and Absorption Spectroscopy. J. Am. Chem. Soc. 1992, 114, 2412. [Google Scholar] [CrossRef]
  97. Taguchi, Y.; Miyake, T.; Margadonna, S.; Kato, K.; Prassides, K.; Iwasa, Y. Synthesis, Structure, and Magnetic Properties of Li-Doped Manganese-Phthalocyanine, Lix[MnPc] (0 ≤ x ≤ 4). J. Am. Chem. Soc. 2006, 128, 3313. [Google Scholar] [CrossRef]
  98. Petraki, F.; Peisert, H.; Hoffmann, P.; Uihlein, J.; Knupfer, M.; Chassé, T. Modification of the 3d-Electronic Configuration of Manganese Phthalocyanine at the Interface to Gold. J. Phys. Chem. C 2012, 116, 5121–5127. [Google Scholar] [CrossRef]
  99. Kataoka, T.; Sakamoto, Y.; Yamazaki, Y.; Singh, V.R.; Fujimori, A.; Takeda, Y.; Ohkochic, T.; Fujimori, S.I.; Okane, T.; Saitoh, Y.; et al. Electronic configuration of Mn ions in the π-d molecular ferromagnet β-Mn phthalocyanine studied by soft X-ray magnetic circular dichroism. Solid State Commun. 2012, 152, 806–809. [Google Scholar] [CrossRef] [Green Version]
  100. Gottfrield, J.M. Surface chemistry of porphyrins and phthalocyanines. Surf. Sci. Reports 2015, 70, 259–379. [Google Scholar] [CrossRef]
  101. Liao, M.S.; Scheiner, S. Electronic structure and bonding in metal phthalocyanines, Metal = Fe, Co, Ni, Cu, Zn, Mg. J. Chem. Phys. 2001, 114, 9780–9791. [Google Scholar] [CrossRef] [Green Version]
  102. Liao, M.S.; Watts, J.D.; Huang, M.J. DFT Study of Unligated and Ligated ManganeseII Porphyrins and Phthalocyanines. Inorg. Chem. 2005, 44, 1941–1949. [Google Scholar] [CrossRef]
  103. Marom, N.; Kronik, L. Density functional theory of transition metal phthalocyanines, II: Electronic structure of MnPc and FePc—symmetry and symmetry breaking. Appl. Phys. A 2009, 95, 165–172. [Google Scholar] [CrossRef]
  104. Shen, X.; Sun, L.; Benassi, E.; Shen, Z.; Zhao, X.; Sanvito, S.; Hou, S. Spin filter effect of manganese phthalocyanine contacted with single-walled carbon nanotube electrodes. J. Chem. Phys. 2010, 132, 054703. [Google Scholar] [CrossRef]
  105. Brumboiu, I.E.; Totani, R.; de Simone, M.; Coreno, M.; Grazioli, C.; Lozzi, L.; Herper, H.C.; Sanyal, B.; Eriksson, O.; Pugli, C.; et al. Elucidating the 3d Electronic Configuration in Manganese Phthalocyanine. J. Phys. Chem. A 2014, 118, 927–932. [Google Scholar] [CrossRef] [Green Version]
  106. Wäckerlin, C.; Donati, F.; Singha, A.; Baltic, R.; Uldry, A.C.; Delley, B.; Rusponi, S.; Dreiser, J. Strong antiferromagnetic exchange between manganese phthalocyanine and ferromagnetic europium oxide. Chem. Comm. 2015, 51, 12958–12961. [Google Scholar] [CrossRef] [Green Version]
  107. Wang, J.; Shi, Y.; Cao, J.; Wu, R. Magnetization and magnetic anisotropy of metallophthalocyanine molecules from the first principles calculations. Appl. Phys. Lett. 2009, 94, 122502. [Google Scholar] [CrossRef]
  108. Grobosch, M.; Mahns, B.; Loose, C.; Friedrich, R.; Schmidt, C.; Kortus, J.; Knupfer, M. Identification of the electronic states of manganese phthalocyanine close to the Fermi level. Chem. Phys. Lett. 2011, 505, 122–125. [Google Scholar] [CrossRef]
  109. Grobosch, M.; Aristov, V.Y.; Molodtsova, O.V.; Schmidt, C.; Doyle, B.P.; Nannarone, S.; Knupfer, M. Engineering of the Energy Level Alignment at Organic Semiconductor Interfaces by Intramolecular Degrees of Freedom: Transition Metal Phthalocyanines. J. Phys. Chem. C 2009, 113, 13219–13222. [Google Scholar] [CrossRef]
  110. Dale, B.W.; Williams, R.J.P.; Johnson, C.E.; Thorp, T.L. S=1 Spin State of Divalent Iron. I. Magnetic Properties of Phthalocyanine Iron (II). J. Chem. Phys. 1968, 49, 3441–3444. [Google Scholar] [CrossRef]
  111. Dale, B.W.; Williams, R.J.P.; Johnson, C.E.; Thorp, T.L. S=1 Spin State of Divalent Iron. II. A Mössbauer-Effect Study of Phthalocyanine Iron (II). J. Chem. Phys. 1968, 49, 3445–3449. [Google Scholar] [CrossRef]
  112. Coppens, P.; Li, L.; Zhu, N.J. Electronic ground state of iron(II) phthalocyanine as determined from accurate diffraction data. J. Am. Chem. Soc. 1983, 105, 6173–6174. [Google Scholar] [CrossRef]
  113. Evangelisti, M.; Bartolomé, J.; de Jongh, L.J.; Filoti, G. Magnetic properties of α-iron(II) phthalocyanine. Phys. Rev. B Condens. Matter Mater. Phys. 2002, 66, 144410. [Google Scholar] [CrossRef] [Green Version]
  114. Filoti, G.; Kuz’min, M.D.; Bartolomé, J. Mössbauer study of the hyperfine interactions and spin dynamics in α-iron(II) phthalocyanin. Phys. Rev. B Condens. Matter Mater. Phys. 2006, 74, 134420. [Google Scholar] [CrossRef] [Green Version]
  115. Liao, M.S.; Scheiner, S. Comparative study of metal-porphyrins, -porphyrazines, and -phthalocyanines. J. Comput. Chem. 2002, 23, 1391–1403. [Google Scholar] [CrossRef]
  116. Mangione, G.; Pandolfo, L.; Sambi, M.; Ligorio, G.; Nardi, M.V.; Cossaro, A.; Floreano, L.; Casarin, M. Ligand-Field Strength and Symmetry-Restricted Covalency in Cu II Complexes—A Near-Edge X-ray Absorption Fine Structure Spectroscopy and Time-Dependent DFT Study. Eur. J. Inorg. Chem. 2015, 2709–2713. [Google Scholar]
  117. Nardi, M.V.; Verucchi, R.; Pasquali, L.; Giglia, A.; Fronzoni, G.; Sambi, M.; Mangione, G.; Casarin, M. XAS of tetrakis(phenyl)- and tetrakis (pentafluorophenyl)-porphyrin: An experimental and theoretical study. Phys. Chem. Chem. Phys. 2015, 17, 2001–2011. [Google Scholar] [CrossRef]
  118. Carlotto, S.; Sambi, M.; Vittadini, A.; Casarin, M. Theoretical modelling of the L2,3-edge X-ray absorption spectra of Mn(acac)2 and Co(acac)2 complexes. Phys. Chem. Chem. Phys. 2016, 18, 2242–2249. [Google Scholar] [CrossRef]
  119. Carlotto, S.; Casella, G.; Floreano, L.; Verdini, A.; Ribeiro, A.P.C.; Martins, L.M.D.R.S.; Casarin, M. Spin state, electronic structure and bonding on C-scorpionate [Fe(II)Cl2(tpm)] catalyst: An experimental and computational study. Catal. Tod. 2020, 358, 403–411. [Google Scholar] [CrossRef]
  120. Cojocariu, I.; Carlotto, S.; Sturmeit, H.M.; Zamborlini, G.; Cinchetti, M.; Cossaro, A.; Verdini, A.; Floreano, L.; Jugovac, M.; Puschnig, P.; et al. Ferrous to ferric transition in Fe-phthalocyanine driven by NO2 exposure. Chem.-Eur. J. in press. [CrossRef]
  121. Ballhausen, C.J. Introduction to Ligand Field Theory; McGraw-Hill Book Company, Inc.: New York, NY, USA, 1962; p. 69. [Google Scholar]
  122. Sugano, S.; Tanabe, Y.; Kamimura, H. Multiplets of Transition-Metal Ions in Crystals; Academic Press: New York, NY, USA, 1970. [Google Scholar]
  123. Carlotto, S.; Sambi, M.; Vittadini, A.; Casarin, M. Mn(acac)2 and Mn(acac)3 complexes, a theoretical modelling of their L2,3-edges X-ray absorption spectra. Polyhedron 2017, 135, 216–223. [Google Scholar] [CrossRef]
  124. Simonov, K.A.; Vinogradov, A.S.; Brzhezinskaya, M.M.; Preobrajenski, A.B.; Generalov, A.V.; Yu Klyushin, A. Features of metal atom 2p excitations and electronic structure of 3d-metal phthalocyanines studied by X-ray absorption and resonant photoemission. Appl. Surf. Sci. 2013, 267, 132–135. [Google Scholar] [CrossRef]
  125. Bartolomé, J.; Bartolomé, F.; Brookes, N.B.; Sedona, F.; Basagni, A.; Forrer, D.; Sambi, M. Reversible Fe Magnetic Moment Switching in Catalytic Oxygen Reduction Reaction of Fe-Phthalocyanine Adsorbed on Ag(110). J. Phys. Chem. C 2015, 119, 12488–12495. [Google Scholar] [CrossRef]
  126. Betti, M.G.; Gargiani, P.; Frisenda, R.; Biagi, R.; Cossaro, A.; Verdini, A.; Floreano, L.; Mariani, C. Localized and Dispersive Electronic States at Ordered FePc and CoPc Chains on Au(110). J. Phys. Chem. C 2010, 114, 21638–21644. [Google Scholar] [CrossRef]
  127. Stepanow, S.; Lodi Rizzini, A.; Krull, C.; Kavich, J.; Cezar, J.C.; Yakhou-Harris, F.; Sheverdyaeva, P.M.; Moras, P.; Carbone, C.; Ceballos, G.; et al. Spin Tuning of Electron-Doped Metal–Phthalocyanine Layers. J. Am. Chem. Soc. 2014, 136, 5451–5459. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of a D4h MPc molecule with the atom numbering recommended by the International Union of Pure and Applied Chemistry (IUPAC). White, grey, blue and yellow spheres correspond to H, C, N, and M atoms, respectively. In the adopted framework, the molecular σh plane corresponds to the xy plane. N(29), N(30), N(31) and N(32) (N(6), N(13), N(20) and N(27)) are collectively labelled NPy (Nm).
Figure 1. Schematic representation of a D4h MPc molecule with the atom numbering recommended by the International Union of Pure and Applied Chemistry (IUPAC). White, grey, blue and yellow spheres correspond to H, C, N, and M atoms, respectively. In the adopted framework, the molecular σh plane corresponds to the xy plane. N(29), N(30), N(31) and N(32) (N(6), N(13), N(20) and N(27)) are collectively labelled NPy (Nm).
Nanomaterials 11 00054 g001
Figure 2. Relative energy positions of VPc (a), CrPc (b), MnPc (c) and FePc (d) frontier MOs. Black (↑)/red (↓) arrows refer to Pc-based selected orbitals, while the blue ones correspond to the M 3d-based MOs.
Figure 2. Relative energy positions of VPc (a), CrPc (b), MnPc (c) and FePc (d) frontier MOs. Black (↑)/red (↓) arrows refer to Pc-based selected orbitals, while the blue ones correspond to the M 3d-based MOs.
Nanomaterials 11 00054 g002
Figure 3. Three-dimensional plots of the VPc and CrPc 6eg MO. Displayed isosurfaces correspond to ± 0.015 e1/2A−3/2.
Figure 3. Three-dimensional plots of the VPc and CrPc 6eg MO. Displayed isosurfaces correspond to ± 0.015 e1/2A−3/2.
Nanomaterials 11 00054 g003
Figure 4. Experimental (dotted lines) and calculated (solid lines) ||/f(EE) distributions for HSVPc (a), HSCrPc (b), ISMnPc (c) and ISFePc (d). Blue and red lines correspond to || and ⊥ components, respectively. Simulated spectra have been shifted by 10.9 (HSVPc), 9.8 (HSCrPc), 8.3 (ISMnPc) and 13.1 (ISFePc) eV and have a Gaussian broadening of 1.8 (HSVPc), 2.0 (HSCrPc), 1.0 (ISMnPc) and 1.0 (ISFePc) eV. Only the total experimental spectrum (black dotted line) is available for CrPc [33]; MnPc ||/f(EE) distributions have been obtained by using the hybrid M06 meta-GGA XC [75] (see text).
Figure 4. Experimental (dotted lines) and calculated (solid lines) ||/f(EE) distributions for HSVPc (a), HSCrPc (b), ISMnPc (c) and ISFePc (d). Blue and red lines correspond to || and ⊥ components, respectively. Simulated spectra have been shifted by 10.9 (HSVPc), 9.8 (HSCrPc), 8.3 (ISMnPc) and 13.1 (ISFePc) eV and have a Gaussian broadening of 1.8 (HSVPc), 2.0 (HSCrPc), 1.0 (ISMnPc) and 1.0 (ISFePc) eV. Only the total experimental spectrum (black dotted line) is available for CrPc [33]; MnPc ||/f(EE) distributions have been obtained by using the hybrid M06 meta-GGA XC [75] (see text).
Nanomaterials 11 00054 g004
Table 1. BP86 ΔBE (kcal/mol) of optimized D4h VPc, CrPc, MnPc and FePc with different spin states. GSs are taken as reference.
Table 1. BP86 ΔBE (kcal/mol) of optimized D4h VPc, CrPc, MnPc and FePc with different spin states. GSs are taken as reference.
SVPcCrPcMnPcFePc
0 NC a 31.9 (1A1g) b
1/216.6 (2B1g) 14.9 (2B2u)
1 25.6b (3Eu) 0 (3A2g)90
3/20 (4Eg) 0 (4Eg)
2 0 (5B1g) NCa
5/2 16.5 (6Eg)
a NC stands for non-converged; b Non-Aufbau.
Table 2. Theoretical and experimental [88] (in parentheses) structural parameters for D4h VPc, CrPc, MnPc and FePc. Bond lengths/bond angles in Å/°, respectively.
Table 2. Theoretical and experimental [88] (in parentheses) structural parameters for D4h VPc, CrPc, MnPc and FePc. Bond lengths/bond angles in Å/°, respectively.
M–NPyNPy–CNm–CM–NPy–CNPy–C–Nm
VPc *1.9961.3921.333125.4127.4
CrPc *1.9821.3871.330125.6127.5
MnPc a1.952
(1.938)
1.396
(1.392)
1.324
(1.315)
126.1
(126.2)
127.4
(127.6)
FePc b1.935
(1.927)
1.393
(1.378)
1.321
(1.322)
126.4
(126.3)
127.3
(127.8)
* Neither VPc nor CrPc crystallographic data are available in the literature; a from Ref. [92]; b from Ref. [93].
Table 3. BP86 Δ E o r b χ (kcal/mol) of GS D4h MPc herein considered (FePc is taken as a reference).
Table 3. BP86 Δ E o r b χ (kcal/mol) of GS D4h MPc herein considered (FePc is taken as a reference).
VPcCrPcMnPcFePc
a1g100.9106.7103.80.0
b1g−86.8−85.9−26.80.0
b2g81.699.686.20.0
eg−16.7−55.0−153.90.0
ΔEorb79.064.89.30.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Carlotto, S.; Sambi, M.; Sedona, F.; Vittadini, A.; Casarin, M. A Theoretical Study of the Occupied and Unoccupied Electronic Structure of High- and Intermediate-Spin Transition Metal Phthalocyaninato (Pc) Complexes: VPc, CrPc, MnPc, and FePc. Nanomaterials 2021, 11, 54. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11010054

AMA Style

Carlotto S, Sambi M, Sedona F, Vittadini A, Casarin M. A Theoretical Study of the Occupied and Unoccupied Electronic Structure of High- and Intermediate-Spin Transition Metal Phthalocyaninato (Pc) Complexes: VPc, CrPc, MnPc, and FePc. Nanomaterials. 2021; 11(1):54. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11010054

Chicago/Turabian Style

Carlotto, Silvia, Mauro Sambi, Francesco Sedona, Andrea Vittadini, and Maurizio Casarin. 2021. "A Theoretical Study of the Occupied and Unoccupied Electronic Structure of High- and Intermediate-Spin Transition Metal Phthalocyaninato (Pc) Complexes: VPc, CrPc, MnPc, and FePc" Nanomaterials 11, no. 1: 54. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11010054

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop