Next Article in Journal
Analysis of Adhesion at the Interface of Steamed Bread and Eggshell
Next Article in Special Issue
Structure and Properties of Heterometallics Based on Lanthanides and Transition Metals with Methoxy-β-Diketonates
Previous Article in Journal
Effects of EOs vs. Antibiotics on E. coli Strains Isolated from Drinking Waters of Grazing Animals in the Upper Molise Region, Italy
Previous Article in Special Issue
Iron Porphyrin as a Cytochrome P450 Model for the Degradation of Dye
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Charge Transfer Chromophores Derived from 3d-Row Transition Metal Complexes

by
Kira I. Pashanova
1,*,
Irina V. Ershova
1,
Olesya Yu. Trofimova
1,
Roman V. Rumyantsev
1,
Georgy K. Fukin
1,
Artem S. Bogomyakov
2,
Maxim V. Arsenyev
1 and
Alexandr V. Piskunov
1,*
1
Laboratory of Metal Complexes with Redox-Active Ligands, G.A. Razuvaev Institute of Organometallic Chemistry, Russian Academy of Sciences, 49 Tropinina Street, 603137 Nizhny Novgorod, Russia
2
International Tomography Center, Siberian Branch of the Russian Academy of Sciences, 3a Institutskaya Street, 630090 Novosibirsk, Russia
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(23), 8175; https://doi.org/10.3390/molecules27238175
Submission received: 3 November 2022 / Revised: 18 November 2022 / Accepted: 20 November 2022 / Published: 24 November 2022
(This article belongs to the Special Issue Applications of Metal Complexes)

Abstract

:
A series of new charge transfer (CT) chromophores of “α-diimine-MII-catecholate” type (where M is 3d-row transition metals—Cu, Ni, Co) were derived from 4,4′-di-tert-butyl-2,2′-bipyridyl and 3,6-di-tert-butyl-o-benzoquinone (3,6-DTBQ) in accordance with three modified synthetic approaches, which provide high yields of products. A square-planar molecular structure is inherent for monomeric [CuII(3,6-Cat)(bipytBu)]∙THF (1) and NiII(3,6-Cat)(bipytBu) (2) chromophores, while dimeric complex [CoII(3,6-Cat)(bipytBu)]2∙toluene (3) units two substantially distorted heteroleptic D-MII-A (where D, M, A are donor, metal and acceptor, respectively) parts through a donation of oxygen atoms from catecholate dianions. Chromophores 13 undergo an effective photoinduced intramolecular charge transfer (λ = 500–715 nm, extinction coefficient up to 104 M−1·cm−1) with a concomitant generation of a less polar excited species, the energy of which is a finely sensitive towards solvent polarity, ensuring a pronounced negative solvatochromic effect. Special attention was paid to energetic characteristics for CT and interacting HOMO/LUMO orbitals that were explored by a synergy of UV-vis-NIR spectroscopy, cyclic voltammetry, and DFT study. The current work sheds light on the dependence of CT peculiarities on the nature of metal centers from various groups of the periodic law. Moreover, the “α-diimine-MII-catecholate” CT chromophores on the base of “late” transition elements with differences in d-level’s electronic structure were compared for the first time.

1. Introduction

Originating in ancient time as a household craft [1], dyes chemistry has received well-deserved recognition and the widest industrial application, working on the diverse needs of society and making our life unthinkable without the existence of colorants (textile, food, cosmetics, paint and varnish manufacturing, etc.). At present, a search for new effective colorants (dyes or chromophoric systems), corresponding to newly emerging challenges, remains one of the most demanded chemical fields, which involves a numerous range of different technologies to generate and design a huge set of chromophoric compounds and materials [2].
In according with K. Müllen and co-workers [3], “functional dyes” should be considered rightfully as a pinnacle in a development of modern colorant chemistry: such dyes are intended to be implied in so-called “high-level photonic applications” for “smart” life, depending on the absorption/emission peculiarities of chromophores, e.g., photovoltaics (photosensitizers in solar cells) [4,5,6,7,8,9,10], optics (optical fibers, liquid-crystal materials) [11,12,13], luminescent-based (bioimaging, LED devices) [14,15,16,17,18], and electrochromic-based directions (e.g., “smart” mirror systems) [19,20,21,22,23,24,25,26,27,28] (references are mainly given on charge transfer complexes, see below).
The well-known and most effective to date chromophores can be classified [29] by their purpose, and by the mechanism of absorption/emission and ways to manage them [2]. In particular, avant-garde trends in working-out luminescent labels, which are indispensable in bioimaging, involve two types of fundamentally different chromophoric systems, quantum dots (QDs) and molecular dyes [29]. Unlike size-dependent absorption/emission of QDs, electron transition energy towards classical molecular colorants (e.g., organics, metal complexes) is available for a fine-tuning through the molecular design. Along with “resonant dyes” [29], displaying resonant absorption/emission owing to a delocalization of optical transitions over the molecular chromophoric system, the colorant with intramolecular charge transfer (CT dyes) deserve special attention from researchers. The sound advantages of CT chromophores [29] are the presence of high-intensive absorption/emission bands, as well as much larger Stokes shift against those for QDs.
The vast class of CT colorants is presented by transition metal complexes, the metal-to-ligand charge transfer (MLCT) of which is facilitated by non-planar molecular geometry [30,31] with high coordination numbers, and driven by a pronounced back donation phenomenon [32]. Remarkably, the molecular design of highly effective MLCT dyes derived from Mo [27], Os [4,6], Re [4,6,33], and mainly Ru [4,6,27,34,35] has been named an absolute priority, which can be explained by their use as photosensitizers (otherwise “redox mediators” [36]) in photovoltaic (PV) devices of Grätzel cell’s type [37,38]. As a rule, these MLCT chromophores demonstrate long-time stability under applied redox potential, as well as high photoelectric conversion efficiency (PCE) at ≈10% [6,39] in the case of third generation (3G) devices. Nevertheless, sound minuses of conventional Ru dyes are toxicity, high cost, and low abundance [40].
In this regard, CT chromophores of D-MII-A type (often this is dye with ligand-to-ligand charge transfer (LL’CT) and less coordination numbers) can provide a competitive alternative towards MLCT colorants due to noteworthy aspects: (1) pronounced charge separation on a molecular level with a planar mutual arrangement of interacting HOMO (donor organic part) and LUMO (acceptor organic part) [32]; (2) higher intense absorptivity; (3) an ability to fine-tune CT energy in the course of molecular design, varying electronic properties of ligands by functionalization; (4) lower energy of effective intramolecular CT in some cases, leading to an absorption/emission even in the NIR region [29] of an undeniable practical importance. It should be noted that a diminution in CT energy up to NIR absorption/emission is one of the priority tasks in the colorant chemistry. In particular, NIR light is ~45% of the solar energy [39] which makes the building of PV devices on the base of NIR photosensitizers reasonable (i.e., have the maximums of CT band in NIR region). In turn, the use of CT chromophores as luminescent labels in bioimaging screening requires NIR luminescence, since NIR radiation can penetrate into body tissues much more efficiently than UV/visible emission [18]. Additionally, an introduction of electrochromic NIR chromophores for electro-optic switching should be a promising project within fiber-optic technologies, especially those working at “telecommunication wavelengths” (from first to third wavelengths, 850, 1300, and 1550 nm, respectively) region [11].
Since a planar molecular geometry (most frequent in CT chromophores of D-MII-A type) is a favorable and necessary condition for realization of effective intramolecular LL’CT [32,41], at present the overwhelming majority of LL’CT chromophores comprises the derivatives of 10 group metals (Ni, Pd, Pt), which prefer a square-planar coordination environment [42]. Against 10 group metal CT dyes, very scant attention is paid to compounds of other transition elements, and a range of such chromophores is vastly limited in number and nature: thus, sporadic sources reported on 3d-row metal (Co [43] and Cu [44,45,46,47]) CT complexes. There are serious reasons for the observed disbalance. Many 3d-row elements tend to be surrounded by non-planar coordination environment, and the compounds often form with high coordination numbers (more than 4). In particular, tetrahedral polyhedron is an exemplary situation for four-coordinated Zn complexes [48,49,50,51], whereas in the case of iminoquinonato Cu derivatives, well-known equilibrium “square-planar—tetrahedron” [30,31] is finely susceptible towards ligands’ nature, i.e., bulkiness and carbon backbone rigidity, electronic properties and “hardness/softness” [52] of heteroatoms from potentially coordinating substituents. Regarding metals with more electron deficient d-orbitals, the tendency of complexes to increase their coordination number exists, and a steric factor plays a lesser role. For instance, quinonato/iminoquinonato Mn and Fe complexes are generally six-coordinated octahedral, while five-coordinated square-pyramidal/trigonal bipyramidal Mn and Fe derivatives are less common and contain more bulky iminoquinone type ligands [30,31], and the cases of four-coordinated Fe species are very scarce [53]. More complicated and an ambiguous situation is implemented for quinonato/iminoquinonato Co complexes: in contradistinction to exemplary six-coordinated quinonato derivatives, iminoquinonato species demonstrate a multeity among coordination numbers (4–6), that depends on slightest fluctuations in ligands’ bulkiness [30]. With a perspective to design CT chromophores of Co, great interest should be paid to coordinatively unsaturated Co complexes. An ability to generate stable square-planar derivatives was confirmed clearly by the example of iminoquinonato species [54,55,56,57,58], being driven by a steric factor. Moreover, the bright instance of tetrahedral Co complex and its isostructural Ni analog were reported recently [59]. Remarkably, a degree of a steric hindrance for coordinating atoms from ligands is an important factor within synthesis of D-MII-A chromophores, which determines their monomeric/oligomeric structure in the case of metals with the ability to form derivatives with higher coordination numbers than 4 (for instance, see ref. [47], as well as Section 2.2, and references in this chapter towards monomeric and oligomeric quinonato Cu/Co complexes).
As another serious omission in dye chemistry of CT chromophores of D-MII-A type, many available works take the main aim to perform a synthesis and routine analysis of general physic-chemical properties. On the other side, there is a scattered and non-systematized information concerning important CT characteristics, such as: (1) lifetime of excited state (all data for PtII chromophores) [10,16,17,60,61,62,63,64,65,66]; (2) CT energy in terms of its dependence on the nature of metal center [67,68,69,70,71]/electronic properties of ligands [11,17,28,32,41,42,60,61,62,63,64,68,70,71,72,73,74,75,76]/solvent polarity [32,42,60,61,62,63,64,67,76,77]. However, knowledge about these key parameters helps turn colorant into “functional dye”. For example, it should be highlighted that one comprehensive study of LL’CT energy [67] determined comparatively an influence of metal center’s nature: 3,5-di-tert-butyl-catecholato-MII-2,2′-bipyridine (where M is Ni, Pd, Pt) complexes display a slight increase of an extinction coefficient and narrowing of HOMO-LUMO gap within [Ni] < [Pd] < [Pt]. Remarkably, observed changes are expectedly minor and conditioned by an identical electronic structure of metal d-levels.
On the way to “functional CT dyes”, a number of recent reports can be identified as cutting-edge research designed to bridge the gap between “synthesis–structure–property” research pattern in attempts to apply or at least to reach/improve practically important characteristics. First of all, the extremely short lifetime of excited states of 3d-row transition metal complexes [40] can cross them off the set of solar cells’ photosensitizers and LED emitters. Being suitable model objects, PtII LL’CT dyes were chosen by many authors for lifetime’s measurements, including such methods, as time-resolved infrared (TRIR) and transient absorption (TA) spectroscopy [10,16,60,62,63,65,66]. Promising chapter within lifetime’s study, as well as within investigations of photoinduced electron spin polarization (ESP), belongs to magnetoactive PdII and PtII LL’CT chromophores, the catechol part of which is equipped by nitronylnitroxyl radical [78,79]. At the same time, an ability to generate long-lived excited species exists, for instance, in the case of dyes based on nodal metals with a full d-level. Such an approach was shown by example of bis(arylimino)acenaphthene CuI complexes [80]. The search for an alternative to the proven cisplatin can considered as a long-standing medical challenge, and is a typical for heteroligand Pt/Pd chromophores, mainly carboxylate [81,82,83,84,85] or quinonato [86,87,88,89,90] derivatives of Pt and Pd, which are augmented by ammine or condensed α-diimine moieties. On the other side, the heeds of modern fields such as solar energy have promoted the testing of high-effective LL’CT colorants in recent years, and the application of Ni [7,39,40,91], Pd [7,91], Pt [62,92,93,94], and Cu [36,80] complexes in dye-sensitized solar cells (DSSCs) is known.
Summarizing, the further molecular design of CT colorants should be continued as a prospective, targeted action, directed to a synergy between fundamental research and a response to practical challenges. Involving the derivatives of 3d-row “late” transition elements (first of all, Cu and Co) other than 10 group metals promises to be interesting: considerable results can be achieved, taking into account the diverse structure of d-levels, and lability in the formation of a coordination environment. Following above trend, we present CT chromophores 13 of 3d-row “late” transition metals (Cu, Ni, Co) based on 4,4′-di-tert-butyl-2,2′-bipyridyl and 3,6-di-tert-butyl-o-benzoquinone in terms of used synthetic approaches, molecular structures’ peculiarities, as well as CT characteristics, CuII(3,6-Cat)(bipytBu), NiII(3,6-Cat)(bipytBu) and [CoII(3,6-Cat)(bipytBu)]2 complexes, respectively. Interrelations between CT energy, structural features, and the nature of nodal metal center are explored in detail.

2. Results and Discussion

2.1. Syntheses of Complexes 13

Three different synthetic approaches (Figure 1) were chosen to obtain a series of new CT chromophores, monomeric CuII(3,6-Cat)(bipytBu), NiII(3,6-Cat)(bipytBu) and dimeric [CoII(3,6-Cat)(bipytBu)]2 complexes (1, 2 and 3 respectively).
An effective synthetic technique was proposed earlier [44,45,46,47] to obtain the “α-diimine-CuII-catecholate” type compounds from a set of 1,2-diols (cathechol, 3,5-di-tert-butyl-catechol, tetrachlorocatechol, 3-n-nonylcatechol, 2,3-dihydroxynaphthalene), and α-diimines (ethylenediamine, 2,2′-bipyridyl, 1,10-phenanthroline, N,N,N′,N′-tetramethylethylenediamine), which differ in electronic properties (donor/acceptor ability) and in a bulkiness. A significant discrepancy of this method from one conventional for heteroleptic NiII analogs (see below) lies in an absence of a redox stage: derivative of the divalent metal ion (copper(II) sulfate) was chosen as a metal-containing compound, and dianion redox forms of donor ligands took part in the interaction. It should be noted that the similar approach is a common pattern for preparing of PdII and PtII LL’CT chromophores. In this case, additional reagents can be needed for an activation of the 1,2-diols. For instance, some authors [60,62,67,68,72,95] used bases, such as triethylamine, potassium tert-butoxide or sodium hydroxide. On the contrary, the synthesis of CuII(3,6-Cat)(bipytBu) (1) was performed through a convenient “one-pot” redox process [96] with participation of neutral 3,6-di-tert-butyl-o-benzoquinone (3,6-DTBQ), 4,4′-di-tert-butyl-2,2′-bipyridyl, and an excess of metallic copper, yielding 81% of targeted product (Figure 1).
Due to the greatest prevalence of 10 group metals’ CT complexes (see Section 1), a set of well-developed synthetic techniques for such metal derivatives is known, depending on a redox state of o-quinone type ligand and an initial metal-containing compound. In particular, synthesis of the Ni chromophores proceeds in some cases with a participation of o-quinones, as well as Ni(0) derivatives coordinated by neutral ligands (tetracarbonyl nickel(0) [43] or bis(cyclooctadiene)nickel(0) [41,42], for example). The above method does not require any additional activating agents and involves a redox stage—a two-electron reduction of the redox-active o-quinone to a catechol form, that is accompanied by an oxidation of a metal center to a divalent state. Thus, following the authors [41,42], we have recently postulated [32] bis(cyclooctadiene)nickel(0) (Ni(cod)2) as a convenient “parent” to prepare the LL’CT chromophores of “α-diimine-NiII-catecholate” type with high yields, applying neutral 3,6-di-tert-butyl-o-benzoquinone (3,6-DTBQ) and related 7,10-di-tert-butyl-2,5-dioxabicyclo(4.4.0)deca-1(10),6-diene-8,9-dione (3,6-DTBQgly). In an absolute concordance with an analogous synthetic route described recently [32], the complex NiII(3,6-Cat)(bipytBu) (2) was prepared, yielding 84% of a crystalline individual product. As can be seen from Figure 1, the first-step exchange interaction between Ni(cod)2 and 4,4′-di-tert-butyl-2,2′-bipyridyl in ratio 1:1 proceeded in a glovebox under argon atmosphere, whereas the second redox stage leading to the final complex was carried out under anaerobic conditions.
Another activating method for the donor organic platform was applied by authors [43] to synthesize CoII derivatives on the base of dry CoCl2, 3,6-di-tert-butyl-o-benzoquinone (3,6-DTBQ), 2,2′-bipyridyl or bulky substituted diazabutadienes: thallium catecholate was incorporated into the reaction mixture, being prepared in situ by a direct interaction between 3,6-di-tert-butyl-o-benzoquinone (3,6-DTBQ) and thallium amalgam (approximately 10-fold excess). Modifying slightly this method in the presented work, the dimeric [CoII(3,6-Cat)(bipytBu)]2 complex was produced under anhydrous conditions with yield 78% in the course of two-step synthetic procedure, using sodium catecholate (CatNa2) (Figure 1).
As an interesting fact, the ability to manage the CT chromophoreʼs composition through a fine-tuning of the synthetic procedure (hydrous/anhydrous conditions), which allows generating monomer or dimeric species, was illustrated by the example of (3,5-di-tert-butyl-catecholato)(2,2′-bipyridine)copper(II) derivatives [47]. Moreover, an oligomerization can be considered as a frequent enough phenomenon for homoleptic transition metal derivatives of 3,5-di-tert-butyl-o-benzoquinone type platform: a dimerization of bis-o-benzosemiquinonato CuII moieties [97], as well as a tetramerization of o-benzosemiquinonato CoII units [98] are fulfilled due to a dative bonding between oxygen atoms and metal ions. In contrast, a formation of monomeric molecular structures is a typical situation for coordination compounds derived from 3,6-di-tert-butyl-o-benzoquinone. A higher degree of the oxygen atom’s steric hindrance prevents the existence of oligomeric species. In this regard, an obtaining of dimeric [CoII(3,6-Cat)(bipytBu)]2 complex (3), bearing two different bulky organic parts, is a noteworthy event. Summarizing, our strategy to prevent the oligomerization of D-MII-A parts by shielding of coordinating atoms from ligands due to an incorporation of substantially bulky tert-butyl groups turned out to be effective only in the CuII derivative 1.
The synthesis of complexes 13 requires conditions excluding air moisture and oxygen. However, the resulting compounds are quite air stable and can be stored in open vessels in a crystalline form.

2.2. Molecular Structures of Complexes 13

It is acknowledged that the “α-diimine-MII-catecholate” type complexes have a square-planar structure. Hence, the planar mutual arrangement of overlapping HOMO and LUMO, facilitates an effective intramolecular ligand-to-ligand charge transfer [30,32,41]. As a consequence, to date, there is a numerous set of LL’CT heteroleptic chromophores, each of which is distinguished by a planar molecular structure [99].
In turn, complexes 1 and 2 were no exception. The coordination polyhedra possess a minimal distortion (Figure 2), in accordance with values τ4(1, 2) = 0.05 [100,101]. Moreover, an ideal planarity (excluding tert-butyl groups) is featured for molecule of 1, since the molecule lies on a special position. As a result, zero angles are formed by ligand planes, as well as by O(1)-Cu(1)-O(2) and N(1)-Cu(1)-N(2) planes of a coordination core. On the contrary, the peculiarities of molecular packing (detailed description is given below) cause a deviation from planarity in 2: O(1A)-Ni(1A)-O(2A) and N(1A)-Ni(1A)-N(2A), O(1B)-Ni(1B)-O(2B) and N(1B)-Ni(1B)-N(2B) planes intersect at angles 2.81° and 3.53°, while the angles between catecholate dianion and bipyridine are found to be 11.96° and 13.98° (for molecules A and B, respectively).
Another molecular structure is inherent for the complex 3: “α-diimine-CoII-catecholate” species are combined into dimer [CoII(3,6-Cat)(bipytBu)]2∙toluene (Figure 2) through a donation from the oxygen atoms of one D-MII-A part to the metal center of an adjacent moiety. As a result, each nodal CoII ion is surrounded by an almost ideal (τ5 = 0.087 [102]) square-piramidal coordination environment with an apical oxygen atom from neighboring D-MII-A part of dimer. Corresponding distance Co(1)⋯O(2*) is equal to 2.0222(18) Å, which is less substantially for those towards cognate heteroleptic [CuII(3,5-Cat)(bipy)]2 [47] and homoleptic [CuII(3,5-SQ)2]2 [97] dimers, however comparable for bridging lengths in tetrameric [CoII(3,5-SQ)2]4 [98] derivative (Table 1) of octahedrally coordinated CoII ion (where 3,5-SQ is 3,5-di-tert-butyl-o-semiquinone, bipy is 2,2′-bipyridine). In turn, intermetallic contacts Co(1)⋯Co(1*) = 2.8847(7) Å are shortened sizably, as follows from Table 1. In general, a formation of tetrameric structures is preferable for CoII compounds derived from o-benzoquinone type ligands, as 3,5-di-tert-butyl-o-benzoquinone [98] and 9,10-phenanthrenequinone [103], but a turn to 3,6-di-tert-butyl-o-benzoquinone with a higher degree of oxygen atoms’ shielding leads usually to monomeric species [104]. However, the complex 3 is quite remarkable: the presence of two bulky tert-butyl groups in both coordinating ligands does not interfere a dimerization. Each chromophoric D-MII-A part in 3 is non-planar. Interligand angle values are sufficiently large (56.97°), almost twice those (32.44°) for [CuII(3,5-SQ)(bipy)]2 dimer mentioned above.
Metal complexes 13 under investigation possess the same unambiguous electronic structures of “α-diimine-MII-catecholate” species (Table 2). In particular, the classical bond lengths for dianionic O-donor ligand platforms (average C–C = 1.39–1.41 Å, C–O = 1.32–1.39 Å [31,105]) are found, with typical relative shortening for C(4)–C(5) distances of catecholate rings, due to a pronounced bonding nature for HOMO between mentioned carbon atoms [105]. In turn, a donor-acceptor coordination mode is implemented for neutral bipyridine organic moieties. Corresponding M-N and C-N bonds (Table 2) are in a good accordance with relative ones in known CT chromophores [32,68,72,87,89].
The “metal-heteroatom” lengths for NiII complex 2 are almost identical to those presented for NiII(3,6-Cat)(bipy) analog [32]. In turn, the same distances in CuII(3,5-Cat)(bipy) [47] slightly exceed ones for CuII derivative 1 (Table 3). A non-equidistance is observed for Co-O lengths owing to dimerization with a participation of oxygen atoms. As expected, the Co-O and Co-N bonds in 3 are elongated substantially compared to CoII(3,6-Cat)(DAD) (Table 3). This fact is in a good agreement with a growth in the ionic radius of cobalt as the coordination number increases [106], as well as with a higher steric hindrance of metal center in complex 3.
The THF molecule in crystal 1 is disordered over two positions and is oriented by the oxygen atom to the hydrogen atoms of the bipyridine ligand. As a result, weak intermolecular hydrogen interaction O(THF)⋯H(16A) (2.653 Å) or O(THF)⋯H(19A) (2.600 Å) is realized.
An ideal lamellar crystal structure without π-stacking is inherent for complex 1 (Figure 3). The distances between molecular layers are equal to 4.559 Å. A similar situation with weakly interacting molecular layers was observed for a series of PdII compounds on the base of perfluoroalkyl-bearing catechol and various substituted bipyridines [68]. In contrast, the stabilization of a crystal structure is characterized for less bulky NiII(3,6-Cat)(bipy) reported recently [32] through intermolecular π-π contacts of 2,2′-bipyridines. Indeed, an existence of π-stacking can be assumed as a distinctive feature for many lamellar CT chromophores, containing conjugated organic platforms with a rigid carbon backbone [32,68,72,107,108]. Nevertheless, the presence of bulky tert-butyl groups in donor/acceptor ligands, as well as solvate THF molecules (one molecule per molecule of complex) in 1 makes any π-π interactions impossible. Molecules of complex 1 in odd layers are strictly under each other, and a similar pattern is observed for molecules of even layers. So, the complex 1 demonstrates a mutual molecular arrangement, minimizing a repulsion of tert-butyl moieties (Figure S1).
Despite the lamellar laying realized in 1, in general, the herringbone-like packing folded from T-motifs become the most preferable “building blocks” for a majority of CT chromophores of D-MII-A type with hindered organic moieties [32,67,73,87,89,90,95,109]. As a rule, molecules within such T-motif are stacked through the π-π interplays involving ligand orbitals, while there are no substantial π-π inter-stack contacts because of a steric shielding effect from bulky substituents in ligands (for instance, tert-butyls [95], iso-propylphenyls [73], perfluoroalkyls [68], etc). Complex 2 demonstrates the same pattern. The crystal structure contains isolated pairs of crystallographically independent molecules A and B (Figure 3). These T-motifs have a mutual arrangement close to perpendicularity, since molecular planes form angles at 72.31°/77.82°/83.37° (between molecules from different pairs—(A and A)/(A and B)/(B and B), respectively. Thus, the shortest contacts in dimeric pair are found: (1) between centroids of bipyridine ring Cdbipy and chelate cycle CdCh, which is adjacent to catecholate dianion (CdCh(A)Cdbipy(B) 3.735 = Å, CdCh(B)Cdbipy(A) 3.783 = Å); (2) between nickel atoms (Ni(1A)⋯Ni(1B) 3.635 Å), Figure S2).
The molecular packing of compound 3, continuing the tendencies for derivatives 1 and 2, is organized according to the principle of least repulsion between complex’ molecules and solvate toluene molecules (Figure 3). The plane of the C6-ring of the catecholate and the NC5-ring of the bipyridine ligand of the neighboring molecule are almost perpendicular to each other. The corresponding dihedral angle is 74.68°. Thus, the possibility of realizing the π-π interaction is completely excluded. However, the shortest contact CdCat⋯C(23) = 3.695 Å (where CdCat is centroid of C6-ring, Figure S3) may indicate the presence of intermolecular C-H⋯π interaction in the crystal of 3.
It should be noted that an incorporation of sterically hindered functions into the ligands can be assumed as a rational strategy to increase the solubility of the complex. A tendency to π-π intermolecular stacking is weakening. For instance, a problem of solubility has been raised for CuII CT chromophores [44]. As a result, all studied coordination compounds 13 possess good solubility in a wide range of solvents with various polarity.

2.3. EPR Investigation for Complex 1

It is known that square-planar NiII complexes are characterized by a low-spin state of the metal ion. As a result, compound 2 is diamagnetic, as evidenced by a well-resolved NMR spectrum. A related CuII derivative 1 has the d9 configuration of a metal ion and is paramagnetic. X-Band EPR spectra recorded for solid sample of 1 (Figure 4a) demonstrate anisotropic signals with an axial symmetry (g = 2.18, g = 2.08, A(63,65Cu) = 220 G), which is characteristic for square-planar CuII compounds [110]. The finely resolved EPR spectrum for complex 1 in CH2Cl2 solution at room temperature also exhibits a four-line pattern (Figure 4b) as the hyperfine splitting (hfs) of unpaired electron with copper magnetic isotopes (I = 3/2). High-field components of spectrum show additional quintets 1:2:3:2:1 which are the result of hfs on two equivalent nitrogen nuclei 14N (I = 1) of bipyridine ligand. Spectrum parameters are gi = 2.100, ai (63,65Cu) = 92 G, ai (2 14N) = 10 G. No half-field signal is observed for either solid or vitrified (T = 150 K) CH2Cl2 matrix. This indicates the absence of any dimerization of 1 in crystals or in solution.
Dimeric compound 3 has two CoII paramagnetic ions. However, the cobalt derivative does not have any X-band EPR signal in solid or in solution at room temperature and being frozen down to 100 K. This illustrates the preservation of the dimeric structure in the solution of 3. The formation of a low-spin CoII (SCo = 1/2) complex would be observed with dissociation. Such CoII derivatives show the corresponding signal in the EPR spectrum of frozen solutions [43].

2.4. Magnetochemical Study for Complex 3

An almost ideal square-pyramidal coordination environment of metal centers in 3 determined their high-spin state (SCo = 3/2), being a usual situation for transition elements in d7 electronic configuration [30]. According to the temperature-variable magnetic susceptibility measurements, the high-temperature (T = 300 K) effective magnetic moment (μeff) value for complex 3 is equal to 6.39 μB, which transcends insignificantly the theoretical spin only value of 6.30 μB for two high-spin CoII ions (SCo = 3/2, quadruplet spin state) with g = 2.30. The magnitude of μeff changes slightly in the temperature range 300–100 K, whereas below 100 K the μeff(T) curve drops sharply down to 0.83 μB at 2 K, that clearly indicates a presence of antiferromagnetic exchange interactions between spins of two CoII ions (Figure 5). Analysis of the μeff(T) dependence using dimer model (Spin Hamiltonian H = –2J·S1S2) allows estimating exchange interaction parameter J. The best fit values of g-factor and J are 2.39 ± 0.01 and −4.76 ± 0.03 cm−1, respectively.

2.5. Electrochemical Study for Complexes 13

The electrochemical behavior of CuII and NiII monomeric derivatives 1 and 2 can be considered as similar, consisting in ligand-centered redox-stages (Figure 6): reversible one-electron oxidative wave corresponds to a transition “catecholate dianion → semiquinone radical anion”, and quasi-reversible one-electron reduction of coordinated bipyridine ligand is observed.
It should be noted that the values of electrochemical potentials for complex 2 and NiII analog NiII(3,6-Cat)(bipy) with unsubstituted 2,2′-bipyridine [32] are very close. However, the HOMO → LUMO gap is wider expectedly for 2 due to the presence of electron-donor tert-butyl fragments in acceptor diimine part (Table 4).
On the contrary, dimeric CoII complex 3 displays another character of voltammogram (Figure 6). Quasi-reversible one-electron redox process, which is detected with significant displacement towards reductive steps for 1 and 2, can be attributed to CoII → CoI reduction. Noticeably, the same metal-centered redox transformations were described in detail for CoII derivatives based on the polypyridyl multidentate ligand platforms [111,112]. Two consecutive one-electron reversible peaks are observed on the oxidative curve of the voltammogram. These processes can be attributed with equal probability to both the oxidation of catecholate ligands to o-semiquinone ones, and to the transition of CoII → CoIII [112].

2.6. UV-vis-NIR Spectroscopy for Complexes 13

As it was highlighted above, a comparison of chromophores from a standpoint of CT energy is a quite rare final goal for corresponding investigations. H.-C. Chang and co-workers have shown [67] that changes in absorptivity among analogous CT dyes on the base of 10 group metals (Ni, Pd, Pt) cannot be considered as “substantial”: the wavelength shift of LL’CT band did not transcend at ~20 nm (in the same solvent) between derivatives of metals from adjacent periods. On the other side, to the best of our knowledge, there are no targeted investigations establishing how the nature (d-level structure, hence a coordination sphere volume and preferable coordination environment) of metals from other groups affects CT energy. However, this task may represent certain perspectives within CT chromophores’ development and design. Thus, in this paper, we will try to work on this issue.
The key feature of absorption spectra for complexes 13 is the presence of high-intense broadened CT bands extending in the visible/NIR regions (Figure 7), and a pronounced negative solvatochromic effect (Table 5).
Regarding the NiII complex 2, the energetic parameters of LL’CT (according to CV data, see above) are identical to those for the related NiII chromophore, containing 3,5-di-tert-butyl-catecholate ligand instead of 3,6-substituted counterpart (Table 5). An incorporation of tert-butyl functions to 2,2′-bipyridine platforms in 2 expectedly initiates a slight hypsochromic shift (within 20–35 nm in dependence on solvent) of the LL’CT bands (Table 5) in comparison with recently published (3,6-di-tert-butyl-catecholato)(2,2′-bipyridine)nickel(II) [32]). As can be seen from the current results (see Table 5), an equivalent effect towards LL’CT (according to CV data, see above) energy can be caused by a replacement of the metal center from NiII to CuII ion with the unchanged ligands and a similar molecular structure (monomer, square-planar coordination core): the LL’CT bands’ shifts to shortwave region are equal to 26–40 nm in the case of CuII analog 1 (Table 5). Another important fact that can be seen when comparing the photophysical characteristics of isostructural complexes 1 and 2 is that replacing the NiII cation with a CuII almost halves the extinction coefficient.
It is necessary to note that compounds 1, 2, and abovementioned metal complexes reported earlier [32,67] can be classified as so-called “visible CT dyes (i.e., have the maximums of CT band in visible region)” based on λmax values (without considering the band width). The generation of effective NIR chromophores is extremely relevant and dictated by different challenges, thereby one of the prior strategies within molecular design of CT dyes should be considered a targeted decrease of CT energy until to NIR absorptivity. One way is a modification of the donor catecholate moiety against acceptor diimine part. This is a quite efficient approach: (1) a slight turn in electronic properties of 3,6-di-tert-butyl-catecholate ligand in the course of an annelation with electron-donor glycol fragment shifted the LL’CT band at 95 nm, producing NIR dye (λmax = 810 nm, in toluene solution) of NiII (Table 5) [32]; (2) a cardinal design (A. F. Heyduk with co-workers) [42]) of catecholates’ composition and structure (3,5-di-tert-butyl-o-benzoquinone, tetrachloro-1,2-quinone and 9,10-phenanthrenequinone were used) among a wide set of “α-diimine-NiII-catecholate” species resulted a record-breaking red LL’CT band’s displacement at about ~430–550 nm up to deep NIR region (the largest value λmax = 1370 nm, in THF solution).
Here, we report another possible strategy of molecular design on the way to NIR CT chromophores: a turn from 10 group metal in NiII complex 2 (convenient standard in LL’CT chemistry due to a preference of a square-planar coordination polyhedron) to the 3d-row transition element of different nature with more labile coordination sphere. Dimeric CoII derivative 3 demonstrate UV-vis-NIR spectra with high-intense (within studied complexes 13) (M + L)L’CT (according to CV data, see above) absorption bands which are shifted from the visible to the NIR region (Table 5). It should be remarked especially that the dimeric structure of chromophore 3 is maintained regardless of the polarity and coordination ability of used solvents (CH3CN, DMF, CH2Cl2, THF) that is proved by a synergy of UV-vis-NIR and EPR-spectroscopy data: (1) the general character of UV-vis-NIR spectra was independent towards concentration effect, as well as solvent nature—high-intense (M + L)L’CT bands are presented (Figure 7); (2) no X-band EPR signal for complex 3 was observed in CH2Cl2 solution. Obviously, dimer destruction should generate “α-diimine-CoII-catecholate” species, and an implementation of a square-planar or tetrahedral molecular geometry is equiprobable for similar four-coordinated cobalt derivatives, as highlighted in Section 1. The tetrahedral mutual arrangement between HOMO and LUMO eliminates CT and, therefore, an observation of corresponding CT band in UV-vis-NIR spectra is impossible. In turn, square-planar coordination environment provides a low spin state of divalent cobalt ion (SCo = 1/2) [30,31], conditioning the paramagnetism of such “α-diimine-CoII-catecholate” monomer. Thus, the existence of a dimeric structure for 3 in solution is beyond doubt.
The pronounced negative solvatochromism is established for chromophores 13 as expressed by a fine linear dependence ETN(λ) (Figure 8) between CT bands and the normalized empirical parameter ETN by Dimroth and Reichardt [113]. The ETN parameter is obtained empirically on the base of the solvent-dependent CT band’s shift for the standard N-phenoxypyridinium betaine dye, and, hence, describes the non-specific solvent polarity. CT bands’ displacement for chromophores 13 is explained by fine-sensitive energy of non-polar excited states towards the solvent polarity. So, a turn from CH3CN to THF or toluene facilitates CT, causing substantial red shifts (Table 5) due to CT energy’s decrease. The found parameters of solvatochromic displacement are comparable with those for known “α-diimine-NiII-catecholate” analogs (Table 5), being one of the largest reported values [32,42,67].

2.7. DFT Calculations

Quantum chemical calculations by the Density functional theory (DFT) method at the B3LYP/6-311++g(2d,2p) level of theory were performed to study the electronic structure of complexes under investigation. The experimental (X-ray) geometries were used as the starting points and optimized. Compound 1 was calculated in an open-shell approximation with S = 1/2 ground state. EPR spectroscopy data confirms doublet spin state for CuII derivative 1. Complexes 2 and 3 were calculated in a closed-shell approximation with S = 0 ground state. The diamagnetic nature of NiII complex 2 was evidenced by well-resolved NMR spectrum with no line broadening. The singlet ground state for dimeric CoII derivative 3 is clear from the solid state magnetochemical measurements. The selected frontier orbitals are presented in Figure 9 and Figure 10.
The LUMO orbital (−2.51 eV) (Figure 9, (4)) is located mainly on the acceptor diimine part of the NiII complex 2. This well-defined HOMO (−4.38 eV) (Figure 9, (2)) occupies catecholate ligand with small nickel dxy and bipyridine system contribution. Orbitals energy and the nature of their distribution is quite obvious for such type of complexes [32]. The energy of the HOMO orbital increases significantly (0.41 eV) when tert-butyl substituents are introduced into the 2,2′-bipyridine [32]. These changes lead to an increase in the HOMO-LUMO energy gap (1.87 eV) and a corresponding hypsochromic shift observed in the electronic spectrum of compound 2 compared to the previously published derivative containing unsubstituted bipyridine. The energy (1.82 eV) of longwave absorption band (680 nm in toluene) is in a good agreement with calculated HOMO-LUMO energy gap. The nearest orbitals with a prevailing contribution of metal AO are HOMO-2 (dz2, −6.19 eV) and LUMO+3 (dx2-y2, −1.14 eV). Thus, complex 2 can be considered as a LL’CT donor-acceptor chromophore which is controlled by the ligand framework.
The situation with the formation of frontier orbitals for the related CuII complex 1 turns out to be very similar. HOMO (−4.28 eV) and α-LUMO (−2.58 eV) are located on the donor catecholate and acceptor diimine ligands respectively (Figure 9, (1) and (3)). The appearance of an additional electron at the d-level of the metal leads to a general decrease in the energy of metal atomic orbitals. Thus, the highest for squire-planar configuration semioccupied dx2-y2 orbital is α-HOMO-2 (−5.93 eV) for complex 1. This orbital is responsible for the doublet spin state of the compound (Figure 9, (5)). A decrease in the energy of the d-orbitals in the complex 1 (compared to the NiII derivative 2) leads to a decrease in the contribution of the dxy orbitals in the HOMO and LUMO. It reduces the efficiency of charge transfer between the donor and acceptor fragments of the molecule and the molar extinction coefficient for the CuII compound 1 is about two times lower than the NiII analog 2.
The composition and energy of the frontier orbitals undergo a significant change in the CoII complex 3 (Figure 10). HOMO (−3.82 eV) and LUMO (−2.25 eV) have a prevailing metal character which is confirmed by the data of electrochemical studies. A noticeable contribution of the diimine fragment is observed for LUMO. The nearby HOMO-1 (−4.19 eV) and HOMO-2 (−4.44 eV), as well as LUMO+1 (−2.00 eV) and LUMO+2 (1.97 eV), are occupied by catecholate and bipyridine ligands respectively. These orbitals have an appreciable contribution of cobalt d-orbitals. Thus, the cobalt ion in the resulting donor-acceptor complex provides effective charge transfer due to the active involvement of metallic d-orbitals into the frontier orbitals formed by redox-active organic fragments. It should be noted that the decrease in the molar extinction coefficient in low-polar solvents for CoII derivative 3 is not observed, despite the formation of a five-coordinate environment of the metal atom. This configuration of the complex can provide charge transfer between the donor and the acceptor [114]. However, it is significantly inferior to structures with a square-planar geometry.

3. Materials and Methods

The comprehensive general information towards used materials and methods is collected in Supporting Information.

Syntheses and Characterization of Complexes 13

To prepare complexes 13, combined conditions were implemented: the second stages of the two-step synthetic procedure in the case of compounds 2 and 3, as well as the synthesis of 1, were performed under Schlenk line, while an argon atmosphere glovebox was used to carry out the second stages for 2 and 3.
Complex 1·THF, [CuII(3,6-Cat)(bipytBu)]∙THF. Equimolar amounts of 3,6-di-tert-butyl-o-benzoquinone (3,6-DTBQ) (0.5 g, 2.27 mmol) and 4,4′-di-tert-butyl-2,2′-bipyridyl (0.61 g, 2.27 mmol) reacted with an excess of sliced copper foil in dry THF (20 mL) in evacuated ampoule. The resulting solution, which turned from bright-green to ink-blue color, was decanted from copper and partially evaporated, after that dark fine-crystalline product formed. Precipitate was collected by filtration and washed with THF under reduced pressure. Single crystals of [CuII(3,6-Cat)(bipytBu)]∙THF with C36H52CuN2O3 composition, which are suitable for X-ray diffraction experiment, were obtained from THF through a slow evaporation of solution under residual pressure. The total yield of product with C36H52CuN2O3 formula is 1.15 g (81%) based on the starting ligands.
The polycrystalline samples of 1 for an elemental analysis, IR-, EPR-, UV-vis-NIR spectroscopy and cyclic voltammetry were prepared from THF.
IR (Nujol, KBr) cm−1: 488 (m), 538 (w), 557 (m), 604 (s), 650 (s), 698 (s), 742 (m), 752 (w), 771 (m), 809 (m), 851 (s), 899 (s), 923 (m), 939 (s), 980 (s), 1023 (m), 1034 (m), 1053 (s), 1117 (m), 1145 (m), 1172 (w), 1202 (m), 1246 (s), 1258 (s), 1279 (m), 1304 (w), 1347 (m), 1367 (m), 1409 (s), 1552 (s), 1582 (w), 1620 (s), 1706 (w), 1787 (w), 1832 (w), 1965 (w), 3062 (m), 3114 (w).
Anal. Calc. for the complex [CuII(3,6-Cat)(bipytBu)]∙THF with C36H52CuN2O3 composition (%): C, 69.25; H, 8.39; N, 4.49. Found: C, 69.56; H, 8.86; N, 4.75.
Complex 2, NiII(3,6-Cat)(bipytBu). A colorless solution of 4,4′-di-tert-butyl-2,2′-bipyridyl (0.5 g, 1.86 mmol) in dry toluene (20 mL) was poured to an equimolar yellow solution of Ni(cod)2 (0.51 g, 1.86 mmol) in the same solvent (20 mL) in glovebox. A resulting reaction mixture was stirred in an evacuated ampoule for an hour under an ambient temperature when a deep purple solution was obtained. An addition of an equimolar green solution of 3,6-DTBQ (0.41 g, 1.86 mmol) in dry toluene (10 mL) changed a reaction mixture’s color to ink blue. A slow removal of the reaction mixture under reduced pressure facilitated a growth of ink rectangular-shaped single crystals of NiII(3,6-Cat)(bipytBu) with C32H44N2NiO2 composition. The polycrystalline precipitate was filtered off and washed with dry hexane under residual pressure. The total yield for product NiII(3,6-Cat)(bipytBu) with C32H44N2NiO2 formula is 0.84 g (84%) based on the starting ligands.
The polycrystalline samples of NiII(3,6-Cat)(bipytBu) for an elemental analysis, IR-, 1H NMR, UV-vis-NIR spectroscopy, and cyclic voltammetry were obtained from toluene, and dried additionally in anaerobic conditions.
Anal. Calc. for the solvent-free complex NiII(3,6-Cat)(bipytBu) with C32H44N2NiO2 composition (%): C, 70.21; H, 8.10; N, 5.12. Found: C, 69.83; H, 8.09; N, 5.00.
IR (Nujol, KBr) cm−1: 462 (m), 485 (w), 520 (m), 540 (m), 569 (w), 599 (s), 612 (m), 648 (s), 670 (w), 710 (s), 727 (s), 743 (m), 776 (m), 810 (m), 846 (s), 887 (m), 901 (m), 927 (w), 941 (m), 982 (s), 1024 (m), 1035 (m), 1076 (w), 1115 (w), 1141 (m), 1173 (w), 1201 (m), 1244 (s), 1263 (m), 1281 (m), 1295 (w), 1310 (m), 1348 (m), 1366 (m), 1412 (s), 1478 (m), 1546 (m), 1590 (m), 1620 (s), 1793 (w), 1804 (w), 1825 (w), 1954 (w), 3073 (m).
1H NMR (300 MHz, CDCl3, 20 °C, δ/ppm): 1.42 (tBu, 18H), 1.47 (tBu, 18H), 6.28 (CHAr, 2H), 7.47 (CHAr, 1H, J = 1.74), 7.49 (CHAr, 1H, J = 1.88), 7.67 (CHAr, 2H, J = 1.46), 8.73 (CHAr, 2H, J = 5.99).
Complex 3∙toluene, [CoII(3,6-Cat)(bipytBu)]2∙toluene. A colorless solution of 4,4′-di-tert-butyl-2,2′-bipyridyl (0.5 mg, 1.86 mmol) in dry THF (5 mL) was mixed with a blue suspension of an anhydrous CoCl2 (0.24 mg, 1.86 mmol) in the same solvent (10 mL) in glovebox, producing an ink-colored solution. Interaction between the resulting mixture and a pale solution (dry THF, 10 mL) of prepared in situ CatNa2 (1.86 mmol) in evacuated ampoule facilitated an immediate turn to a green color of reaction. Solvent was changed to dry CH2Cl2 with a subsequent removing of formed white NaCl precipitate, after that the filtrate was stored in refrigerator for an hour. Obtained purple fine-crystalline product was separated by filtration in vacuum, and dried. The total yield for product [CoII(3,6-Cat)(bipytBu)]2∙CH2Cl2 with C65H90Cl2Co2N4O4 formula is 1.41 g (78%) based on the starting ligands.
A hot toluene solution of complex was evaporated slowly under anaerobic conditions to grow single crystals ([CoII(3,6-Cat)(bipytBu)]2∙toluene, C71H96Co2N4O4 composition) suitable for X-ray diffraction measurements.
The synthesis in situ of CatNa2 was carried out through the direct interaction of 3,6-DTBQ with an excess of metallic Na in THF. Reagents were stirred for an hour in evacuated ampoule, while the bright-green initial color of the reaction mixture turned to pale. The corresponding technique is well-known and described earlier [115].
The polycrystalline samples of 3 for an elemental analysis, IR-, EPR-, UV-vis-NIR spectroscopy, cyclic voltammetry, and magnetic susceptibility measurements were prepared from CH2Cl2.
Anal. Calc. for the complex [CoII(3,6-Cat)(bipytBu)]2∙CH2Cl2 with C65H90Cl2Co2N4O4 composition (%): C, 66.15; H, 7.69; N, 4.75. Found: C, 65.74; H, 7.87; N, 4.95.
IR (Nujol, KBr) cm−1: 486 (m), 501 (m), 539 (m), 546 (w), 559 (s), 608 (s), 652 (s), 676 (s), 696 (m), 742 (s), 760 (w), 785 (s), 804 (m), 847 (s), 862 (m), 887 (m), 900 (m), 922 (m), 931 (s), 972 (s), 1016 (m), 1080 (w), 1121 (m), 1144 (s), 1170 (w), 1192 (m), 1198 (w), 1214 (s), 1249 (s), 1265 (w), 1277 (m), 1346 (w), 1367 (m), 1394 (s), 1407 (m), 1478 (s), 1530 (m), 1551 (s), 1567 (w), 1614 (s), 1694 (w), 1748 (w), 1770 (w), 1823 (m), 1843 (w), 1948 (w), 1969 (w), 3046 (w), 3065 (m).

4. Conclusions

A novel strategy within a molecular design of CT chromophores aimed at the CT energy management was implemented by the example of new “α-diimine-MII-catecholate” CT dyes 13 (where M is 3d-row transition metals—Cu, Ni, Co): the varying of the metal center’s nature (electronic structure of d-level) was performed. As regards the preparation of CT dyes, three synthetic routes were applied and optimized, providing a high yield of complexes with different composition and molecular structure. So, the chosen approaches allow obtaining the square-planar monomeric LL’CT species (CuII and NiII derivatives 1 and 2, respectively), as well as dimeric CoII (M + L)L’CT chromophore 3 containing two “α-diimine-MII-catecholate” moieties united through the donor ability of oxygen atoms from catecholate dianions.
Corresponding changes in CT energy were investigated in detail by the synergy of experimental and theoretic tools, namely UV-vis-NIR spectroscopy, cyclic voltammetry, and DFT study. In particular, a considerable red CT band’s displacement at approximately 130–150 nm was achieved by the turn from CuII to CoII center with the same ligand platforms (neutral 4,4′-di-tert-butyl-2,2′-bipyridyl and 3,6-di-tert-butyl-catecholate dianion). Moreover, a fine-tunable interrelation between CT energy and solvent polarity was established, expressed in a pronounced negative solvatochromic effect with substantial blue CT band’s shifts. Total values are equal to 153, 137, and 58 nm for complexes CuII (1), NiII (2), and CoII (3), respectively (from toluene to CH3CN for 1, 2; from THF to CH3CN for 3).

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules27238175/s1, Figure S1: The fragment of crystal packing for complex 1—the molecules from odd and even layers are represented as bright and faded pictures, respectively. Structures are given with 30% thermal probability ellipsoids. Hydrogen atoms, tert-butyl groups and THF molecules are omitted for clarity; Figure S2: The shortest intra-stack contacts CdCat(A)Cdbipy(B), CdCat(B)Cdbipy(A), and Ni(1A)⋯Ni(1B) for complex 2. Structures are given with 30% thermal probability ellipsoids. Hydrogen atoms and tert-butyl groups are omitted for clarity; Figure S3. The shortest contact CdCat⋯C(23) = 3.695 Å for complex 3. Structures are given with 30% thermal probability ellipsoids. Hydrogen atoms and tert-butyl groups are omitted for clarity; Table S1: X-ray diffraction data collection and structure refinement for complexes 13. CIF and CheckCIF files for 1 and 2. Crystallographic data for the structural analysis has been deposited with the Cambridge Crystallographic Data Centre, CCDC 2213958 (1∙THF), 2213959 (2) and 2213960 (3∙toluene). Copies of the above information may be received free of charge at ccdc.cam.ac.uk/getstructures from the Cambridge Crystallographic Data Centre. References [113,116,117,118,119,120,121] are cited in the supplementary materials.

Author Contributions

Conceptualization, supervision, writing—review and editing, A.V.P.; synthesis and investigation, formal analysis and discussion, writing—original draft preparation, K.I.P.; synthesis and investigation, I.V.E. and O.Y.T.; formal analysis and discussion, R.V.R., G.K.F., A.S.B., and M.V.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation, grant number 22-13-00351.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article and supplementary materials.

Acknowledgments

The studies were carried out using the equipment of the center for collective use “Analytical Center of the IOMC RAS” with the financial support of the grant “Ensuring the development of the material and technical infrastructure of the centers for collective use of scientific equipment” (Unique identifier RF—2296.61321X0017, Agreement Number 075-15-2021-670).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of compounds 13 are available from the authors.

References

  1. Bechtold, T.; Mussak, R. Handbook of Natural Colorants; John Wiley & Sons: New York, NY, USA, 2009. [Google Scholar]
  2. Zollinger, H. Color Chemistry: Syntheses, Properties, and Applications of Organic Dyes and Pigments; John Wiley & Sons: New York, NY, USA, 2003. [Google Scholar]
  3. Weil, T.; Vosch, T.; Hofkens, J.; Peneva, K.; Müllen, K. The rylene colorant family—tailored nanoemitters for photonics research and applications. Angew. Chem. Int. Ed. 2010, 49, 9068–9093. [Google Scholar] [CrossRef] [PubMed]
  4. Giribabu, L.; Kanaparthi, R.K.; Velkannan, V. Molecular engineering of sensitizers for dye-sensitized solar cell applications. Chem. Rec. 2012, 12, 306–328. [Google Scholar] [CrossRef] [PubMed]
  5. Smirnova, E.; Besedina, M.; Karushev, M.; Vasil’ev, V.; Timonov, A. Photogalvanic and photovoltaic effects in systems based on metal complexes of Schiff bases. Russ. J. Phys. Chem. A 2016, 90, 1088–1094. [Google Scholar] [CrossRef]
  6. Sekar, N.; Gehlot, V.Y. Metal complex dyes for dye-sensitized solar cells: Recent developments. Resonance 2010, 15, 819–831. [Google Scholar] [CrossRef]
  7. Neuthe, K.; Popeney, C.; Bialecka, K.; Hinsch, A.; Sokolowski, A.; Veurmann, W.; Haag, R. Simple NIR complexes and their applicability in dye-sensitized solar cells. Polyhedron 2014, 81, 583–587. [Google Scholar] [CrossRef]
  8. Shubbak, M.H. Advances in solar photovoltaics: Technology review and patent trends. Renew. Sustain. Energy Rev. 2019, 115, 109383. [Google Scholar] [CrossRef]
  9. Tyagi, V.; Rahim, N.A.; Rahim, N.; Jeyraj, A.; Selvaraj, L. Progress in solar PV technology: Research and achievement. Renew. Sustain. Energy Rev. 2013, 20, 443–461. [Google Scholar] [CrossRef]
  10. Archer, S.; Weinstein, J.A. Charge-separated excited states in platinum(II) chromophores: Photophysics, formation, stabilization and utilization in solar energy conversion. Coord. Chem. Rev. 2012, 256, 2530–2561. [Google Scholar] [CrossRef]
  11. Tahara, K.; Kadowaki, T.; Kikuchi, J.-I.; Ozawa, Y.; Yoshimoto, S.; Abe, M. Synthesis and characterization of a new series of binuclear Pd(II) biscatecholato complexes: Non-innocent ligand-based approach to a wide range of variation in near-infrared absorptions of mixed-valence complexes. Bull. Chem. Soc. Jpn. 2018, 91, 1630–1639. [Google Scholar] [CrossRef] [Green Version]
  12. Baise, A.; Teucher, I.; Labes, M. Effect of charge-transfer acceptors on dynamic scattering in a nematic liquid crystal. Appl. Phys. Lett. 1972, 21, 142–143. [Google Scholar] [CrossRef]
  13. Krebs, P.; Sackmann, E.; Schwarz, J. Triplet ESR-spectra of charge-transfer complexes in liquid and single crystals. Chem. Phys. Lett. 1971, 8, 417–420. [Google Scholar] [CrossRef]
  14. Yam, V.W.-W.; Au, V.K.-M.; Leung, S.Y.-L. Light-emitting self-assembled materials based on d8 and d10 transition metal complexes. Chem. Rev. 2015, 115, 7589–7728. [Google Scholar] [CrossRef]
  15. Shavaleev, N.M.; Accorsi, G.; Virgili, D.; Bell, Z.R.; Lazarides, T.; Calogero, G.; Armaroli, N.; Ward, M.D. Syntheses and crystal structures of dinuclear complexes containing d-block and f-block luminophores. Sensitization of NIR luminescence from Yb(III), Nd(III), and Er(III) centers by energy transfer from Re(I)− and Pt(II)−bipyrimidine metal centers. Inorg. Chem. 2005, 44, 61–72. [Google Scholar] [CrossRef]
  16. Baggaley, E.; Weinstein, J.A.; Williams, J.G. Lighting the way to see inside the live cell with luminescent transition metal complexes. Coord. Chem. Rev. 2012, 256, 1762–1785. [Google Scholar] [CrossRef]
  17. Heinze, K.; Reinhardt, S. Platinum(II) complexes with non-innocent ligands: Solid-phase synthesis, redox chemistry and luminescence. Chem. Eur. J. 2008, 14, 9482–9486. [Google Scholar] [CrossRef]
  18. McKenzie, L.K.; Bryant, H.E.; Weinstein, J.A. Transition metal complexes as photosensitisers in one-and two-photon photodynamic therapy. Coord. Chem. Rev. 2019, 379, 2–29. [Google Scholar] [CrossRef] [Green Version]
  19. Mortimer, R.J. Electrochromic materials. Chem. Soc. Rev. 1997, 26, 147–156. [Google Scholar] [CrossRef]
  20. Rosseinsky, D.R.; Mortimer, R.J. Electrochromic systems and the prospects for devices. Adv. Mater. 2001, 13, 783–793. [Google Scholar] [CrossRef]
  21. Bange, K.; Gambke, T. Electrochromic materials for optical switching devices. Adv. Mater. 1990, 2, 10–16. [Google Scholar] [CrossRef]
  22. Monk, P.; Mortimer, R.; Rosseinsky, D. Electrochromism: Fundamentals and Applications, Angewandte Chemie-English Edition; Wiley-VCH: Weinheim, Germany, 1995. [Google Scholar]
  23. Granqvist, C.G. Handbook of inorganic Electrochromic Materials; Elsevier: Amsterdam, The Netherlands, 1995. [Google Scholar]
  24. Mortimer, R.J. Organic electrochromic materials. Electrochim. Acta 1999, 44, 2971–2981. [Google Scholar] [CrossRef]
  25. Granqvist, C. Electrochromism and smart window design. Solid State 1992, 53, 479–489. [Google Scholar] [CrossRef]
  26. Nejad, M.A.F.; Ranjbar, S.; Parolo, C.; Nguyen, E.P.; Álvarez-Diduk, R.; Hormozi-Nezhad, M.R.; Merkoçi, A. Electrochromism: An emerging and promising approach in (bio)sensing technology. Today 2021, 50, 476–498. [Google Scholar]
  27. Ward, M.D. Near-infrared electrochromic materials for optical attenuation based on transition-metal coordination complexes. J. Solid State Electrochem. 2005, 9, 778–787. [Google Scholar] [CrossRef]
  28. Sobottka, S.; Nößler, M.; Ostericher, A.L.; Hermann, G.; Subat, N.Z.; Beerhues, J.; Behr-van der Meer, M.; Suntrup, L.; Albold, U.; Hohloch, S. Tuning PtII-based donor–acceptor systems through ligand design: Effects on frontier orbitals, redox potentials, UV/Vis/NIR absorptions, electrochromism, and photocatalysis. Chemistry 2020, 26, 1314. [Google Scholar] [CrossRef] [PubMed]
  29. Resch-Genger, U.; Grabolle, M.; Cavaliere-Jaricot, S.; Nitschke, R.; Nann, T. Quantum dots versus organic dyes as fluorescent labels. Nat. Methods 2008, 5, 763–775. [Google Scholar] [CrossRef]
  30. Pashanova, K.I.; Poddel′sky, A.I.; Piskunov, A.V. Complexes of “late” transition metals of the 3d row based on functionalized o-iminobenzoquinone type ligands: Interrelation of molecular and electronic structure, magnetic behaviour. Coord. Chem. Rev. 2022, 459, 214399. [Google Scholar] [CrossRef]
  31. Poddel’sky, A.I.; Cherkasov, V.K.; Abakumov, G.A. Transition metal complexes with bulky 4,6-di-tert-butyl-N-aryl(alkyl)-o-iminobenzoquinonato ligands: Structure, EPR and magnetism. Coord. Chem. Rev. 2009, 253, 291–324. [Google Scholar] [CrossRef]
  32. Pashanova, K.I.; Bitkina, V.O.; Yakushev, I.A.; Arsenyev, M.V.; Piskunov, A.V. Square-planar heteroleptic complexes of α-diimine-NiII-catecholate type: Intramolecular ligand-to-ligand charge transfer. Molecules 2021, 26, 4622. [Google Scholar] [CrossRef] [PubMed]
  33. Atallah, H.; Taliaferro, C.M.; Wells, K.A.; Castellano, F.N. Photophysics and ultrafast processes in rhenium(I) diimine dicarbonyls. Dalton Trans. 2020, 49, 11565–11576. [Google Scholar] [CrossRef] [PubMed]
  34. García-Cañadas, J.; Meacham, A.P.; Peter, L.M.; Ward, M.D. A Near-infrared electrochromic window based on an Sb-doped SnO2 electrode modified with a Ru–dioxolene complex. Angew. Chem. 2003, 115, 3119–3122. [Google Scholar] [CrossRef]
  35. Tsai, M.-K.; Rochford, J.; Polyansky, D.E.; Wada, T.; Tanaka, K.; Fujita, E.; Muckerman, J.T. Characterization of redox states of Ru(OH2)(Q)(tpy)2+ (Q = 3,5-di-tert-butyl-1,2-benzoquinone, tpy = 2, 2′: 6′, 2″-terpyridine) and related species through experimental and theoretical studies. Inorg. Chem. 2009, 48, 4372–4383. [Google Scholar] [CrossRef] [PubMed]
  36. Michaels, H.; Benesperi, I.; Edvinsson, T.; Muñoz-Garcia, A.B.; Pavone, M.; Boschloo, G.; Freitag, M. Copper complexes with tetradentate ligands for enhanced charge transport in dye-sensitized solar cells. Inorganics 2018, 6, 53. [Google Scholar] [CrossRef] [Green Version]
  37. Grätzel, M. Solar energy conversion by dye-sensitized photovoltaic cells. Inorg. Chem. 2005, 44, 6841–6851. [Google Scholar] [CrossRef]
  38. O’Regan, B.; Grätzel, M. A Low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. In Renewable Energy; Routledge: London, UK, 2018; pp. 208–213. [Google Scholar]
  39. Miao, Q.; Gao, J.; Wang, Z.; Yu, H.; Luo, Y.; Ma, T. Syntheses and characterization of several nickel bis(dithiolene) complexes with strong and broad near-IR absorption. Inorg. Chim. Acta 2011, 376, 619–627. [Google Scholar] [CrossRef]
  40. Linfoot, C.L.; Richardson, P.; McCall, K.L.; Durrant, J.R.; Morandeira, A.; Robertson, N. A nickel-complex sensitiser for dye-sensitised solar cells. Sol. Energy 2011, 85, 1195–1203. [Google Scholar] [CrossRef] [Green Version]
  41. Cameron, L.A.; Ziller, J.W.; Heyduk, A.F. Near-IR absorbing donor–acceptor ligand-to-ligand charge-transfer complexes of nickel(II). Chem. Sci. 2016, 7, 1807–1814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Kramer, W.W.; Cameron, L.A.; Zarkesh, R.A.; Ziller, J.W.; Heyduk, A.F. Donor–acceptor ligand-to-ligand charge-transfer coordination complexes of nickel(II). Inorg. Chem. 2014, 53, 8825–8837. [Google Scholar] [CrossRef] [PubMed]
  43. Bubnov, M.P.; Teplova, I.A.; Druzhkov, N.O.; Fukin, G.K.; Cherkasova, A.V.; Cherkasov, V.K. Catecholato complexes of cobalt and nickel with 1,4-disubstituted-1,4-diazabutadiens-1,3 and 1,2-bis(diphenylphosphino) ethane. J. Chem. Sci. 2015, 127, 527–535. [Google Scholar] [CrossRef] [Green Version]
  44. Brown, D.G.; Reinprecht, J.T.; Vogel, G.C. Synthesis and characterization of copper(II) catecholato complexes. Inorg. Nucl. Chem. Lett. 1976, 12, 399–404. [Google Scholar] [CrossRef]
  45. Brown, D.G.; Hughes, W.J. Copper(II) complexes containing 3-n-nonylcatechol: Preparation and ligand oxidation. Z. Für Nat. B 1979, 34, 1408–1412. [Google Scholar] [CrossRef]
  46. Brown, D.G.; Hughes, W.J.; Knerr, G. Electronic and X-ray photoelectron spectra of copper catecholate complexes. Inorg. Chim. Acta 1980, 46, 123–126. [Google Scholar] [CrossRef]
  47. Buchanan, R.M.; Wilson-Blumenberg, C.; Trapp, C.; Larsen, S.K.; Greene, D.L.; Pierpont, C.G. Counter ligand dependence of charge distribution in copper-quinone complexes. Structural and magnetic properties of (3, 5-di-tert-butylcatecholato)(bipyridine)copper(II). Inorg. Chem. 1986, 25, 3070–3076. [Google Scholar] [CrossRef]
  48. Poddel’sky, A.; Smolyaninov, I.; Skatova, A.; Lukoyanov, A.; Fukin, G.; Berberova, N.; Cherkasov, V.; Abakumov, G. Diradical bis-o-iminosemiquinonato zinc complex: Spectroscopy, magneto-and electrochemistry. Z. Für Anorg. Und Allg. Chem. 2008, 634, 1154–1160. [Google Scholar] [CrossRef]
  49. Spiro, T.G. Zinc Enzymes. Curr. Opin. Chem. Biol. 1983, 2, 222–234. [Google Scholar]
  50. Alsfasser, R.; Powell, A.K.; Vahrenkamp, H. Functional tetrahedral zinc complexes. Angew. Chem. Int. Ed. Engl. 1990, 29, 898–899. [Google Scholar] [CrossRef]
  51. Brock, S.L.; Kauzlarich, S.M. A2Zn3As2O2 (A = Ba, Sr): A rare example of square planar zinc. Inorg. Chem. 1994, 33, 2491–2492. [Google Scholar] [CrossRef]
  52. Rakshit, R.; Mukherjee, C. Secondary interactions versus intramolecular π–π interactions in CuII–diradical complexes. Eur. J. Inorg. Chem. 2016, 2016, 2731–2737. [Google Scholar] [CrossRef]
  53. Poddel’sky, A.I.; Bubnov, M.P.; Fukin, G.K.; Cherkasov, V.K.; Abakumov, G.A. New nitrosyl bis-o-iminobenzosemiquinonato complexes of M(ISQ)2(NO) type. Z. Für Anorg. Und Allg. Chem. 2008, 634, 1205–1209. [Google Scholar] [CrossRef]
  54. Poddel’sky, A.I.; Cherkasov, V.K.; Fukin, G.K.; Bubnov, M.P.; Abakumova, L.G.; Abakumov, G.A. New four-and five-coordinated complexes of cobalt with sterically hindered o-iminobenzoquinone ligands: Synthesis and structure. Inorg. Chim. Acta 2004, 357, 3632–3640. [Google Scholar] [CrossRef]
  55. Piskunov, A.V.; Pashanova, K.I.; Bogomyakov, A.S.; Smolyaninov, I.V.; Starikov, A.G.; Fukin, G.K. Cobalt complexes with hemilabile o-iminobenzoquinonate ligands: A novel example of redox-induced electron transfer. Dalton Trans. 2018, 47, 15049–15060. [Google Scholar] [CrossRef] [PubMed]
  56. Paul, G.C.; Ghorai, S.; Mukherjee, C. Monoradical-containing four-coordinate Co(III) complexes: Homolytic S–S and Se–Se bond cleavage and catalytic isocyanate to urea conversion under sunlight. Chem. Commun. 2017, 53, 8022–8025. [Google Scholar] [CrossRef] [PubMed]
  57. Piskunov, A.; Pashanova, K.; Ershova, I.; Bogomyakov, A.; Starikov, A.; Cherkasov, A. Synthesis of four-, five-, and six-coordinate cobalt(III) bis-o-iminobenzosemiquinone complexes. Russ. Chem. Bull. 2019, 68, 757–769. [Google Scholar] [CrossRef]
  58. Bill, E.; Bothe, E.; Chaudhuri, P.; Chlopek, K.; Herebian, D.; Kokatam, S.; Ray, K.; Weyhermüller, T.; Neese, F.; Wieghardt, K. Molecular and electronic structure of four-and five-coordinate cobalt complexes containing two o-phenylenediamine- or two o-aminophenol-type ligands at various oxidation levels: An experimental, density functional, and correlated ab initio study. Chem. Eur. J. 2005, 11, 204–224. [Google Scholar] [CrossRef] [PubMed]
  59. Ershova, I.V.; Smolyaninov, I.V.; Bogomyakov, A.S.; Fedin, M.V.; Starikov, A.G.; Cherkasov, A.V.; Fukin, G.K.; Piskunov, A.V. Tetrahedral nickel(II) and cobalt(II) bis-o-iminobenzosemiquinonates. Dalton Trans. 2019, 48, 10723–10732. [Google Scholar] [CrossRef] [PubMed]
  60. Best, J.; Sazanovich, I.V.; Adams, H.; Bennett, R.D.; Davies, E.S.; Meijer, A.J.; Towrie, M.; Tikhomirov, S.A.; Bouganov, O.V.; Ward, M.D. Structure and ultrafast dynamics of the charge-transfer excited state and redox activity of the ground state of mono-and binuclear platinum(II) diimine catecholate and bis-catecholate complexes: A transient absorption, TRIR, DFT, and electrochemical study. Inorg. Chem. 2010, 49, 10041–10056. [Google Scholar] [CrossRef] [PubMed]
  61. Cummings, S.D.; Eisenberg, R. Tuning the excited-state properties of platinum(II) diimine dithiolate complexes. J. Am. Chem. Soc. 1996, 118, 1949–1960. [Google Scholar] [CrossRef]
  62. Scattergood, P.A.; Jesus, P.; Adams, H.; Delor, M.; Sazanovich, I.V.; Burrows, H.D.; Serpa, C.; Weinstein, J.A. Exploring excited states of Pt(II) diimine catecholates for photoinduced charge separation. Dalton Trans. 2015, 44, 11705–11716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Shavaleev, N.M.; Davies, E.S.; Adams, H.; Best, J.; Weinstein, J.A. Platinum(II) diimine complexes with catecholate ligands bearing imide electron-acceptor groups: Synthesis, crystal structures, (spectro)electrochemical and EPR studies, and electronic structure. Inorg. Chem. 2008, 47, 1532–1547. [Google Scholar] [CrossRef]
  64. Weinstein, J.A.; Tierney, M.T.; Davies, E.S.; Base, K.; Robeiro, A.A.; Grinstaff, M.W. Probing the electronic structure of platinum(II) chromophores: Crystal structures, NMR structures, and photophysical properties of six new bis- and di-phenolate/thiolate Pt(II) diimine chromophores. Inorg. Chem. 2006, 45, 4544–4555. [Google Scholar] [CrossRef]
  65. Tichnell, C.R.; Daley, D.R.; Stein, B.W.; Shultz, D.A.; Kirk, M.L.; Danilov, E.O. Wave function control of charge-separated excited-state lifetimes. J. Am. Chem. Soc. 2019, 141, 3986–3992. [Google Scholar] [CrossRef]
  66. Yang, J.; Kersi, D.K.; Giles, L.J.; Stein, B.W.; Feng, C.; Tichnell, C.R.; Shultz, D.A.; Kirk, M.L. Ligand control of donor–acceptor excited-state lifetimes. Inorg. Chem. 2014, 53, 4791–4793. [Google Scholar] [CrossRef] [PubMed]
  67. Yamada, S.; Matsumoto, T.; Chang, H.C. Impact of group 10 metals on the solvent-induced disproportionation of o-semiquinonato complexes. Chem. Eur. J. 2019, 25, 8268–8278. [Google Scholar] [CrossRef] [PubMed]
  68. BaniKhaled, M.O.; Becker, J.D.; Koppang, M.; Sun, H. Perfluoroalkylation of square-planar transition metal complexes: A strategy to assemble them into solid state materials with a π–π stacked lamellar structure. Cryst. Growth Des. 2016, 16, 1869–1878. [Google Scholar] [CrossRef]
  69. Sun, X.; Chun, H.; Hildenbrand, K.; Bothe, E.; Weyhermüller, T.; Neese, F.; Wieghardt, K. o-Iminobenzosemiquinonato(1−) and o-amidophenolato(2−) complexes of palladium(II) and platinum(II): A combined experimental and density functional theoretical study. Inorg. Chem. 2002, 41, 4295–4303. [Google Scholar] [CrossRef]
  70. Deibel, N.; Schweinfurth, D.; Fiedler, J.; Záliš, S.; Sarkar, B. Isomeric separation in donor–acceptor systems of Pd(II) and Pt(II) and a combined structural, electrochemical and spectroelectrochemical study. Dalton Trans. 2011, 40, 9925–9934. [Google Scholar] [CrossRef] [Green Version]
  71. Tahara, K.; Ashihara, Y.; Higashino, T.; Ozawa, Y.; Kadoya, T.; Sugimoto, K.; Ueda, A.; Mori, H.; Abe, M. New π-extended catecholato complexes of Pt(II) and Pd(II) containing a benzothienobenzothiophene (BTBT) moiety: Synthesis, electrochemical behavior and charge transfer properties. Dalton Trans. 2019, 48, 7367–7377. [Google Scholar] [CrossRef]
  72. Ghosh, P.; Begum, A.; Herebian, D.; Bothe, E.; Hildenbrand, K.; Weyhermüller, T.; Wieghardt, K. Coordinated o-dithio-and o-iminothiobenzosemiquinonate(1−) π radicals in [MII(bpy)(L.)](PF6) complexes. Angew. Chem. Int. Ed. 2003, 42, 563–567. [Google Scholar] [CrossRef]
  73. Yang, H.; Zhao, Y.; Liu, B.; Su, J.-H.; Fedushkin, I.L.; Wu, B.; Yang, X.-J. Noninnocent ligands: Heteroleptic nickel complexes with α-diimine and 1,2-diketone derivatives. Dalton Trans. 2017, 46, 7857–7865. [Google Scholar] [CrossRef]
  74. Rauth, G.K.; Pal, S.; Das, D.; Sinha, C.; Slawin, A.M.; Woollins, J.D. Synthesis, spectral characterization and electrochemical studies of mixed-ligand complexes of platinum(II) with 2-(arylazo)pyridines and catechols. Single-crystal X-ray structure of dichloro{2-(phenylazo) pyridine}platinum(II). Polyhedron 2001, 20, 363–372. [Google Scholar] [CrossRef]
  75. Roy, R.; Chattopadhyay, P.; Sinha, C.; Chattopadhyay, S. Synthesis, spectral and electrochemical studies of arylazopyridine complexes of palladium(II) with dioxolenes. Polyhedron 1996, 15, 3361–3369. [Google Scholar] [CrossRef]
  76. Weinstein, J.; Zheligovskaya, N.; Mel’nikov, M. Spectroscopic (UV/VIS, resonance Raman) and spectroelectrochemical study of platinum(II) complexes with 2,2′-bipyridine and aromatic thiolate ligands. J. Chem. Soc. Dalton Trans. 1998, 15, 2459–2466. [Google Scholar] [CrossRef]
  77. Liu, W.; Heinze, K. Rhenium(I) and platinum(II) complexes with diimine ligands bearing acidic phenol substituents: Hydrogen-bonding, acid–base chemistry and optical properties. Dalton Trans. 2010, 39, 9554–9564. [Google Scholar] [CrossRef] [PubMed]
  78. Kirk, M.L.; Shultz, D.A.; Chen, J.; Hewitt, P.; Daley, D.; Paudel, S.; van Der Est, A. Metal ion control of photoinduced electron spin polarization in electronic ground states. J. Am. Chem. Soc. 2021, 143, 10519–10523. [Google Scholar] [CrossRef] [PubMed]
  79. Stein, B.W.; Tichnell, C.R.; Chen, J.; Shultz, D.A.; Kirk, M.L. Excited state magnetic exchange interactions enable large spin polarization effects. J. Am. Chem. Soc. 2018, 140, 2221–2228. [Google Scholar] [CrossRef] [PubMed]
  80. Kee, J.; Ng, Y.; Kulkarni, S.; Muduli, S.; Xu, K.; Ganguly, R.; Lu, Y.; Hirao, H.; Soo, H. Development of bis(arylimino)acenaphthene (BIAN) copper complexes as visible light harvesters for potential photovoltaic applications. Inorg. Chem. Front. 2016, 3, 651–662. [Google Scholar] [CrossRef] [Green Version]
  81. Barnham, K.J.; Djuran, M.I.; Frey, U.; Mazid, M.A.; Sadler, P.J. [Pd(CBDCA-O,O′)(NH3)2]: The PdII analogue of a platinum anticancer drug (CBDCA= cyclobutane-1,1-dicarboxylate). J. Chem. Soc. Chem. Commun. 1994, 65–66. [Google Scholar] [CrossRef]
  82. Beagley, B.; Cruickshank, D.; McAuliffe, C.; Pritchard, R.; Zaki, A.; Beddoes, R.; Cernik, R.; Mills, O. The crystal and molecular structure of cis-diammine-1,1-cyclobutanedicarboxoplatinum(II) [cis-Pt(NH3)2CBDCA]. Dynamic puckering of the cyclobutane ring. J. Mol. Struct. 1985, 130, 97–102. [Google Scholar] [CrossRef]
  83. Neidle, S.; Ismail, I.M.; Sadler, P.J. The structure of the antitumor complex cis-(diammino)(1,1-cyclobutanedicarboxylato)-Pt(II): X ray and NMR studies. J. Inorg. Biochem. 1980, 13, 205–212. [Google Scholar] [CrossRef]
  84. Mansuri-Torshizi, H.; Ghadimy, S.; Akbarzadeh, N. Synthesis, characterization, DNA binding and cytotoxic studies of platinum(II) and palladium(II) complexes of the 2, 2′-bipyridine and an anion of 1,1-cyclobutanedicarboxylic acid. Chem. Pharm. Bull. 2001, 49, 1517–1520. [Google Scholar] [CrossRef] [Green Version]
  85. Jin, V.X.; Ranford, J.D. Complexes of platinum (II) or palladium (II) with 1, 10-phenanthroline and amino acids. Inorg. Chim. Acta 2000, 304, 38–44. [Google Scholar] [CrossRef]
  86. Nobuo, O.; Yu, M.; Mamiko, O. (Benzene-1,2-diolato-κ2O,O′)(di-2-pyridylamine-κ2N,N′)palladium(II) hemihydrate. Acta Crystallogr. Sect. E Struct. Rep. Online 2006, 62, m2778–m2780. [Google Scholar] [CrossRef] [Green Version]
  87. Okabe, N.; Aziyama, T.; Odoko, M. (Benzene-1,2-diolato-κ2O,O′)(2,2′-bipyridine-κ2N,N′)palladium(II). Acta Crystallogr. Sect. E Struct. Rep. Online 2005, 61, m1943–m1945. [Google Scholar] [CrossRef]
  88. Okabe, N.; Aziyama, T.; Odoko, M. (1,2-Benzenediolato-κ2O,O′)(2,2′-biquinoline-κ2N,N′)palladium(II). Acta Crystallogr. Sect. E Struct. Rep. Online 2005, 61, m2154–m2156. [Google Scholar] [CrossRef] [Green Version]
  89. Okabe, N.; Hagihara, K.; Odoko, M.; Muranishi, Y. (2,2′-Bipyridine-κ2N,N′)(2,3-naphthalenediolato-κ2O,O′)palladium(II) and (2,2′-biquinoline-κ2N,N′)(2,3-naphthalenediolato-κ2O,O′)palladium(II). Acta Crystallogr. Sect. C Cryst. Struct. Commun. 2004, 60, m150–m152. [Google Scholar] [CrossRef] [Green Version]
  90. Okabe, N.; Muranishi, Y.; Aziyama, T. (1,2-Benzenediolato-κ2O,O′)(1,10-phenanthroline-κ2N,N′)palladium(II) dihydrate. Acta Crystallogr. Sect. E Struct. Rep. Online 2003, 59, m936–m938. [Google Scholar] [CrossRef] [Green Version]
  91. Imer, A.G.; Syan, R.H.B.; Gülcan, M.; Ocak, Y.S.; Tombak, A. The novel pyridine based symmetrical Schiff base ligand and its transition metal complexes: Synthesis, spectral definitions and application in dye sensitized solar cells (DSSCs). J. Mater. Sci. Mater. Electron. 2018, 29, 898–905. [Google Scholar] [CrossRef]
  92. Islam, A.; Sugihara, H.; Hara, K.; Singh, L.P.; Katoh, R.; Yanagida, M.; Takahashi, Y.; Murata, S.; Arakawa, H. New platinum(II) polypyridyl photosensitizers for TiO2 solar cells. New J. Chem. 2000, 24, 343–345. [Google Scholar] [CrossRef]
  93. Gauthier, S.; Caro, B.; Robin-Le Guen, F.; Bhuvanesh, N.; Gladysz, J.A.; Wojcik, L.; Le Poul, N.; Planchat, A.; Pellegrin, Y.; Blart, E. Synthesis, photovoltaic performances and TD-DFT modeling of push–pull diacetylide platinum complexes in TiO2 based dye-sensitized solar cells. Dalton Trans. 2014, 43, 11233–11242. [Google Scholar] [CrossRef]
  94. Liu, Q.; Zhu, N.; Ho, C.-L.; Fu, Y.; Lau, W.-S.; Xie, Z.; Wang, L.; Wong, W.-Y. Synthesis, characterization, photophysical and photovoltaic properties of new donor–acceptor platinum(II) acetylide complexes. J. Organomet. Chem. 2016, 812, 2–12. [Google Scholar] [CrossRef]
  95. Brasse, M.; Campora, J.; Davies, M.; Teuma, E.; Palma, P.; Alvarez, E.; Sanz, E.; Reyes, M.L. Integrating catalyst and co-catalyst design in olefin polymerization catalysis: Transferable dianionic ligands for the activation of late transition metal polymerization catalysts. Adv. Synth. Catal. 2007, 349, 2111–2120. [Google Scholar] [CrossRef]
  96. Abakumov, G.; Cherkasov, V.; Lobanov, A.; Razuvaev, G. Oxidation of metallic copper and silver by orthoquinones in organic solvents. Bull. Acad. Sci. USSR Div. Chem. Sci. 1984, 33, 1478–1485. [Google Scholar] [CrossRef]
  97. Thompson, J.S.; Calabrese, J.C. Copper-catechol chemistry. Synthesis, spectroscopy, and structure of bis(3,5-di-tert-butyl-o-semiquinato)copper (II). J. Am. Chem. Soc. 1986, 108, 1903–1907. [Google Scholar] [CrossRef]
  98. Buchanan, R.M.; Fitzgerald, B.J.; Pierpont, C.G. Semiquinone radical anion coordination to divalent cobalt and nickel. Structural features of the bis (3,5-di-tert-butyl-1,2-semiquinone)cobalt(II) tetramer. Inorg. Chem. 1979, 18, 3439–3444. [Google Scholar] [CrossRef]
  99. Abakumov, G.; Cherkasov, V.; Piskunov, A.; Meshcheryakova, I.; Maleeva, A.; Poddel’skii, A.; Fukin, G. Zinc molecular complexes with sterically hindered o-quinone and o-iminoquinone. Dokl. Chem. 2009, 427, 168–171. [Google Scholar] [CrossRef]
  100. Yang, L.; Powell, D.R.; Houser, R.P. Structural variation in copper(I) complexes with pyridylmethylamide ligands: Structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, 9, 955–964. [Google Scholar] [CrossRef] [PubMed]
  101. Okuniewski, A.; Rosiak, D.; Chojnacki, J.; Becker, B. Coordination polymers and molecular structures among complexes of mercury(II) halides with selected 1-benzoylthioureas. Polyhedron 2015, 90, 47–57. [Google Scholar] [CrossRef]
  102. Addison, A.W.; Rao, T.N.; Reedijk, J.; van Rijn, J.; Verschoor, G.C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua [1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc. Dalton Trans. 1984, 7, 1349–1356. [Google Scholar] [CrossRef]
  103. Lynch, M.W.; Buchanan, R.M.; Pierpont, C.G.; Hendrickson, D.N. Weak magnetic exchange interactions between paramagnetic metal ions and coordinated o-semiquinones in M(9, 10-phenanthrenesemiquinone)2(pyridine)2 [M = nickel(II) and cobalt(II)] and tetranuclear M4(o-semiquinone)8 complexes. Inorg. Chem. 1981, 20, 1038–1046. [Google Scholar] [CrossRef]
  104. Lange, C.W.; Conklin, B.J.; Pierpont, C.G. Radical superexchange in semiquinone complexes containing diamagnetic metal ions. 3,6-Di-tert-butyl-1,2-semiquinoate complexes of zinc(II), cobalt(III), gallium(III), and aluminum(III). Inorg. Chem. 1994, 33, 1276–1283. [Google Scholar] [CrossRef]
  105. Brown, S.N. Metrical oxidation states of 2-amidophenoxide and catecholate ligands: Structural signatures of metal–ligand π bonding in potentially noninnocent ligands. Inorg. Chem. 2012, 51, 1251–1260. [Google Scholar] [CrossRef]
  106. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A Cryst. Phys. Diffr. Theor. Gen. Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  107. Sarkar, B.; Hübner, R.; Pattacini, R.; Hartenbach, I. Combining two non-innocent ligands in isomeric complexes [Pt(pap)m Qn]0 (pap = phenylazopyridine, Q = 3,5-di-tert-butyl-benzoquinone. Dalton Trans. 2009, 4653–4655. [Google Scholar] [CrossRef] [PubMed]
  108. Deibel, N.; Schweinfurth, D.; Hohloch, S.; Fiedler, J.; Sarkar, B. Donor–acceptor systems of Pt(II) and redox-induced reactivity towards small molecules. Chem. Commun. 2012, 48, 2388–2390. [Google Scholar] [CrossRef] [PubMed]
  109. Wang, Y.; Mizubayashi, Y.; Odoko, M.; Okabe, N. (Di-2-pyridylamine-κ2N,N′)(naphthalene-2,3-diolato-κ2O,O′)palladium (II) monohydrate and (di-2-pyridylamine-κ2N,N′)(3-oxidonaphthalene-2-carboxylato-κ2O,O′)palladium(II). Acta Crystallogr. Sect. C Cryst. Struct. Commun. 2005, 61, m67–m70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Hathaway, B.; Billing, D. The electronic properties and stereochemistry of mono-nuclear complexes of the copper(II) ion. Coord. Chem. Rev. 1970, 5, 143–207. [Google Scholar] [CrossRef]
  111. Nippe, M.; Khnayzer, R.S.; Panetier, J.A.; Zee, D.Z.; Olaiya, B.S.; Head-Gordon, M.; Chang, C.J.; Castellano, F.N.; Long, J.R. Catalytic proton reduction with transition metal complexes of the redox-active ligand bpy2PYMe. Chem. Sci. 2013, 4, 3934–3945. [Google Scholar] [CrossRef] [Green Version]
  112. De La Torre, P.; Derrick, J.S.; Snider, A.; Smith, P.T.; Loipersberger, M.; Head-Gordon, M.; Chang, C.J. Exchange coupling determines metal-dependent efficiency for iron-and cobalt-catalyzed photochemical CO2 reduction. ACS Catal. 2022, 12, 8484–8493. [Google Scholar] [CrossRef]
  113. Reichardt, C.; Welton, T. Solvents and Solvent Effects in Organic Chemistry; Reichardt, C., Welton, T., Eds.; John Wiley & Sons: New York, NY, USA, 2011. [Google Scholar]
  114. Maleeva, A.V.; Ershova, I.V.; Trofimova, O.Y.; Arsenyeva, K.V.; Yakushev, I.A.; Piskunov, A.V. Near-IR absorbing donor–acceptor charge-transfer gallium complex, an example from non-transition metal chemistry. Mendeleev Commun. 2022, 32, 83–86. [Google Scholar] [CrossRef]
  115. Lado, A.; Piskunov, A.; Cherkasov, V.; Fukin, G.; Abakumov, G. Free radical fixation by tin(IV) diphenylcatecholate complexes. Russ. J. Coord. Chem. 2006, 32, 173–179. [Google Scholar] [CrossRef]
  116. Fukin, G.; Cherkasov, A.V.; Shurygina, M.P.; Druzhkov, N.O.; Kuropatov, V.A.; Chesnokov, S.A.; Abakumov, G.A. Geometrical and energetical aspects of structure of 3, 6-di-tert-butyl-o-benzoquinones. Struct. Chem. 2010, 21, 607–611. [Google Scholar]
  117. Gordon, A.; Ford, R. The Chemist’s Companion; Wiley & Sons: New York, NY, USA, 1972; p. 143. [Google Scholar]
  118. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.; Scalmani, R.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. GAUSSIAN 09; Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  119. Chilton, N.F.; Anderson, R.P.; Turner, D.; Soncini, A.; Murray, S. PHI: A powerful new program for the analysis of anisotropic monomeric and exchange-coupled polynuclear d-and f-block complexes. J. Comput. Chem. 2013, 34, 1164–1175. [Google Scholar]
  120. Rigaku Oxford Diffraction. CrysAlis Pro Software System, Version 1.171.41.122a; Rigaku Corporation: Wroclaw, Poland, 2021.
  121. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3. [Google Scholar]
Figure 1. Synthetic routes for complexes 13.
Figure 1. Synthetic routes for complexes 13.
Molecules 27 08175 g001
Figure 2. Views of (1), (2) and (3) molecular structures with 30% thermal probability ellipsoids. Hydrogen atoms are omitted for clarity.
Figure 2. Views of (1), (2) and (3) molecular structures with 30% thermal probability ellipsoids. Hydrogen atoms are omitted for clarity.
Molecules 27 08175 g002
Figure 3. The fragments of crystal packings for complexes 13. Faded views are given for: the molecules from even layer (in the case of compound 1); the molecules B (in the case of complex 2); one monomeric D-MII-A part (in the case of complex 3). Structures are given with 30% thermal probability ellipsoids. Hydrogen atoms and tert-butyl groups are omitted for clarity.
Figure 3. The fragments of crystal packings for complexes 13. Faded views are given for: the molecules from even layer (in the case of compound 1); the molecules B (in the case of complex 2); one monomeric D-MII-A part (in the case of complex 3). Structures are given with 30% thermal probability ellipsoids. Hydrogen atoms and tert-butyl groups are omitted for clarity.
Molecules 27 08175 g003
Figure 4. EPR spectra for complex 1 in solid state (a) and in CH2Cl2 solution (b) at room temperature.
Figure 4. EPR spectra for complex 1 in solid state (a) and in CH2Cl2 solution (b) at room temperature.
Molecules 27 08175 g004
Figure 5. The temperature dependence of the effective magnetic moment μeff(T) for complex 3. Solid line is a theoretical curve.
Figure 5. The temperature dependence of the effective magnetic moment μeff(T) for complex 3. Solid line is a theoretical curve.
Molecules 27 08175 g005
Figure 6. The CV data for complexes 13 (GC electrode, Ag/AgCl/KCl, 0.2 M Bu4NClO4, CH2Cl2, C = 2·10−3 mol/dm3, V = 0.2 V·s−1, Ar). Corrections to Fc/Fc+ couple as a standard were made.
Figure 6. The CV data for complexes 13 (GC electrode, Ag/AgCl/KCl, 0.2 M Bu4NClO4, CH2Cl2, C = 2·10−3 mol/dm3, V = 0.2 V·s−1, Ar). Corrections to Fc/Fc+ couple as a standard were made.
Molecules 27 08175 g006
Figure 7. UV-vis-NIR spectra for complexes 13.
Figure 7. UV-vis-NIR spectra for complexes 13.
Molecules 27 08175 g007
Figure 8. ETN(λ) dependences and linear fits for complexes 13.
Figure 8. ETN(λ) dependences and linear fits for complexes 13.
Molecules 27 08175 g008
Figure 9. Views of calculated frontier orbitals for complexes 1 (1, 3; isovalue = 0.02), and 2 (2, 4; 1, 3; isovalue = 0.02) and the spin density distribution for complex 1 (5; isovalue = 0.005).
Figure 9. Views of calculated frontier orbitals for complexes 1 (1, 3; isovalue = 0.02), and 2 (2, 4; 1, 3; isovalue = 0.02) and the spin density distribution for complex 1 (5; isovalue = 0.005).
Molecules 27 08175 g009
Figure 10. View of calculated frontier orbitals for complex 3 (isovalue = 0.03).
Figure 10. View of calculated frontier orbitals for complex 3 (isovalue = 0.03).
Molecules 27 08175 g010
Table 1. Characteristics of a coordination polyhedron and bridging lengths for dimer 3∙toluene and reported CuII and CoII oligomers on the base of 3,5-di-tert-butyl-o-benzoquinone type ligand.
Table 1. Characteristics of a coordination polyhedron and bridging lengths for dimer 3∙toluene and reported CuII and CoII oligomers on the base of 3,5-di-tert-butyl-o-benzoquinone type ligand.
Contact3∙Toluene[CuII(3,5-Cat)(bipy)]2 [47][CuII(3,5-SQ)2]2 [97][CoII(3,5-SQ)2]4 [98]
dimerdimerdimertetramer
M⋯M *, Å2.8847 (7)3.0742 (8)3.3603.163/3.306/3.317
M⋯O *, Å2.0222 (18)2.330 (2)2.416 (4)2.075/2.133/2.175
Coordination polyhedron
τ5 [102] **0.087,
square pyramid
0.228,
square pyramid
0.420,
square pyramid
-
----octahedron
Notes: *—atoms from the second part of oligomer (Symmetry transformations used to generate equivalent atoms: −x + 1, −y, −z + 1); **—calculated by the authors of this work; [CuII(3,5-Cat)(bipy)]2-bis((3,5-di-tert-butyl-catecholato)(2,2′-bipyridine)copper(II)); [CuII(3,5-SQ)2]2-bis((bis(3,5-di-tert-butyl-semiquinonato))(2,2′-bipyridine)copper(II)); [CoII(3,5-SQ)2]4-tetrakis((bis(3,5-di-tert-butyl-semiquinonato))(2,2′-bipyridine)cobalt(II)).
Table 2. Selected bond lengths for complexes 13.
Table 2. Selected bond lengths for complexes 13.
1·THF2 (Molecule A)/(Molecule B)3∙Toluene
BondÅBondÅBondÅ
Cu(1)-O(1)1.860 (2)Ni(1)-O(1)1.8137 (14)/1.8110 (14)Co(1)-O(1)1.9375 (18)
Cu(1)-O(2)1.857 (3)Ni(1)-O(2)1.8096 (15)/1.8118 (15)Co(1)-O(2)/Co(1)-O(2) *2.0552 (17)/2.0222 (18)
Cu(1)-N(1)1.975 (3)Ni(1)-N(1)1.8755 (17)/1.8773 (18)Co(1)-N(2)2.113 (2)
Cu(1)-N(2)1.975 (3)Ni(1)-N(2)1.8782 (17)/1.8768 (17)Co(1)-N(1)2.115 (2)
O(1)-C(1)1.348 (5)O(1)-C(1)1.353 (3)/1.351 (3)O(1)-C(1)1.335 (3)
O(2)-C(2)1.347 (4)O(2)-C(2)1.357 (2)/1.356 (2)O(2)-C(2)1.364 (3)
N(1)-C(13)1.335 (5)N(1)-C(15)1.342 (3)/1.343 (3)N(2)-C(24)1.340 (3)
N(1)-C(17)1.340 (4)N(1)-C(19)1.349 (3)/1.352 (3)N(2)-C(20)1.341 (3)
N(2)-C(18)1.346 (4)N(2)-C(20)1.355 (3)/1.349 (3)N(1)-C(19)1.344 (3)
N(2)-C(22)1.336 (5)N(2)-C(24)1.339 (3)/1.339 (3)N(1)-C(15)1.341 (3)
C(1)-C(2)1.408 (6)C(1)-C(2)1.411 (3)/1.406 (3)C(1)-C(2)1.434 (4)
C(2)-C(3)1.399 (6)C(2)-C(3)1.393 (3)/1.391 (3)C(2)-C(3)1.400 (4)
C(3)-C(4)1.388 (6)C(3)-C(4)1.393 (3)/1.397 (3)C(3)-C(4)1.401 (4)
C(4)-C(5)1.367 (7)C(4)-C(5)1.378(3)/1.383 (3)C(4)-C(5)1.367 (4)
C(5)-C(6)1.397 (6)C(5)-C(6)1.394 (3)/1.393 (3)C(5)-C(6)1.392 (4)
C(1)-C(6)1.401 (5)C(1)-C(6)1.392 (3)/1.396 (3)C(1)-C(6)1.407 (4)
Notes: *—atoms from the second part of [CoII(3,6-Cat)(bipytBu)]2 dimer (Symmetry transformations used to generate equivalent atoms: −x + 1, −y, −z + 1).
Table 3. Values of “metal-heteroatom” distances in some known monomeric “α-diimine-MII-catecholate” type complexes.
Table 3. Values of “metal-heteroatom” distances in some known monomeric “α-diimine-MII-catecholate” type complexes.
BondCuII(3,5-Cat)(bipy) [47]NiII(3,6-Cat)(bipy) [32]CoII(3,6-Cat)(DAD) [43]
monomermonomermonomer
M-O, Å1.901 (5), 1.870 (5)1.814 (3), 1.815 (3)1.8066 (13), 1.8104 (13)
M-N, Å1.993 (6), 1.999 (6)1.882 (4), 1.891 (3)1.8743 (16), 1.8711 (16)
Notes: CuII(3,5-Cat)(bipy)-(3,5-di-tert-butyl-catecholato)(2,2′-bipyridine)copper(II); NiII(3,6-Cat)(bipy)-(3,6-di-tert-butyl-catecholato)(2,2′-bipyridine)nickel(II); CoII(3,6-Cat)(DAD)-(3,6-di-tert-butyl-catecholato)(2,3-dimethyl-1,4-bis-(2,6-di-iso-propylphenyl)-1,4-diaza-1,3-butadiene)cobalt(II).
Table 4. The comparison of CV data for complex 2 and NiII(3,6-Cat)(bipy).
Table 4. The comparison of CV data for complex 2 and NiII(3,6-Cat)(bipy).
CompoundE1/2Red, VE1/2Ox, VΔEHOMO→LUMO, eV
2−2.01−0.201.81
NiII(3,6-Cat)(bipy)−1.86−0.211.65
Table 5. The characteristics of the CT peaks for complexes 13 in comparison with related known chromophores.
Table 5. The characteristics of the CT peaks for complexes 13 in comparison with related known chromophores.
Solvent1·THFNiII(3,5-Cat)
(bipytBu) [67]
2NiII(3,6-Cat)
(bipy) [32]
3∙TolueneNiII(3,6-CatGly)
(bipy) [32]
λ (ε)Eλ (ε)λ (ε)Eλ (ε)λ (ε)E
CH3CN501 (1.29)2.475-543 (3.94)2.284-657 (1.67)1.887-
DMF523 (1.28)2.371554 (1.83)560 (2.98)2.214580 (3.50)674 (2.74)1.840618 (3.08)
CH2Cl2540 (1.31)2.296572 (2.09)580 (4.89)2.138610 (4.35)695 (5.19)1.784685 (4.04)
THF583 (0.87)2.127618 (1.95)617 (3.98)2.010640 (3.43)715 (9.36)1.734720 (3.18)
Toluene654 (0.95)1.896683 (1.70)680 (2.24)1.824715 (3.57)insoluble-810 (3.38)
Parameters of solvatochromic shift
Δλ(Tol–CH3CN)153-137---
Δλ(Tol–DMF)131129120135-192
Δλ(THF–CH3CN)82-74-58-
Δλ(THF–DMF)6064576041102
Notes: λ—wavelength, nm; ε—extinction coefficient, 103 M−1·cm−1; E—LL’CT energy, eV; Δλ—solvatochromic shifts within two different solvents (given in brackets), nm.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pashanova, K.I.; Ershova, I.V.; Trofimova, O.Y.; Rumyantsev, R.V.; Fukin, G.K.; Bogomyakov, A.S.; Arsenyev, M.V.; Piskunov, A.V. Charge Transfer Chromophores Derived from 3d-Row Transition Metal Complexes. Molecules 2022, 27, 8175. https://doi.org/10.3390/molecules27238175

AMA Style

Pashanova KI, Ershova IV, Trofimova OY, Rumyantsev RV, Fukin GK, Bogomyakov AS, Arsenyev MV, Piskunov AV. Charge Transfer Chromophores Derived from 3d-Row Transition Metal Complexes. Molecules. 2022; 27(23):8175. https://doi.org/10.3390/molecules27238175

Chicago/Turabian Style

Pashanova, Kira I., Irina V. Ershova, Olesya Yu. Trofimova, Roman V. Rumyantsev, Georgy K. Fukin, Artem S. Bogomyakov, Maxim V. Arsenyev, and Alexandr V. Piskunov. 2022. "Charge Transfer Chromophores Derived from 3d-Row Transition Metal Complexes" Molecules 27, no. 23: 8175. https://doi.org/10.3390/molecules27238175

Article Metrics

Back to TopTop