Next Article in Journal
Recent Advances in the BiVO4 Photocatalyst for Sun-Driven Water Oxidation: Top-Performing Photoanodes and Scale-Up Challenges
Next Article in Special Issue
Highly Active and Selective Supported Rhenium Catalysts for Aerobic Oxidation of n-Hexane and n-Heptane
Previous Article in Journal
The Distribution and Strength of Brönsted Acid Sites on the Multi-Aluminum Model of FER Zeolite: A Theoretical Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

C-Homoscorpionate Oxidation Catalysts—Electrochemical and Catalytic Activity

by
Luísa M. D. R. S. Martins
1,2
1
Chemical Engineering Department, Instituto Superior de Engenharia de Lisboa, Instituto Politécnico de Lisboa, R. Conselheiro Emídio Navarro, 1959-007 Lisboa, Portugal
2
Centro de Química Estrutural, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049–001 Lisboa, Portugal
Catalysts 2017, 7(1), 12; https://doi.org/10.3390/catal7010012
Submission received: 18 November 2016 / Revised: 22 December 2016 / Accepted: 23 December 2016 / Published: 1 January 2017
(This article belongs to the Special Issue New Trends in Scorpionate Catalysts)

Abstract

:
A survey of the electrochemical properties of homoscorpionate tris(pyrazol-1-yl)methane complexes is presented. The relationship between structural features and catalytic efficiency toward the oxidative functionalization of inexpensive and abundant raw-materials to added-value products is also addressed.

Graphical Abstract

1. Introduction

Scorpionate compounds (Figure 1), in particular, poly(pyrazol-1-yl)borates, R1BXn(R2pz)3−n (pz = pyrazol-1-yl, n = 0 or 1), and poly(pyrazol-1-yl)methanes, R1CXn(R2pz)3−n (n = 0 or 1), are undoubtedly among the most important N-donor ligands in coordination chemistry [1,2,3,4,5,6,7,8]. The latter are considerably less well studied than the analogous borate species [3]. However, in the last two decades, mainly driven by improved syntheses [9,10], coordination behavior and physicochemical properties of poly(pyrazol-1-yl)methanes have attracted considerable interest [5,6,7,8] in order to perform the precise tuning of target scorpionates towards a desired function [3,6,7]. Applications of this highly versatile class of metal compounds range from organic synthesis, analytical, bio-inorganic or catalytic chemistry to material sciences [3,6,7,8,11,12,13,14,15,16,17,18].
The development of sustainable efficient catalytic processes for the activation of abundant and inexpensive raw-materials into high-added-value products remains a great challenge for both academic and industrial purposes. In this respect, the use of metal complexes bearing C-scorpionate poly(pyrazol-1-yl)methane ligands as catalysts is currently experiencing significant development [6,7,8,16]. Transition metals are important in this topic participating e.g., in redox processes, which can be applied in specific steps. The electronic interaction between transition metals and scorpionate ligands can play a key role in improving the redox process, and the type of scorpionate ligand can be determinant in achieving the desired properties in such complexes. Thus, one advantage of this catalytic system (over e.g., the metallocene based one) is the ease of modifying the scorpionate ligand to change the steric and electronic properties of the metal complex and therefore, its catalytic performance.
Industrially important reactions catalyzed by C-scorpionate complexes include [6,7,8,17,18,19,20,21,22] (i) mild partial oxidation of alkanes to alkyl hydroperoxides, alcohols and ketones; (ii) epoxidation of alkenes; (iii) oxidation of primary or secondary alcohols to aldehydes or ketones, respectively; (iv) the Baeyer-Villiger oxidation of linear or cyclic ketones to the corresponding esters and lactones, respectively; (v) the single pot carboxylation of gaseous alkanes into the corresponding Cn+1 carboxylic acids; (vi) the stereo-selective nitroaldol Henry C–C coupling reaction; and (vii) olefin polymerization.
Whereas the main catalytic applications of metal complexes with tris(pyrazol-1-yl)borates or heteroscorpionate ligands based on the bis(pyrazol-1-yl)methane moiety are found in olefin polymerization reactions [1,6], tris(pyrazol-1-yl)methane-type complexes of several transition metals are mainly used as catalysts or catalyst precursors for alkane, alkene, alcohol, and ketone oxidation reactions directed toward single-pot organic synthesis [7]. Their use as catalysts for the C–C coupling Henry reaction (a non-redox process) [8] has also proved to be a very promising strategy, in particular for those metals (e.g., Zn) that exhibit no redox flexibility but can behave as Lewis acid catalysts.
Moreover, tris(pyrazol-1-yl)methane metal complexes can exhibit remarkable versatile catalytic activity for oxidation reactions [16]. It is believed that the interchange between bidentate and tridentate coordination modes of the C-scorpionate ligands is at the core of the structural and chemical versatility of many metal complexes of this kind and is essential for their catalytic applications.
Electron transfer plays a fundamental role in governing the pathway of most of the above chemical reactions. In fact, the activity of metal-based catalysts depends largely on their ligand environment and coordination geometry, which also rule their oxidation/reduction properties, with the redox potential as a determining parameter. Thus, quantification of the net electron donation of the ligands to a metal center would allow predicting metal-centered redox potentials, and vice-versa, providing a powerful tool for the design of metal-based catalysts within a desired redox window.
Determination of redox potentials can be conveniently done by e.g., the easy and fast cyclic voltammetry technique, provided the redox signals lie within the available solvent/electrolyte potential window and the species have a sufficient lifetime for signal detection. However, to date, the useful information associated to the redox potential of a metal complex has not yet found a common application as a characterization or identification tool [23]. Moreover, a survey of the redox properties of known C-homoscorpionate metal complexes is missing.
Systematic approaches to establish redox potential/structure relationships, following the recognition of full additive ligand (L) effects on that potential have been proposed [24,25,26,27,28]. For example, Lever’s parametrization approach (Equation (1)) [27,28] allows for the prediction of an Mn+1/n redox potential (E) of a six-coordinate metal complex in V vs. SHE (standard hydrogen electrode), where EL is an additive ligand parameter obtained by a statistical analysis on the known redox potentials of a high number of Mn+1/n complexes [27,28]. The slope, SM, and intercept, IM, are dependent upon the metal and redox couple, the polygon of the complex, the spin state, and the stereochemistry [23].
E = SMEL) + IM/V vs. SHE
On the other hand, Equation (1) can be applied to estimate the EL value of a ligand (L) provided one knows the redox potential of a complex with that ligand L bound to a Mn+1/n metal redox couple with known IM and SM parameters, and the EL values of the co-ligands.
Herein, the electrochemical properties of homoscorpionate tris(pyrazol-1-yl)methane metal complexes that act as catalysts for the above industrial oxidation reactions are presented.
Moreover, the EL ligand parameter, a measure of the electron donor character of ligand L (the lower its value, the stronger that character), is used to establish redox/structure as well as redox/catalytic activity relationships, a very important tool for the design of improved catalysts to address some of the problems presented by current large-scale industrial partial oxidation processes.

2. C-Homoscorpionates and Their Metal Catalysts

Firstly reported by Trofimenko (1966) [29,30] as “a new and fertile field of remarkable scope”, B-scorpionate tris(pyrazol-1-yl)borates (Tp, Figure 2a) indeed were revealed to be a class of compounds that became a precious ligand system in modern coordination chemistry [1,2].
Although discovered earlier by Hückel et al. (1937) [31], the analogous C-scorpionate tris(pyrazol-1-yl)methanes (Tpm, Figure 2b) remained dormant with respect to coordination chemistry until 1966 [32], mainly due to synthetic difficulties and usually very low yields associated with the preparation of functionalized tris(pyrazol-1-yl)methanes where substituents on the pyrazolyl rings are larger than methyl. In fact, until Elguero’s report (1984) of an improved synthetic strategy [33] and its subsequent application in the formation of functionalized derivatives bearing bulky substituents [34,35], only few (less than 20) reports appeared pertaining to first-row transition metal complexes of tris(pyrazol-1-yl)methanes, mostly homoleptic ones.
It is commonly agreed to compare [2,3,30,36,37,38] the main characteristics of tris(pyrazol-1-yl) type scorpionate ligands with other face-capping ligands. In particular, the parallel between Tp and cyclopendadienyl (Cp, Figure 2c) ligands is established in that both are mononegative, six-electron (ionic model) or five-electron donor (covalent model) ligands. They are also formally isolobal [2,3]. The former are weak-field hard σ-N donors which tend to behave as fac-capping chelating ligands (i.e., occupy three coordination positions), while Cp are typically 5-fold π-donors and tend to form tetrahedral complexes [3,38].
Importantly, it has been shown that there is no systematic trend in comparative electron donor ability of Tp relative to Cp [39]. Their electron-donating abilities are dependent upon the identity and oxidation state of the metal center as well as the properties of the other ligands in the complex [40]. Tris(pyrazol-1-yl)borates are also bulkier than the formally analogous Cp and, in six-coordinate environments, enforce nearly octahedral coordination to the metal with N-M-N bite angles close to the ideal value (90°). This has been suggested to be the primary source of the different reactivity between comparable Tp and Cp complexes [40,41]. For example, the greater steric profile of Tp ligands has permitted the isolation of molecular species whose pentamethylcyclopentadienyl congeners proved too reactive [42]. In addition, Tp is coordinatively flexible, presenting κ2- or κ3-coordination modes (i.e., the scorpionate feature). The carbon analogues to Tp, tris(pyrazol-1-yl)methanes, maintain the tripodal face capping aspect and the same electro-donor ability, but differ from Tp and Cp in the charge they hold (Figure 2).
Since 2005 a considerable interest in the development of a fast an efficient synthetic route for hydrotris(pyrazol-1-yl)methane, HC(pz)3 (pz = pyrazol-1-yl), Tpm [10], as well as on the design and synthesis of brand new poly-functionalized C-homoscorpionates, R1C(R2pz)3, overcoming the lack in the chemistry of such species (Figure 3), has been found [43,44,45]. For example, new tris(pyrazol-1-yl)methanes functionalized at the methine carbon atom (in order to vary the coordination behavior and physicochemical properties) were successfully prepared: CH3SO3CH2C(pz)3 [43] or PyCH2OCH2C(pz)3 (Py = pyridine) [44]. The functionalization of pyrazol-1-yl rings (to modulate the coordination properties) was also achieved, as well as those derivatives that combine the two types of functionalization: e.g., SO3C(3-Phpz)3 [45], HOCH2C(3-Phpz)3 [44], or PyCH2OCH2C(3-Phpz)3 [44].
A systematic investigation of the coordination behavior of new C-scorpionates, as well as some of the known ones, toward a variety of transition metals (e.g., V [44,46,47,48,49], Mo [17,19,50], Re [51,52], Fe [44,46,49,53], Ru [54], Co [12,13], Ni [44,49,55], Pd [44], Cu [43,45,46,56,57], Ag [14], Au [58] or Zn [44,56]) followed, leading to new classes of complexes exhibiting different types of coordination modes (Figure 4 and Figure 5). Like the pincer of a scorpion, these versatile tripodal ligands bind metal centers with nitrogen atoms from two pyrazolyl rings attached to the central carbon atom; the third pyrazolyl attached to carbon rotates forward like a scorpion’s tail to “sting” the metal; hence the name of “scorpionates” (Figure 4).
Some of the above new C-homoscorpionates (with three identical pyrazol-1-yl rings), such as e.g., PyCH2OCH2C(pz)3 (Figure 5a) exhibit extended coordination ability, including tetradentate characteristics [44], where the extra coordination moiety has different affinity towards metal centers. Therefore, it leads to a sort of metal-supported scorpionate ligand that forms easily heterobimetallic species, opening to a large variety of applications (such as catalysis or supramolecular chemistry).
On the other hand, tris(pyrazol-1-yl)methane derivatives bearing bulky substituents at the pyrazol-1-yl rings (especially at the 3-position), when ligating a metal center, such a bulky species provides a steric control on the other coordination position(s) of the complex, selecting the suitable ligands on the opposite side, namely preventing the formation of full-sandwich complexes (with two such scorpionate ligands) [45]. Moreover, they also offer the opportunity to tailor the coordination behavior toward different metal centers. This important feature is directly correlated to further research in catalytic synthetic chemistry.
It was also found, from electrochemical experiments [57,59], that changes on the functionalized methine group of tris(pyrazol-1-yl) scorpionates have a much smaller influence on the ligand properties than when performed at the pyrazol-1-yl rings (see below).
The coordination versatility of tris(pyrazol-1-yl)methanes, namely the interchange between bidentate and tridentate coordination modes (Figure 5b,c for κ2-coordination) was found to be tuned by metal center as well as by the electronic properties of the co-ligands present at the coordination sphere. In addition, the tripodal functionalized coordination flexibility (e.g., N3- or N2O-coordination modes for sulfonated derivatives, Figure 5d) involving the functionalized methine carbon [7,8,14,45,49,50] is also tailored by such co-ligands, an important pre-requisite for their catalytic activity.
One might consider that the coordination behavior of the tris(pyrazol-1-yl)methane complexes would mirror that exhibited by the corresponding tris(pyrazol-1-yl)borate compounds, the major difference being in the charge between the methane and the borate counterpart. However, large differences appear in some cases [60]: for example, the RC(pz)3 ligands react with Group 6 metal hexacarbonyls to afford insoluble and non-volatile species, whereas [M{RB(pz)3}(CO)3] are very soluble and sublime easily. The tripodal ligand HC(pz)3 produces a relatively strong ligand field, consistent with the rather short metal–nitrogen bond lengths in the complexes. The pyrazol-1-yl group acts as moderately strong σ donor and a weak out-of-plane π donor, with the π interaction in the plane of the amine ligand probably being close to zero [61].
The main applications of C-homoscorpionate complexes as catalysts for oxidative reactions, where the involvement of metal redox processes is crucial for the catalytic activity, are the important and challenging single-pot oxidation of gaseous (e.g., direct oxidation of methane to carboxylic acids, Scheme 1) [47,62]) and liquid (e.g., cyclohexane to the correspondent alcohol-ketone mixture, Scheme 2) alkanes [7,8,13,43,46,47,48,49,53,57,58,63,64,65,66,67,68]. Indeed, oxidation of alkanes has been the object of considerable research [69,70,71,72,73,74,75,76,77,78], but still constitutes a serious challenge to modern chemistry, owing to the high inertness of these substrates. Currently, alkanes are mainly applied as fuels but it would be desirable to direct their application to the synthesis of organic products of a high added value. The feasibility of this approach is supported by the industrial application of cyclohexane in the production of cyclohexanone and cyclohexanol (KA oil, Scheme 2), with dioxygen as oxidizing agent and catalysts based on cobalt [79]. However, this industrial process has a very low yield to ensure an acceptable selectivity. Another case is the industrial production of acetic acid, a known commodity of large-scale demand. Currently mainly obtained by the improved CativaTM methanol carbonylation process [69,71,80], it nevertheless requires three steps from natural gas and considerably harsh, pollutant rich and costly conditions. The above examples explain the interest in finding more efficient processes and in understanding the involved mechanisms [81,82,83].
Among the new catalysts recently found for the above reactions are C-homoscorpionate metal complexes which have been successfully applied either as catalysts for oxygenations, with H2O2, to produce the respective alcohols and ketones or, with K2S2O8, to directly yield carboxylic acids [7,8].
The reactions leading to the above oxygenated species are believed to proceed mainly via both carbon- and oxygen-centered radicals. Interestingly, although occurring via the formation of reactive radicals, such reactions are rather selective [7,8,16]. The C-scorpionate catalyst initially activates not the alkane but reacts with another reactant, usually the oxidant (e.g., hydrogen-peroxide) [7,8]. The formed reactive species (e.g., hydroxyl radical) attacks the alkane molecule without any participation in the latter process of the metal complex. Thus, the metal catalyst does not take part in the direct “activation” of the carbon-hydrogen bond by the radical.
A possible mechanism [69,70] for the generation of carboxylic acids is represented in Scheme 3 for an oxo-V complex: a potassium peroxodisulfate salt (K2S2O8) is essential for the formation of alkyl radicals whereas the scorpionate catalyst is not needed for this purpose. Nevertheless, no carboxylic acid is detected in the absence of the catalyst. Under the experimental conditions, peroxodisulfate undergoes thermolysis into sulfate (or its protonated form HSO4 if in acidic medium) radicals which are known alkane hydrogen abstractors, leading to R. Further conversion of alkyl radical to carboxylic acid includes carbonylation of the former by CO to form the acyl radical RCO. The latter may then be converted, in the presence of the metal catalyst, by its oxygenation to give RCOO, involving a peroxo metal species (Scheme 3) derived from the reaction of catalyst with HS2O8 or with H2SO5 (peroxomonosulfuric acid) formed upon reaction of HS2O8 with TFA or hydrolysis by traces of water. Then RCOOC abstracts a hydrogen atom from, for example, excess TFA or alkane to afford the desired carboxylic acid.
In the case of the oxidation of alkanes with hydrogen peroxide, detailed investigation of the effects of various experimental parameters in this reaction, the use of radical traps, kinetic and selectivity studies complemented with theoretical calculations [73,84,85,86] indicated the interest of using controlled amounts of water and acid, and assured the involvement of hydroxyl (OH) and alkyl (R) radicals in a radical type mechanism (see Equations (2)–(10) and Scheme 4 for a more detailed formation of hydroxyl and hydroperoxyl radicals).
The proposed route for the metal-catalyzed decomposition of hydrogen peroxide (Haber–Weiss mechanism) [70,87] includes the following two key stages (Equations (2) and (3)) of formation of the oxygen-centered radicals HOO and HO:
Mn+ + H2O2 → HO + M(n+1)+ + HO
M(n+1)+ + H2O2 → HOO + H+ + Mn+
It is the hydroxyl radical (derived from the reduction of H2O2 by the reduced form of the metal catalyst, Equation (2) and Scheme 4 for a C-homoscorpionate oxo-V complex) that reacts with the alkane generating the alkyl radical R (Equation (4)) which, in turn, reacts with dioxygen (Equation (5)) to form the alkylperoxyl radical ROO. The latter gives rise to the alkyl-hydroperoxide (ROOH) (Equation (6)) which, in the presence of both the reduced and oxidized forms of the metal catalyst, decomposes (Equations (7)–(10)) to the ketone and/or the alcohol.
HO + RH → H2O + R
R+ O2 → ROO
ROO + H2O2 → ROOH + HOO
ROOH + Mn+ → RO + HO + M(n+1)+
ROOH + M(n+1)+ → ROO + H+ + Mn+
RO + RH → ROH + R
2ROO → ROH + R-H = O + O2
The use of C-homoscorpionate complexes to catalyze epoxidation of alkenes, a very useful synthetic transformation to produce fine chemicals [79,80], is also significant. Selective catalytic epoxidation of cis-cyclooctene to 1,2-epoxy-cyclooctane was achieved in the presence of tris(pyrazol-1-yl)methane Mo(VI) complexes [19,20,88], in particular if water is rigorously excluded from the reaction mixture. Other alkene substrates such as R-(+)-limonene, 1-octene, trans-2-octene, cyclododecene, 3-carene, and 4-vinyl-1-cyclohexene are also selectively converted into the corresponding epoxides.
In the presence of the sacrificial oxidant PhI(OAc)2, aqua Ru(II) tris(pyrazol-1-yl)methane compounds catalyze the aerobic epoxidation of a wide variety of alkenes [89]. The mechanistic pathway for the epoxidation (Scheme 5) proceeds via the formation of an active metal–oxo intermediate through the mediation of iodobenzene diacetate. The active catalytic species is a formally Ru(IV)=O one, resulting from the oxidation of the aqua Ru(II) complex by PhI(OAc)2. The electrophilic metal bound oxo group subsequently interacts with the incoming olefinic double bond with transfer of the oxo group. The involvement of a concerted transition state for the transfer of the oxygen atom from the metal-oxido complex to the olefinic double bond is suggested [89,90].
From the above, it turns out that an effective catalyst for the above oxidation reactions requires the ability to undergo reversible redox processes involving electron transfer at accessible potentials. This redox potential—oxidative catalytic activity relationship will be addressed in detail in the following section.

3. Electrochemical Properties of C-Scorpionate Metal Complexes

The electrochemical approach is a very powerful tool for fundamental chemical characterization of species that can be oxidized or reduced. By continuously changing the working potential, its cycling or keeping constant, enables not only the determination of the respective oxidation or reduction potentials but also revelation of the reversibility of the redox processes, the nature, kinetics and equilibrium constants of the follow-up reactions, the stability and structure of intermediates, the type and yield of products, etc. In fact, an electron transfer in a coordination compound can induce very diverse chemical reactivity, ultimately with catalytic significance.
Some C-homoscorpionate complexes underwent systematic electrochemical investigation usually by cyclic voltammetry (CV) and controlled potential electrolysis (CPE) techniques, at platinum working electrodes (disk or gauze, respectively). Glassy carbon working electrodes for CV were also used [89]. Experiments were performed in a three-electrode system whose potential was controlled vs. a Luggin capillary connected to a silver wire pseudo-reference electrode and a Pt auxiliary electrode. The complexes were added to a 0.1–0.2 M [nBu4N][X] (X = BF4, PF6 or ClO4) or [Et4N][ClO4]/aprotic non-aqueous medium (e.g., CH2Cl2, NCMe, DMF or DMSO), at room temperature, under dinitrogen [46,49,50,51,52,53,54,89,91,92,93].
Their measured redox potentials in volts vs. saturated calomel electrode (V vs. SCE) and the eventual reversibility of the redox process, are indicated in Table 1.
All authors found that C-homoscorpionate ligands are electrochemically inert in the potential range of −2.0 V to 2.0 V vs. SCE, at the used experimental conditions [46,49,50,51,52,53,54,89,91,92,93], thus no ligand centered oxidation or reduction has been reported to date.
Most of the metallic compounds bearing tris(pyrazol-1-yl)methane ligands exhibit at least a single-electron (determined by exhaustive CPE) oxidation wave, assigned to the dndn−1 metal oxidation. Exceptions are, as expected, V(V), Re(VII), Ni(II), Cu(II), Au(III), and Zn(II) complexes. The said oxidation waves can meet the reversibility criteria [94] or be irreversible due to chemical reactions that follow the electron-transfer process (Table 1). Most of the C-homoscorpionate complexes also exhibit (Table 1) a reduction wave which usually is followed, at a lower potential, by a second one. These waves often (e.g., for V, Re, Fe, Ru or Co complexes) correspond to single-electron processes, being assigned to the dndn+1 and dn+1dn+2 metal reductions.
The highest known first oxidation potential of all C-scorpionate metal complexes is shown by the 15-electron Re(IV) complex [ReCl42-HC(pz)3}] (IEpox = 1.79 V vs. SCE, Table 1) per its electron deficiency. Such oxidation potential value is even higher than the one of the oxo-Re(V) 16-electron complex [ReOCl{κ3-SO3C(pz)3}(PPh3)]Cl (IEpox = 1.45 vs. SCE, Table 1) in spite of the higher metal oxidation state of the latter. The presence of the strong electron-donor oxo-ligand provides another reason for the lower oxidation potentials of this oxo-complex. [ReCl42-HC(pz)3}] is also the one that exhibits the most favorable (highest) reduction potential (IE1/2ox = −0.06 V vs. SCE, Table 1) in accord with its low electron-count. Harder to reduce are the oxo-Re species [ReO33-SO3C(pz)3}] and [ReOCl{κ3-SO3C(pz)3}(PPh3)]Cl, in agreement with the presence of the strong electron-donor oxide ligand and with their higher electron count. Among the rhenium complexes, Re(III) 16-electron complexes [ReCl33-HC(pz)3}] and [ReCl33-HC(3,5-Me2pz)3}] are those that present the lowest oxidation potential (IEpox = 1.14 and IE1/2ox = 1.25 V vs. SCE, Table 1), consistent with the lower metal oxidation state. In contrast with the measured values, complex [ReCl33-HC(pz)3}] would be expected to have a higher oxidation potential than the analogous [ReCl33-HC(3,5-Me2pz)3}], on account of the weaker electron-donor character of HC(pz)3 in the former in comparison with HC(3,5-Me2pz)3 in the latter. However, the irreversible character of the oxidation wave of the former (indicative of a chemical reaction following the electron-transfer step, with a resulting shift of the oxidation potential) preclude a reliable comparison between the measured potentials for these complexes.
In the case of Mo(0 or II) complexes a second single-electron oxidation process is detected (not shown in Table 1) in the potential range of 0.18 to 0.6 V vs. SCE. In Li[Mo{κ3-SO3C(pz)3}(CO)3] yields the 16-electron Mo(II) complex [Mo{κ3-SO3C(pz)3}(CO)3]+, its irreversibility being associated to fast coordination of a solvent molecule, leading to an electronically saturated product.
The irreversibility of the first oxidation wave of compounds [MoI{κ3-SO3C(pz)3}(CO)3] and [MoH{κ3-SO3C(pz)3}(CO)3] signals the instability of the resulting cationic Mo(III) complexes, which then rapidly decompose with probable CO loss [95] and, for the hydride compound, by deprotonation [96,97,98]. The first oxidation potentials of all these tricarbonyl complexes are much lower than that of the parent hexacarbonyl compound, on account of the replacement of three carbonyls in the latter by the more electron-donating C-scorpionate ligands [24,27,28]. Moreover, the first oxidation potential of [MoH{κ3-SO3C(pz)3}(CO)3] in comparison with [MoI{κ3-SO3C(pz)3}(CO)3] reflects the stronger electron-donor character of the hydride relatively to the iodide ligand [28].
These Mo (0 or II) compounds have not yet been used for catalytic oxidation reactions. Nevertheless, their low oxidation potentials (first oxidation wave in the range 0.09–0.44 V vs. SCE, Table 1) and the detected easy coordination/decoordination of substrates are promising features for a possible good oxidative catalytic performance.
The interest in electron transfer induced reactivity of C-scorpionate metal compounds is demonstrated in the following catalytic systems where such complexes provide unprecedented examples.

3.1. Oxidation of Alkanes to Alcohols and Ketones

In the case of the oxidation of alkanes with peroxides, the availability of reducible metal species, easily detectable by electrochemical experiments, was found very important for the catalytic performance of C-homoscorpionate complexes.
As previously mentioned, the formation of RO and ROO radicals (Equations (2) and (3), and Scheme 4) involves the reaction of both reduced and oxidized forms of the metal catalyst and is a key step for the occurrence of the C–H abstraction from the alkane. Therefore, C-homoscorpionate complexes that undergo redox processes at accessible potential values are expected to display better oxidative catalytic performance than those harder to oxidize or reduce.
In fact, V(V) complexes [VO23-SO3C(pz)3}] and [VO23-HC(pz)3}][BF4], whose accessible potential values for the first single-electron [V(V) to V(IV)] reduction process are −0.46 and −0.48 V vs. SCE, lead to quite similar (19% and 18.6%, respectively [47]) KA oil yields (among the highest values obtained for this class of catalysts) by catalytic oxidation of cyclohexane. The turnover number (TON, moles of product per mole of catalyst) values also follow the trend: 117 and 112, respectively, for [VO23-SO3C(pz)3}] and [VO23-HC(pz)3}][BF4].
A further example comes from V(III or IV) complexes. [VCl33-SO3C(pz)3}] is easier to oxidize than [VOCl23-CH3SO2OCH2C(pz)3}] (1.14 vs. 1.35 V, Table 1) and thus yields higher KA oil amounts (13% (TON = 121) [46] vs. 7% (TON = 89) in the presence of [VOCl23-CH3SO2OCH2C(pz)3}] [49]). Moreover, trichlorovanadium(III) [VCl33-HC(pz)3}] leads to higher yield (18%) and TON (167) values [47] than the related [VCl33-SO3C(pz)3}] (13% yield and a TON of 121, [46,47]), in accordance with its lower oxidation state and the neutral scorpionate ligand in[VCl33-HC(pz)3}].
Likewise, for the Co(II) complexes [CoCl2(H2O){κ3-PyCH2OCH2C(pz)3}] and [CoCl2(H2O){κ3-CH3SO2OCH2C(pz)3}]; the latter presents lower oxidation potential (1.10 V vs. SCE, Table 1) and thus exhibits better catalytic performance: 10.5% vs. 3.2% yield of KA oil in the presence of [CoCl2(H2O){κ3-PyCH2OCH2C(pz)3}] [12] which is oxidized at 1.28 V vs. SCE (Table 1).
A similar behavior is found for the chloro-Au(III) complexes: [AuCl22-HC(pz)3}]Cl and [AuCl22-HOCH2C(pz)3}]Cl which present very close reduction potentials for the first irreversible two electrons Au(III) → Au(I) reduction process (Table 1) and are the most active, leading to 8.1% and 10.3%, respectively, of KA oil [57]. The hardest to reduce (−0.11 V vs. SCE, Table 1) yields only 7.5% of the oxygenated mixture under the same conditions [64]. The lower reduction potential of [AuCl22-HC(3,5-Me2pz)3}]Cl in comparison with the one of [AuCl22-HC(pz)3}]Cl is consistent with the stronger electron-donor ability of the methyl-substituted κ2-HC(3,5-Me2pz)3 ligand than that of κ2-HC(pz)3 [24]. However, an accurate comparison cannot be established due to the irreversibility of the reduction waves (the reduction potential is not the thermodynamic one). Moreover, whereas the CH2OH substituent at the apical methine carbon appears to have limited influence on the redox potential of the gold complexes (Table 1), the replacement of hydrogens by an electron donor group (Me) at the pyrazolyl rings of the C-scorpionate leads to an electronically richer Au(III) center, resulting in a measurable (ca. 0.1 V) cathodic shift of the potential. A second irreversible reduction (Table 1) assigned to the Au(I) → Au(0) reduction leads to the appearance of gold metal at the platinum electrodes surface after exhaustive controlled potential electrolysis and an irreversible anodic wave (in the range 0.44–0.50 V vs. SCE) observed upon scan reversal after the second reduction wave, corresponding to the oxidation of the Au(0) species formed in the second reduction process.

3.2. Oxidation of Alkanes to Carboxylic Acids

The catalytic activity of the only to date tested [51] Re(III) complexes, [ReCl33-HC(pz)3}] and [ReCl23-HC(pz)3}(PPh3)][BF4], for the for the direct oxidation of ethane to acetic acid follows their oxidation behavior (Table 1 and Figure 6): [ReCl23-HC(pz)3}(PPh3)][BF4] presents considerably lower oxidation potential (E1/2ox = 0.54 V vs. SCE) and leads to higher acetic acid yield (16%) and TON (8) values than [ReCl33-HC(pz)3}] (E1/2ox = 1.14 V vs. SCE; 5% acetic acid yield and TON = 2) [51,52].

3.3. Baeyer-Villiger Oxidation of Ketones

The regioselective Baeyer-Villiger (BV) oxidation of 2-methylhexanone to 6-methylhexanolide (as a result of the formal insertion of the oxygen atom between the carbonyl and the more substituted Cα atom) was found [21] to be favored by strong Lewis acid Re catalysts (Table 1 and Figure 6). The non-radical mechanism of BV ketone oxidation proceeds via the activation of the ketone, upon coordination to the metal catalyst, to nucleophilic attack of the peroxide oxidant, followed by heterolytic peroxo-bond cleavage and carbanion migration. In fact, for the same Re oxidation state (III ( ) or VII ( ), Figure 6) higher lactone yields are obtained when the catalyst presents a higher (more positive) reduction potential. The electron deficiency of the catalyst (stronger Lewis acid character) activates to a greater extent the carbonyl group of the ketone for the nucleophilic attack by hydrogen peroxide.

3.4. Oxidation of 1,2-Diols

The Fe(III)/Fe(II) redox potentials of the chloro-iron(III) complexes follow the trend [FeCl33-HC(3-iPrpz)3}] (3-iPr = iso-propyl group on the 3-position of pyrazole rings) > [FeCl33-HC(pz)3}] > [FeCl33-HC(3,5-Me2pz)3}] (Table 1), which represents a decrease in Lewis acidity of the iron(III) center along this series. In [FeCl33-HC(3-iPrpz)3}], the sterically hindering iso-propyl group weakens the coordination of nitrogen of pyrazol-1-yl ring conferring an enhanced Lewis acidity of the iron(III) center. The electron-releasing methyl groups on the pyrazol-1-yl ring in [FeCl33-HC(3,5-Me2pz)3}] increase the electron density on pyrazol-1-yl nitrogen and hence decreases the Lewis acidity of the iron(III) center [92].
The catechol dioxygenase activity of the above iron(III) complexes was tested and the electrochemical properties of the catecholate adducts of the complexes reveal that a systematic variation in the ligand donor atom type significantly influences the Lewis acidity of the iron(III) center and hence the interaction of the complexes with simple and substituted catechols.
The rate of oxygenation increases upon increasing the Lewis acidity of the iron(III) center by modifying the ligand substituents. One of the pyrazolyl arms in the catecholate adducts is sterically constrained by the 6,6,6-chelate ring system and appears to dissociate from the coordination sphere upon binding to the catecholate substrate, which is followed by dioxygen attack at the equatorial plane leading to the formation of benzoquinone [92].

3.5. Carboxylation of Alkanes

The catalytic activity of C-homoscorpionate V(V) complexes for the one-pot carboxylation of methane to acetic acid (Scheme 1) was found [47] to be in accordance with the order of their V(V) → V(IV) reduction potentials, which follows the electron-donor characters of the scorpionate ligands and the charge of the complex (see Table 1 and Figure 7). The stronger vanadium(V) Lewis acid (the easiest to reduce) favors the carboxylation mechanism represented in Scheme 3 and allows the highest product yield (under the same experimental conditions of the other two V(V) complexes) to be achieved. Moreover, turnover number values follow the yields trend [47].

3.6. Epoxidation of Alkenes

For the o-benzoquinonediimine (bqdi) Ru(II) complexes [RuCl{κ3-HC(pz)3}(bqdi)][ClO4] and [Ru(H2O){κ3-HC(pz)3}(bqdi)][ClO4]2, electrochemical and DFT calculations established [89] that the redox non-innocent bqdi was stabilized in its fully oxidized quinone state in both the chloro complex ([Ru{κ3-HC(pz)3}(bqdi)(Cl)]+) and the aqua ([Ru){κ3-HC(pz)3}(bqdi)(H2O)]2+) derivative. The chloro complex exhibits metal based Ru(II)/(III) oxidation and bqdi centered reduction.
The aqua complex [Ru(H2O){κ3-HC(pz)3}(bqdi)][ClO4]2 exhibits two one electron oxidations at pH 7, suggesting the formation of a {Ru(IV)=O} species, the supposed active species of the alkene epoxidation catalytic cycle (see Scheme 5). Thus, [Ru(H2O){κ3-HC(pz)3}(bqdi)]2+ functions as an efficient pre-catalyst for the selective epoxidation of a wide variety of alkenes in the presence of iodobenzene diacaetate as the sacrificial oxidant.

3.7. Redox Potential Parametrization

The values of the Ru(II/III) oxidation potential (in the range of 0.95–1.37 V vs. SCE, Table 1) of [Ru(L)(L′)]X complexes [L = p-cymene, benzene, hexamethylbenzene (HMB), or cyclooctadiene (cod), L′ = tris(pyrazol-1-yl)methanesulfonate or the 3-phenylpyrazolyl-substituted derivative, X = Cl or BF4] reflect [54] the electron-donor characters of their ligands: for the cationic complexes, with the common [Ru{κ3-SO3C(pz)3}]+ center, the order of the oxidation potentials follows that (in the opposite direction) of the electron-releasing character of the corresponding variable ligand (cymene > benzene) as measured by the electrochemical Lever EL ligand parameter (+1.48 and +1.59 V vs. NHE for cymene and benzene, respectively) [54].
As mentioned in the introduction, EL is a measure of the electron-donor character of the ligand, the stronger this character, the lower is EL. Moreover, the experimental oxidation potentials are in accordance with those predicted from the knowledge of EL values for cymene [54], benzene [54] and κ3-SO3C(pz)3 (Table 2) [52] by applying the Lever method. Accordingly, the higher oxidation potentials of [Ru{κ3-SO3C(3-Phpz)3}(benzene)]Cl or [RuCl{κ3-SO3C(3-Phpz)3}(cod)], bearing the 3-phenyl substituted tris(pyrazol-1-yl)methanesulfonate ligand, than those of the analogous [Ru(benzene){κ3-SO3C(pz)3}]Cl or [RuCl(cod){κ3-SO3C(pz)3}] reflect the expected weaker electron-donor character of {κ3-SO3C(3-Phpz)3} ligand in comparison with that of {κ3-SO3C(pz)3}. Hence, the former ligand should present a higher EL value than the latter [52] (Table 2).
The electrochemical EL Lever parameters for tris(pyrazol-1-yl)methane ligands [23,51,52,54] to date, found possible to estimate from the oxidation potential values of C-homoscorpionate complexes, by applying the linear (valid for octahedral complexes) relationship (1) and considering its extension to square-planar coordination and to full- and half-sandwich complexes [28,99,100,101,102,103], are presented in Table 2. These values correspond to partial EL parameters assigned to each metal ligated arm (2-electron-donor) of the scorpionate ligand. Thus, the overall EL value of a scorpionate ligand will depend on its coordination mode to the metal center in a complex.
Since the EL parameter is a measure of the electron donor character of a ligand (the lower the parameter value, the stronger is that character), each ligated pyrazol-1-yl arm in {SO3C(pz)3} (EL = −0.09 V vs. SHE, Table 2) is clearly a stronger electron donor than in HC(pz)3 (EL = 0.14 V vs. SHE), indicating a much stronger electron-releasing ability of the anionic CSO3 group at {SO3C(pz)3} than the methine HC group in the neutral HC(pz)3. That is consistent with the above reported electrochemical behavior for V, Re, and Au complexes. Moreover, the value of –0.05 V vs. SHE (for each coordinating pyrazolyl arm) agrees with the expected slightly weaker electron-donor character of SO3C(3-Phpz)3 relative to SO3C(pz)3 due to the phenyl substituent at the pyrazol-1-yl rings in the former ligand.
The above two series Ru(II) complexes, bearing the tris(pyrazol-1-yl)methanesulfonate ligand and its 3-phenyl substituted derivative, have not yet been tested as catalysts for partial oxidation reactions. Nevertheless, based on the reported electrochemical studies [54] we would expect a better oxidative catalytic performance for the tris(pyrazol-1-yl)methanesulfonate complexes [Ru(p-cymene){κ3-SO3C(pz)3}]Cl and [Ru(cod)Cl{κ3-SO3C(pz)3}].
A comparison of the effect of HC(pz)3 or HB(pz)3 ligands on the redox potential of a metal complex was reported for acetonitrile-Ru(II) complexes [93]. [Ru{κ3-HB(3,5-Me2pz)3}(NCCH3)3][OTf] exhibits a higher Ru(III)/(II) potential than its carbon analogue [Ru{κ3-HC(3,5-Me2pz)3}(NCCH3)3][BF4]2 (0.59 and 0.42 V vs. SCE, respectively), indicating that the charged {κ3-HB(3,5-Me2pz)3} ligand stabilizes Ru(III) relative to Ru(II) compared to the neutral {κ3-HC(3,5-Me2pz)3}. In contrast, the Ru(III)/(II) reduction potential observed for the [Ru{κ3-HB(3,5-Ph2pz)3}(NCCH3)3][BF4] complex (0.57 V vs. SCE) is lower than the same potential for the [Ru{κ3-HC(3,5-Ph2pz)3}(NCCH3)3][BF4]2 complex (0.71 V vs. SCE, Table 1), indicating that ligand charge is not as significant a factor as steric in determining the stability of ruthenium oxidation states for complexes with these bulky ligands. Since the redox potential of [Ru{κ3-HB(3,5-Ph2pz)3}(NCCH3)3][BF4] complex is smaller than that of the [Ru{κ3-HB(3,5-Me2pz)3}(NCCH3)3][OTf] complex, but the reverse is true for the [Ru{κ3-HC(3,5-Ph2pz)3}(NCCH3)3][BF4]2 and [Ru{κ3-HC(3,5-Me2pz)3}(NCCH3)3][BF4]2 complexes, both electronic and steric factors of these ligands affect the redox potentials of their Ru(II) complexes. Overall, Ru(II) coordination by negatively charged {κ3-HB(3,5-R2pz)3} (R = Me or Ph) or neutral {κ3-HC(3,5-R2pz)3} ligands with varying steric bulk alters the Ru(III)/(II) potential by over 400 mV. The ability to alter the stability of ruthenium +2 or +3 oxidation states may be used to tune catalytic reactions.
The electronic and structural properties of scorpionate ligands, such as poly(pyrazol-1-yl)methane ligands, play an important role in the ability of several transition metal complexes to mediate C–H activation and functionalization as well as other partial oxidations. Thus, the knowledge of the redox behavior of a certain C-homoscorpionate catalyst, as well as its relationship with the structure of the catalyst, may allow tailoring of its structural design to present a favorable value of potential to enhance its catalytic performance. Moreover, tris(pyrazol-1-yl)methane ligands may act as more than simple spectators during chemical reactions experienced by their metal complexes, and have an important influence on their reactivity by means of temporary changes of denticity.
Of course, other factors are involved in the catalytic activity exhibited by the metal complex. Importantly, the tripodal C-scorpionate ligand, bearing three pyrazol-1-yl moieties (via their N atoms) is found to assist proton-transfer steps (see Scheme 4) that are involved in key catalytic oxidation processes. Such factors should additionally be considered in the design of a catalyst with expected improved activity for the above oxidation reactions.

Acknowledgments

The author gratefully acknowledges all the co-authors cited in the joint publications. The work in this area has been partially supported by the Fundação para a Ciência e a Tecnologia (FCT), Portugal, and its projects PTDC/QEQ-ERQ/1648/2014 and UID/QUI/00100/2013.

Conflicts of Interest

The author declares no conflict of interest. The founding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

References

  1. Pettinari, C. Scorpionates II: Chelating Borate Ligands—Dedicated to Swiatoslaw Trofimenko; Imperial College Press, World Scientific Pub.: London, UK, 2008; ISBN 13-1-86094-876-3. [Google Scholar]
  2. Trofimenko, S. Scorpionates: The Coordination Chemistry of Polypyrazolylborates Ligands; Imperial College Press: London, UK, 1999; ISBN 1-86094-172-9. [Google Scholar]
  3. Pettinari, C.; Santini, C. Comprehensive Coordination Chemistry II: From Biology to Nanotechnology; McCleverty, J.A., Meyer, T.J., Eds.; Elsevier Pergamon (APS): Amsterdam, The Netherlands, 2003; Volume 1, pp. 159–172. [Google Scholar]
  4. Pettinari, C.; Pettinari, R. Metal derivatives of poly(pyrazolyl)alkanes. I. Tris(pyrazolyl)alkanes and related systems. Coord. Chem. Rev. 2005, 249, 525–543. [Google Scholar] [CrossRef]
  5. Bigmore, H.R.; Lawrence, S.C.; Mountford, P.; Tredget, C.S. Coordination, organometallic and related chemistry of tris(pyrazolyl)methane ligands. Dalton Trans. 2005, 635–651. [Google Scholar] [CrossRef] [PubMed]
  6. Otero, A.; Fernández-Baeza, J.; Lara-Sánchez, A.; Sánchez-Barba, L.F. Metal complexes with heteroscorpionate ligands based on the bis(pyrazol-1-yl)methane moiety: Catalytic chemistry. Coord. Chem. Rev. 2013, 257, 1806–1868. [Google Scholar] [CrossRef]
  7. Martins, L.M.D.R.S.; Pombeiro, A.J.L. Tris(pyrazol-1yl)methane metal complexes for catalytic mild oxidative functionalizations of alkanes, alkenes and ketones. Coord. Chem. Rev. 2014, 265, 74–88. [Google Scholar] [CrossRef]
  8. Martins, L.M.D.R.S.; Pombeiro, A.J.L. Water-soluble C-scorpionate complexes: Catalytic and biological applications. Eur. J. Inorg. Chem. 2016, 2016, 2236–2252. [Google Scholar] [CrossRef]
  9. Reger, D.L.; Grattan, T.C.; Brown, K.J.; Little, C.A.; Lamba, J.J.S.; Rheingold, A.L.; Sommer, R.D. Syntheses of tris(pyrazolyl)methane ligands and {[tris(pyrazolyl)methane]Mn(CO)3}SO3CF3 complexes: Comparison of ligand donor properties. J. Organomet. Chem. 2000, 607, 120–128. [Google Scholar] [CrossRef]
  10. Pombeiro, A.J.L.; Martins, L.M.D.R.S.; Alegria, E.C.B.A. Use of Microwaves for the Synthesis of Substituted Tris(pyrazolyl)methanes. Patent 103,681, 4 December 2009. [Google Scholar]
  11. Martins, L.M.D.R.S.; Alegria, E.C.B.A.; Pombeiro, A.J.L. Ligands: Synthesis, Characterization and Role in Biotechnology; Smolenski, P., Gawryszewska, P., Eds.; Nova Science Publishers, Inc.: New York, NY, USA, 2014; Chapter 4; pp. 117–140. [Google Scholar]
  12. Silva, T.F.S.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Fernandes, A.R.; Silva, A.; Borralho, P.M.; Santos, S.; Rodrigues, C.M.P.; Pombeiro, A.J.L. Cobalt complexes bearing scorpionate ligands: Synthesis, characterization, cytotoxicity and DNA cleavage. Dalton Trans. 2012, 41, 12888–12897. [Google Scholar] [CrossRef] [PubMed]
  13. Silva, T.F.S.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Kuznetsov, M.L.; Fernandes, A.R.; Silva, A.; Santos, S.; Pan, C.-J.; Lee, J.-F.; Hwang, B.-J.; et al. Cobalt complexes with pyrazole ligands as catalysts for the peroxidative oxidation of cyclohexane. XAS studies and biological applications. Chem. Asian J. 2014, 9, 1132–1143. [Google Scholar] [CrossRef] [PubMed]
  14. Pettinari, C.; Marchetti, F.; Lupidi, G.; Quassinti, L.; Bramucci, M.; Petrelli, D.; Vitali, L.A.; Guedes da Silva, M.F.C.; Martins, L.M.D.R.S.; Smoleński, P.; et al. Synthesis, Antimicrobial and antiproliferative activity of novel silver(I) tris(pyrazolyl)methanesulfonate and 1,3,5-triaza-7-phosphadamantane complexes. Inorg. Chem. 2011, 50, 11173–11183. [Google Scholar] [CrossRef] [PubMed]
  15. Niesel, J.; Pinto, A.; N’Dongo, H.W.P.; Merz, K.; Ott, I.; Gust, R.; Schatzschneider, U. Photoinduced CO release, cellular uptake and cytotoxicity of a tris(pyrazolyl)methane (tpm) manganese tricarbonyl complex. Chem. Commun. 2008, 1798–1800. [Google Scholar] [CrossRef] [PubMed]
  16. Martins, L.M.D.R.S.; Pombeiro, A.J.L. Carbon-scorpionate Complexes in Oxidation Catalysis. In Advances in Organometallic Chemistry and Catalysis, The Silver/Gold Jubilee ICOMC Celebratory Book; Pombeiro, A.J.L., Ed.; J. Wiley & Sons: Hoboken, NJ, USA, 2014; Chapter 22; pp. 285–294. [Google Scholar]
  17. Martins, L.M.D.R.S.; Ribeiro, A.P.C.; Carabineiro, S.A.C.; Figueiredo, J.L.; Pombeiro, A.J.L. Highly efficient and reusable CNT supported iron(II) catalyst for microwave assisted alcohol oxidation. Dalton Trans. 2016, 45, 6816–6819. [Google Scholar] [CrossRef] [PubMed]
  18. Pombeiro, A.J.L.; Martins, L.M.D.R.S.; Ribeiro, A.P.C.; Carabineiro, S.A.C.; Figueiredo, J.L. Production Process of Ketones from Secondary Alcohols. Patent 109,062, 29 December 2015. [Google Scholar]
  19. Santos, A.M.; Kuhn, F.E.; Bruus-Jensen, K.; Lucas, I.; Romão, C.C.; Herdtweck, E. Molybdenum(VI) cis-dioxo complexes bearing (poly)pyrazolylmethane and -borate ligands: Syntheses, characterization and catalytic applications. J. Chem. Soc. Dalton Trans. 2001, 1332–1337. [Google Scholar] [CrossRef]
  20. Neves, P.; Gago, S.; Balula, S.S.; Lopes, A.D.; Valente, A.A.; Cunha-Silva, L.; Paz, F.A.A.; Pillinger, M.; Rocha, J.; Silva, C.M.; et al. Synthesis and catalytic properties of molybdenum(VI) complexes with tris(3,5-dimethyl-1-pyrazolyl)methane. Inorg. Chem. 2011, 50, 3490–3500. [Google Scholar] [CrossRef] [PubMed]
  21. Martins, L.M.D.R.S.; Alegria, E.C.B.A.; Smoleński, P.; Kuznetsov, M.L.; Pombeiro, A.J.L. Oxorhenium complexes bearing the water-soluble tris(pyrazol-1-yl)methanesulfonate, 1,3,5-triaza-7-phosphaadamantane or related ligands, as catalysts for the Baeyer-Villiger oxidation of ketones. Inorg. Chem. 2013, 52, 4534–4546. [Google Scholar] [CrossRef] [PubMed]
  22. Martins, L.M.D.R.S.; Pombeiro, A.J.L. C-scorpionate rhenium complexes and their application as catalysts in Baeyer-Villiger oxidation of ketones. Inorg. Chim. Acta 2016. [Google Scholar] [CrossRef]
  23. Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Redox potential parameterization in coordination compounds with polydentate scorpionate and benzene ligands. Electrochim. Acta 2012, 82, 478–483. [Google Scholar] [CrossRef]
  24. Pombeiro, A.J.L. Characterization of coordination compounds by electrochemical parameters. Eur. J. Inorg. Chem. 2007, 2007, 1473–1482. [Google Scholar] [CrossRef]
  25. Guedes da Silva, M.F.C.; Martins, L.M.D.R.S.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L. Redox potential-structure relationships in 18-and 17-electron mononitrile (or monocarbonyl) diphosphine complexes of Re and Fe. Collect. Czech. Chem. Commun. 2001, 66, 139–154. [Google Scholar] [CrossRef]
  26. Bursten, B.E. Ligand additivity: Applications to the electrochemistry and photoelectron spectroscopy of D6 octahedral complexes. J. Am. Chem. Soc. 1982, 104, 1299–1304. [Google Scholar] [CrossRef]
  27. Lever, A.B.P. Electrochemical parametrization of metal complex redox potentials, using the ruthenium(III)/ruthenium(II) couple to generate a ligand electrochemical series. Inorg. Chem. 1990, 29, 1271–1285. [Google Scholar] [CrossRef]
  28. Lever, A.B.P.; Dodsworth, E.S. Inorganic Electronic Structure and Spectroscopy; Solomon, E.I., Lever, A.B.P., Eds.; Wiley: New York, NY, USA, 1999; Volume 2, p. 227. [Google Scholar]
  29. Trofimenko, S. Boron-Pyrazole Chemistry. J. Am. Chem. Soc. 1966, 88, 1842–1844. [Google Scholar] [CrossRef]
  30. Trofimenko, S. Boron-pyrazole chemistry. II. Poly(1-pyrazolyl)-borates. J. Am. Chem. Soc. 1967, 89, 3170–3177. [Google Scholar] [CrossRef]
  31. Hückel, W.; Bretschneider, H. N-tripyrazolyl-methane. Ber. Dtsch. Chem. Ges. 1937, 70, 2024–2026. [Google Scholar] [CrossRef]
  32. Jesson, J.P. Isotropic nuclear resonance shifts in some trigonal Co(II) and Ni(II) chelate systems. J. Chem. Phys. 1966, 45, 1049–1056. [Google Scholar] [CrossRef]
  33. Juliá, S.; del Mazo, J.M.; Avila, L.; Elguero, J. Improved synthesis of polyazolylmethanes under solid-liquid phase-transfer catalysis. Org. Prep. Proced. Int. 1984, 16, 299–307. [Google Scholar] [CrossRef]
  34. Titze, C.; Hermann, J.; Vahrenkamp, H. Highly substituted tris(pyrazolyl)methane ligands and some zinc-complexes thereof. Chem. Ber. 1995, 128, 1095–1103. [Google Scholar] [CrossRef]
  35. Reger, D.L.; Collins, J.E.; Rheingold, A.L.; Liable-Sands, L.M. Synthesis and characterization of cationic [tris(pyrazolyl)methane]copper(I) carbonyl and acetonitrile complexes. Organometallics 1996, 15, 2029–2032. [Google Scholar] [CrossRef]
  36. Trofimenko, S. Recent advances in poly(pyrazolyl)borate (scorpionate) chemistry. Chem. Rev. 1993, 93, 943–980. [Google Scholar] [CrossRef]
  37. Otero, A.; Fernández-Baeza, J.; Antiñolo, A.; Tejeda, J.; Lara-Sánchez, A. Heteroscorpionate ligands based on bis(pyrazol-1-yl)methane: Design and coordination chemistry. Dalton Trans. 2004, 1499–1510. [Google Scholar] [CrossRef] [PubMed]
  38. Brunker, T.J.; Cowley, A.R.; O’Hare, D. Synthesis, structures, and redox properties of mixed-sandwich complexes of cyclopentadienyl and hydrotris(pyrazolyl)borate ligands with first-row transition metals. Organometallics 2002, 21, 3123–3138. [Google Scholar] [CrossRef]
  39. McKeown, B.A.; Lee, J.P.; Mei, J.; Cundari, T.R.; Gunnoe, T.B. Transition metal mediated C–H activation and functionalization: The role of poly(pyrazolyl)borate and poly(pyrazolyl)alkane ligands. Eur. J. Inorg. Chem. 2016, 2016, 2296–2311. [Google Scholar] [CrossRef]
  40. Tellers, D.M.; Skoog, S.J.; Bergman, R.G.; Gunnoe, T.B.; Harman, W.D. Comparison of the relative electron-donating abilities of hydridotris(pyrazolyl)borate and cyclopentadienyl ligands: Different interactions with different transition metals. Organometallics 2000, 19, 2428–2432. [Google Scholar] [CrossRef]
  41. Jones, W.D.; Feher, F.J. Comparative reactivities of hydrocarbon C–H bonds with a transition-metal complex. Acc. Chem. Res. 1989, 22, 91–100. [Google Scholar] [CrossRef]
  42. Vahrenkamp, H. Transitions, transition states, transition state analogues: Zinc pyrazolylborate chemistry related to zinc enzymes. Acc. Chem. Res. 1999, 32, 589–596. [Google Scholar] [CrossRef]
  43. Silva, T.F.S.; Mishra, G.S.; Silva, M.F.G.; Wanke, R.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Cu(II) complexes bearing the 2,2,2-tris(1-pyrazolyl)ethanol or 2,2,2-tris(1-pyrazolyl)ethyl methanesulfonate scorpionates. X-ray structural characterization and application in the mild catalytic peroxidative oxidation of cyclohexane. Dalton Trans. 2009, 42, 9207–9215. [Google Scholar] [CrossRef] [PubMed]
  44. Wanke, R.; Silva, M.F.C.G.; Lancianesi, S.; Silva, T.F.S.; Martins, L.M.D.R.S.; Pettinari, C.; Pombeiro, A.J.L. Synthesis and coordination chemistry of a new N4-Polydentate class of pyridyl-functionalized scorpionate ligands: Complexes of FeII, ZnII, NiII, VIV, PdII and use for heterobimetallic systems. Inorg. Chem. 2010, 49, 7941–7952. [Google Scholar] [CrossRef] [PubMed]
  45. Wanke, R.; Smoleński, P.; Guedes da Silva, M.F.C.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Cu(I) complexes bearing the new sterically hindered and coordination flexible tris(3-phenyl-1-pyrazolyl)methanesulfonate (TpmsPh) ligand and the water-soluble phosphine 1,3,5-triaza-7-phosphaadamantane (PTA) or related ligands. Inorg. Chem. 2008, 47, 10158–10168. [Google Scholar] [CrossRef] [PubMed]
  46. Silva, T.F.S.; Alegria, E.C.B.A.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Scorpionate vanadium, iron and copper complexes as selective catalysts for the peroxidative oxidation of cyclohexane under mild conditions. Adv. Synth. Catal. 2008, 350, 706–716. [Google Scholar] [CrossRef]
  47. Silva, T.F.S.; Luzyanin, K.V.; Kirilova, M.V.; Silva, M.F.C.G.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Novel scorpionate and pyrazole dioxovanadium complexes, catalysts for carboxylation and peroxidative oxidation of alkanes. Adv. Synth. Catal. 2010, 352, 171–187. [Google Scholar] [CrossRef]
  48. Silva, T.F.S.; Mac Leod, T.C.O.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.; Pombeiro, A.J.L. Pyrazole or tris(pyrazolyl)ethanol oxo-vanadium(IV) complexes as homogeneous or supported catalysts for oxidation of cyclohexane under mild conditions. J. Mol. Catal. A Chem. 2013, 367, 52–60. [Google Scholar] [CrossRef]
  49. Silva, T.F.S.; Rocha, B.G.M.; Guedes da Silva, M.F.C.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. VIV, FeII, NiII and CuII complexes bearing 2,2,2-tris(pyrazol-1-yl)ethyl methanesulfonate: Application as catalysts for the cyclooctane oxidation. New J. Chem. 2016, 40, 528–537. [Google Scholar] [CrossRef]
  50. Dinoi, C.; Silva, M.F.C.G.; Alegria, E.C.B.A.; Smoleński, P.; Martins, L.M.D.R.S.; Poli, R.; Pombeiro, A.J.L. Molybdenum complexes bearing the tris(1-pyrazolyl)methanesulfonate ligand: Synthesis, characterization and electrochemical behavior. Eur. J. Inorg. Chem. 2010, 16, 2415–2424. [Google Scholar] [CrossRef]
  51. Alegria, E.C.B.A.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Syntheses and properties of Re(III) complexes derived from hydrotris(1-pyrazolyl)methanes. Molecular structure of [ReCl2(HCpz3)(PPh3)][BF4]. J. Organomet. Chem. 2005, 690, 1947–1958. [Google Scholar] [CrossRef]
  52. Alegria, E.C.B.A.; Martins, L.M.D.R.S.; Haukka, M.; Pombeiro, A.J.L. Rhenium complexes of tris(pyrazolyl)methanes and sulfonate derivative. Dalton Trans. 2006, 4954–4961. [Google Scholar] [CrossRef] [PubMed]
  53. Silva, T.F.S.; Silva, M.F.C.G.; Mishra, G.S.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Synthesis and structural characterization of iron complexes with 2,2,2-tris(1-pyrazolyl)ethanol ligands: Application in the peroxidative oxidation of cyclohexane under mild conditions. J. Organomet. Chem. 2011, 696, 1310–1318. [Google Scholar] [CrossRef]
  54. Marchetti, F.; Pettinari, C.; Pettinari, R.; Cerquetella, A.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.; Silva, T.F.S.; Pombeiro, A.J.L. Ru(II) arene complexes bearing tris(pyrazolyl)methanesulfonate capping ligands. Electrochemistry, spectroscopic and X-ray structural characterization. Organometallics 2011, 30, 6180–6188. [Google Scholar] [CrossRef]
  55. Rocha, B.G.M.; Mac Leod, T.C.O.; Guedes da Silva, M.F.C.; Luzyanin, K.V.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. NiII, CuII and ZnII complexes with a sterically hindered scorpionate ligand (TpmsPh) and catalytic application in the diasteroselective nitroaldol (Henry) reaction. Dalton Trans. 2014, 43, 15192–15200. [Google Scholar] [CrossRef] [PubMed]
  56. Rocha, B.G.M.; Wanke, R.; Guedes da Silva, M.F.C.; Luzyanin, K.V.; Martins, L.M.D.R.S.; Smolénski, P.; Pombeiro, A.J.L. Reactivity of bulky tris(phenylpyrazolyl)methanesulfonate copper(I) complexes towards small unsaturated molecules. J. Organomet. Chem. 2012, 714, 47–52. [Google Scholar] [CrossRef]
  57. Peixoto de Almeida, M.; Martins, L.M.D.R.S.; Carabineiro, S.A.C.; Lauterbach, T.; Rominger, F.; Hashmi, A.S.K.; Pombeiro, A.J.L.; Figueiredo, J.L. Homogeneous and heterogenised new gold C-scorpionate complexes as catalysts for cyclohexane oxidation. Catal. Sci. Technol. 2013, 3, 3056–3069. [Google Scholar] [CrossRef]
  58. Pombeiro, A.J.L.; Martins, L.M.D.R.S.; Ribeiro, A.P. Process for the Microwave-Assisted Conversion of Cycloalkanes to the Corresponding Alcohol-Ketone Mixtures, with Hydrogen Peroxide, and Using a Scorpionate Chloro-Complex of Iron(II) as Catalyst. Patent 107,797, 25 July 2014. [Google Scholar]
  59. Vitze, H.; Bolte, M.; Lerner, H.-W.; Wagner, M. Third-generation scorpionates [RBpz3]—How influential is the nondonor substituent R? Eur. J. Inorg. Chem. 2016, 2016, 2443–2454. [Google Scholar] [CrossRef]
  60. Trofimenko, S. Geminal poly(1-pyrazolyl)alkanes and their coordination chemistry. J. Am. Chem. Soc. 1970, 92, 5118–5126. [Google Scholar] [CrossRef]
  61. Astley, T.; Gulbis, J.M.; Hitchman, M.A.; Tiekink, E.R.T. Syntheses and characterisation of tris(3-(pyridin-2-yl)-1H-pyrazol-1-yl)methane and its bis(µ-hydroxo) dicobalt(II) complex. J. Chem. Soc. Dalton Trans. 1993, 509–515. [Google Scholar] [CrossRef]
  62. Pombeiro, A.J.L.; Martins, L.M.D.R.S.; Silva, T.F.S.; Silva, M.F.C.G.; Luzyanin, K.V.; Kirillova, M.V. Oxocomplexes of Vanadium(IV-V) with Scorpionate or Pyrazole Ligands and Their Application as Catalysts for the Peroxidative Oxidation of Cycloalkanes and the Carboxylation of Gaseous Alkanes. Patent 104,887, 15 September 2011. [Google Scholar]
  63. Pombeiro, A.J.L.; Martins, L.M.D.R.S.; Silva, M.F.C.G.; Mishra, G.S.; Silva, T.F.S.; Wanke, R. Copper(II) Complexes with Hydrophilic C-Functionalized Scorpionate Ligands and Their Application as Catalysts for the Peroxidative Oxidation of Cyclohexane under Environmentally Tolerable Conditions, in Particular in Aqueous Medium. Patent 104713, 8 September 2010. [Google Scholar]
  64. Pombeiro, A.J.L.; Martins, L.M.D.R.S.; Silva, T.F.S.; Mishra, G.S. Process for conversion of cyclohexane to cyclohexanol and cyclohexanone using scorpionate chloro-complexes of vanadium(III or IV) as catalysts, with oxygen in the absence of solvents. Patent 104,447, 30 December 2009. [Google Scholar]
  65. Pombeiro, A.J.L.; Martins, L.M.D.R.S.; Alegria, E.C.B.A.; Mishra, G.S.; Fraústo da Silva, J.J.R. Complexes of Rhenium and Pyrazole Supported on Functionalized Silica as Catalysts for the Partial Oxidation of n-hexane and Cyclohexane with Dioxygen and under Environmentally Acceptable Conditions. Patent 104,197, 27 July 2009. [Google Scholar]
  66. Pombeiro, A.J.L.; Martins, L.M.D.R.S.; Alegria, E.C.B.A.; Silva, T.F.S. Scorpionate Chloro-Complexes of Iron and Vanadium and Their Application as Catalysts for the Partial Oxidation, under Mild and Environmentally Tolerable Conditions, of Cyclohexane to Cyclohexanol and Cyclohexanone. Patent 104153, 1 July 2009. [Google Scholar]
  67. Alegria, E.C.B.A.; Kirillova, M.V.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Pyrazole and tris(pyrazolyl)methane rhenium complexes as catalysts for ethane and cyclohexane oxidations. Appl. Catal. A Gen. 2007, 317, 43–52. [Google Scholar] [CrossRef]
  68. Pombeiro, A.J.L.; Martins, L.M.D.R.S.; Alegria, E.C.B.A.; Kirillova, M.V. New Complexes of Rhenium with Pyrazole or Tris(1-pyrazolyl)methanes and Their Application as Catalysts for the Partial Oxidation, under Mild Conditions, of Ethane to Acetic and Acetaldehyde and of Cyclohexane to Cyclohexanol and Cyclohexanone. Patent 103,735, 31 March 2008. [Google Scholar]
  69. Labinger, J.A. Catalysis by Metal Complexes; Pérez, P.J., Ed.; Springer Science + Business Media: Dordrecht, The Netherlands, 2012; Chapter 2. [Google Scholar]
  70. Shilov, A.E.; Shul’pin, G.B. Activation of Catalytic Reactions of Saturated Hydrocarbons in the Presence of Metal Complexes; Kluwer Academic Press: New York, NY, USA, 2000. [Google Scholar]
  71. Derouane, E.D.; Haber, J.; Lemos, F.; Ramôa Ribeiro, F.; Guinet, M. (Eds.) Catalytic Activation and Functionalisation of Light Alkanes; NATO ASI Series; Kluwer Academic Publ.: Dordrecht, The Netherlands, 1998; Volume 44.
  72. Shul’pin, G.B. New trends in oxidative functionalization of carbon–hydrogen bonds: A review. Catalysts 2016, 6, 50. [Google Scholar] [CrossRef]
  73. Shul’pin, G.B. C–H functionalization: Thoroughly tuning ligands at a metal ion, a chemist can greatly enhance catalyst’s activity and selectivity. Dalton Trans. 2013, 42, 12794–12818. [Google Scholar] [CrossRef] [PubMed]
  74. Shul’pin, G.B. Metal-catalyzed hydrocarbon oxidations. C. R. Chim. 2003, 6, 163–178. [Google Scholar] [CrossRef]
  75. Shul’pin, G.B. Hydrocarbon oxygenations with peroxides catalyzed by metal compounds. Mini-Rev. Org. Chem. 2009, 6, 95–104. [Google Scholar] [CrossRef]
  76. Kirillova, M.V.; Kuznetsov, M.L.; Reis, P.M.; Silva, J.A.L.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L. Direct and remarkably efficient conversion of methane into acetic acid catalyzed by amavadine and related vanadium complexes. A synthetic and a theoretical DFT mechanistic study. J. Am. Chem. Soc. 2007, 129, 10531–10545. [Google Scholar] [CrossRef] [PubMed]
  77. Kirillova, M.V.; Kuznetsov, M.L.; Silva, J.A.L.; Guedes da Silva, M.F.C.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L. Amavadin and other vanadium complexes as remarkably efficient catalysts for one-pot conversion of ethane to propionic and acetic acids. Chem. Eur. J. 2008, 14, 1828–1842. [Google Scholar] [CrossRef] [PubMed]
  78. Kirillov, A.M.; Kopylovich, M.N.; Kirillova, M.V.; Haukka, M.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Multinuclear copper triethanolamine complexes as selective catalysts for the peroxidative oxidation of alkanes under mild conditions. Angew. Chem. Int. Ed. 2005, 44, 4345–4349. [Google Scholar] [CrossRef] [PubMed]
  79. Kirk-Othmer Encyclopedia of Chemical Technology; Seidel, A.; Bickford, M. (Eds.) J. Wiley & Sons: New York, NY, USA, 2014.
  80. Weissermel, W.; Horpe, H.J. Industrial Organic Chemistry, 2nd ed.; VCH Press: Weinheim, Germany, 1993. [Google Scholar]
  81. Schuchardt, U.; Cardoso, D.; Sercheli, R.; Pereira, R.; da Cruz, R.S.; Guerreiro, M.C.; Mandelli, D.; Spinace, E.V.; Pires, E.L. Cyclohexane oxidation continues to be a challenge. Appl. Catal. A Gen. 2001, 211, 1–17. [Google Scholar] [CrossRef]
  82. Martins, L.M.D.R.S.; Peixoto de Almeida, M.; Carabineiro, S.A.C.; Figueiredo, J.L.; Pombeiro, A.J.L. Heterogenisation of a C-scorpionate Fe(II) complex in carbon materials for cyclohexane oxidation with hydrogen peroxide. ChemCatChem 2013, 5, 3847–3856. [Google Scholar] [CrossRef]
  83. Martins, L.M.D.R.S.; Martins, A.; Alegria, E.C.B.A.; Carvalho, A.P.; Pombeiro, A.J.L. Efficient cyclohexane oxidation with hydrogen peroxide catalyzed by a C-scorpionate iron(II) complex immobilized on desilicated MOR zeolite. Appl. Catal. A Gen. 2013, 464–465, 43–50. [Google Scholar] [CrossRef]
  84. Pombeiro, A.J.L. Vanadium-catalyzed alkane functionalization reactions under mild conditions. In Vanadium: The Versatile Metal; Kustin, K., Pessoa, J.C., Crans, D.C., Eds.; ACS Symposium Series, Nº 974; American Chemical Society; Oxford University Press: Oxford, UK, 2007; Chapter 4; p. 51. [Google Scholar]
  85. Kirillova, M.V.; Kuznetsov, M.L.; Romakh, V.B.; Shul’pina, L.S.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L.; Shul’pin, G.B. Mechanism of H2O2 oxidations catalyzed by vanadate anion or oxovanadium(V) triethanolaminate (vanadatrane) in combination with pyrazine-2-carboxylic acid (PCA): Kinetic and DFT studies. J. Catal. 2009, 267, 140–157. [Google Scholar] [CrossRef]
  86. Kirillov, A.M.; Kirillova, M.V.; Pombeiro, A.J.L. Multicopper complexes and coordination polymers for mild oxidative functionalization of alkanes. Coord. Chem. Rev. 2012, 256, 2741–2759. [Google Scholar] [CrossRef]
  87. Rachmilovich-Calis, S.; Masarwa, A.; Meyerstein, N.; Meyerstein, D.; van Eldik, R. New Mechanistic Aspects of the Fenton Reaction. Chem. Eur. J. 2009, 15, 8303–8309. [Google Scholar] [CrossRef] [PubMed]
  88. Gomes, A.C.; Neves, P.; Figueiredo, S.; Fernandes, J.A.; Valente, A.A.; Paz, F.A.A.; Pillinger, M.; Lopes, A.D.; Gonçalves, I.S. Tris(pyrazolyl)methane molybdenum tricarbonyl complexes as catalyst precursors for olefin epoxidation. J. Mol. Catal. A Chem. 2013, 370, 64–74. [Google Scholar] [CrossRef]
  89. Agarwala, H.; Ehret, F.; Chowdhury, A.D.; Maji, S.; Mobin, S.M.; Kaim, W.; Lahiri, G.K. Electronic structure and catalytic aspects of [Ru(tpm)(bqdi)(Cl/H2O)]n, tpm = tris(1-pyrazolyl)methane and bqdi = o-benzoquinonediimine. Dalton Trans. 2013, 42, 3721–3734. [Google Scholar] [CrossRef] [PubMed]
  90. Lane, B.S.; Burgess, K. Metal-catalyzed epoxidations of alkenes with hydrogen peroxide. Chem. Rev. 2003, 103, 2457–2474. [Google Scholar] [CrossRef] [PubMed]
  91. Silva, T.F.S.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Redox behaviour of a tris(pyrazolyl) methanesulfonate vanadium complex, a preliminary study. Port. Electrochim. Acta 2006, 24, 257–259. [Google Scholar] [CrossRef]
  92. Dhanalakshmi, T.; Sureshb, E.; Palaniandavar, M. Synthesis, structure, spectra and reactivity of iron(III) complexes of imidazole and pyrazole containing ligands as functional models for catechol dioxygenases. Dalton Trans. 2009, 8317–8328. [Google Scholar] [CrossRef] [PubMed]
  93. Underwood, C.C.; Stadelman, B.S.; Sleeper, M.L.; Brumaghim, J.L. Synthesis and electrochemical characterization of [Ru(NCCH3)6]2+, tris(acetonitrile) tris(pyrazolyl)borate, and tris(acetonitrile) tris(pyrazolyl)methane ruthenium(II) complexes. Inorg. Chim. Acta 2013, 405, 470–476. [Google Scholar] [CrossRef]
  94. Bard, A.J.; Faulkner, L.R. Electrochemical Methods–Fundamentals and Applications, 2nd ed.; John Wiley & Sons: New York, NY, USA, 2001. [Google Scholar]
  95. Fettinger, J.C.; Keogh, D.W.; Poli, R. Stable mononuclear, 17-electron molybdenum(III) carbonyl complexes. Synthesis, structure, thermal decomposition, and Cl addition reactions. J. Am. Chem. Soc. 1996, 118, 3617–3625. [Google Scholar] [CrossRef]
  96. Quadrelli, E.A.; Kraatz, H.-B.; Poli, R. Oxidation and protonation of transition metal hydrides: Role of an added base as proton shuttle and nature of protonated water in acetonitrile. Inorg. Chem. 1996, 35, 5154–5162. [Google Scholar] [CrossRef]
  97. Venâncio, A.I.F.; Kuznetsov, M.L.; Guedes da Silva, M.F.C.; Martins, L.M.D.R.S.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L. Metal-hydride bond activation and metal-metal interaction in dinuclear iron complexes with linking dinitriles: A synthetic, electrochemical and theoretical study. Inorg. Chem. 2002, 41, 6456–6467. [Google Scholar] [CrossRef] [PubMed]
  98. Martins, L.M.D.R.S.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L.; Henderson, R.A.; Evans, D.J.; Benetollo, F.; Bombieri, G.; Michelin, R.A. Syntheses, properties and mössbauer studies of cyanamide and cyanoguanidine complexes of iron(II). Crystal structures of trans-[FeH(NCNH2)(Ph2PCH2CH2PPh2)2] [BF4] and trans-[Fe(NCNEt2)2(Et2PCH2CH2PEt2)2][BF4]2. Inorg. Chim. Acta 1999, 291, 39–48. [Google Scholar] [CrossRef]
  99. Marchetti, F.; Pettinari, C.; Pettinari, R.; Cerquetella, A.; Cingolani, A.; Chan, E.J.; Kozawa, K.; Skelton, B.W.; White, A.H.; Wanke, R.; et al. Areneruthenium(II) 4-acyl-5-pyrazolonate derivatives: Coordination chemistry and reactivity. Inorg. Chem. 2007, 46, 8245–8257. [Google Scholar] [CrossRef] [PubMed]
  100. Lever, A.B.P. Electrochemical parametrization of rhenium redox couples. Inorg. Chem. 1991, 30, 1980–1985. [Google Scholar] [CrossRef]
  101. Lever, A.B.P. (Ed.) Comprehensive Coordination Chemistry II; Elsevier: Oxford, UK, 2004; Chapter 2.19; Volume 2, p. 251.
  102. Lu, S.; Strelets, V.V.; Ryan, M.F.; Pietro, W.J.; Lever, A.B.P. Electrochemical Parametrization in Sandwich Complexes of the First Row Transition Metals. Inorg. Chem. 1996, 35, 1013–1023. [Google Scholar] [CrossRef] [PubMed]
  103. Kopylovich, M.N.; Mahmudov, K.T.; Silva, M.F.C.G.; Martins, L.M.D.R.S.; Kuznetsov, M.L.; Silva, T.F.S.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L. Trends in properties of para-substituted 3-(phenylhydrazo)pentane-2,4-diones. J. Phys. Org. Chem. 2011, 24, 764–773. [Google Scholar] [CrossRef]
Figure 1. General scorpionate structure: poly(pyrazol-1-yl)borates for Z = B; poly(pyrazol-1-yl)methanes for Z = C.
Figure 1. General scorpionate structure: poly(pyrazol-1-yl)borates for Z = B; poly(pyrazol-1-yl)methanes for Z = C.
Catalysts 07 00012 g001
Figure 2. Structural general representation of tris(pyrazol-1-yl)borate (a); tris(pyrazol-1-yl)methane (b); and cyclopentadienyl (c) ligands.
Figure 2. Structural general representation of tris(pyrazol-1-yl)borate (a); tris(pyrazol-1-yl)methane (b); and cyclopentadienyl (c) ligands.
Catalysts 07 00012 g002
Figure 3. Structures of functionalized tris(pyrazol-1-yl)methanes: (a) CH3SO3CH2C(pz)3; (b) PyCH2OCH2C(pz)3 (Py = pyridine) or PyCH2OCH2C(3-Phpz)3; (c) SO3C(3-Phpz)3 and (d) HOCH2C(3-Phpz)3.
Figure 3. Structures of functionalized tris(pyrazol-1-yl)methanes: (a) CH3SO3CH2C(pz)3; (b) PyCH2OCH2C(pz)3 (Py = pyridine) or PyCH2OCH2C(3-Phpz)3; (c) SO3C(3-Phpz)3 and (d) HOCH2C(3-Phpz)3.
Catalysts 07 00012 g003
Figure 4. κ2–κ3 interchange coordination modes of a tris(pyrazol-1-yl)methane scorpionate ligand and comparison with a scorpion.
Figure 4. κ2–κ3 interchange coordination modes of a tris(pyrazol-1-yl)methane scorpionate ligand and comparison with a scorpion.
Catalysts 07 00012 g004
Figure 5. Selected C-homoscorpionate complexes exhibiting: (a) tetradentate coordination ability of the scorpionate ligand; (b) κ2-coordination of the scorpionate ligand at an octahedral geometry; (c) κ2-coordination of the scorpionate ligand at a square planar geometry (d) N3- or N2O-coordination of the scorpionate ligand.
Figure 5. Selected C-homoscorpionate complexes exhibiting: (a) tetradentate coordination ability of the scorpionate ligand; (b) κ2-coordination of the scorpionate ligand at an octahedral geometry; (c) κ2-coordination of the scorpionate ligand at a square planar geometry (d) N3- or N2O-coordination of the scorpionate ligand.
Catalysts 07 00012 g005
Scheme 1. One-pot carboxylation of methane to acetic acid catalyzed by C-scorpionate complexes [7,8,47,62].
Scheme 1. One-pot carboxylation of methane to acetic acid catalyzed by C-scorpionate complexes [7,8,47,62].
Catalysts 07 00012 sch001
Scheme 2. Peroxidative oxidation of cyclohexane to cyclohexanone and cyclohexanol (KA oil) in aqueous medium, catalyzed by C-scorpionate catalysts [7,8,13,43,46,47,48,49,57,58,59,60,61,62,63,64,65,66,67,68].
Scheme 2. Peroxidative oxidation of cyclohexane to cyclohexanone and cyclohexanol (KA oil) in aqueous medium, catalyzed by C-scorpionate catalysts [7,8,13,43,46,47,48,49,57,58,59,60,61,62,63,64,65,66,67,68].
Catalysts 07 00012 sch002
Scheme 3. Proposed mechanism for the carboxylation of an alkane to the corresponding carboxylic acid catalyzed by a C-homoscorpionate oxo-V complex.
Scheme 3. Proposed mechanism for the carboxylation of an alkane to the corresponding carboxylic acid catalyzed by a C-homoscorpionate oxo-V complex.
Catalysts 07 00012 sch003
Scheme 4. Proposed mechanism for the formation of OH and HOO radicals in the oxidation of an alkane with hydrogen peroxide, catalyzed by a C-homoscorpionate oxo-V complex.
Scheme 4. Proposed mechanism for the formation of OH and HOO radicals in the oxidation of an alkane with hydrogen peroxide, catalyzed by a C-homoscorpionate oxo-V complex.
Catalysts 07 00012 sch004
Scheme 5. Proposed mechanism for the epoxidation of alkenes catalyzed by a C-homoscorpionate aqua-Ru complex.
Scheme 5. Proposed mechanism for the epoxidation of alkenes catalyzed by a C-homoscorpionate aqua-Ru complex.
Catalysts 07 00012 sch005
Figure 6. Yields of: ( ) acetic acid produced from the one-pot oxidation of ethane catalyzed by the Re(III) complexes vs. their Re(III) oxidation potentials; ( ) 6-methylhexanolide obtained from Baeyer-Villiger (BV) oxidation of 2-methylcyclohexanone, vs. their Re(VII) oxidation potentials; and ( ) 6-methylhexanolide from BV oxidation of 2-methylcyclohexanone, vs. their Re(III) oxidation potentials.
Figure 6. Yields of: ( ) acetic acid produced from the one-pot oxidation of ethane catalyzed by the Re(III) complexes vs. their Re(III) oxidation potentials; ( ) 6-methylhexanolide obtained from Baeyer-Villiger (BV) oxidation of 2-methylcyclohexanone, vs. their Re(VII) oxidation potentials; and ( ) 6-methylhexanolide from BV oxidation of 2-methylcyclohexanone, vs. their Re(III) oxidation potentials.
Catalysts 07 00012 g006
Figure 7. Yields of acetic acid produced from the one-pot carboxylation of methane catalyzed by the V(V) complexes [VO23-HC(pz)3}][BF4], [VO23-HC(3,5-Me2pz)3}][BF4] and [VO23-SO3C(pz)3}] vs. their V(V) to V(IV) reduction potentials.
Figure 7. Yields of acetic acid produced from the one-pot carboxylation of methane catalyzed by the V(V) complexes [VO23-HC(pz)3}][BF4], [VO23-HC(3,5-Me2pz)3}][BF4] and [VO23-SO3C(pz)3}] vs. their V(V) to V(IV) reduction potentials.
Catalysts 07 00012 g007
Table 1. Cyclic voltammetric data a for metal C-homoscorpionate complexes.
Table 1. Cyclic voltammetric data a for metal C-homoscorpionate complexes.
C-Scorpionate CompoundRedox Potential/V vs. SCERef.
IEpox (IE½ox)IEpred (IE½red)IIEpred (IIE½red)
[VCl33-SO3C(pz)3}](1.14)--[91]
[VO23-SO3C(pz)3}] b-−0.46−1.82[47]
[VOCl23-CH3SO2OCH2C(pz)3}] c(1.35)−0.78-[49]
[VO23-HC(pz)3}][BF4] b-−0.28−1.70[47]
[VO23-HC(3,5-Me2pz)3}][BF4]-−0.37−1.75[47]
Li[Mo{κ3-SO3C(pz)3}(CO)3](0.18)--[50]
[Mo{κ3-SO3C(pz)3}I(CO)3]0.44--[50]
[Mo{κ3-SO3C(pz)3}H(CO)3]0.09--[50]
[ReCl23-HC(pz)3}(PPh3)][BF4] d(0.54)(−0.74)-[51]
[ReCl33-HC(pz)3}]1.14−0.62−1.70[52]
[ReCl33-HC(3,5-Me2pz)3}](1.25)(−0.13)(−0.72)[52]
[ReCl42-HC(pz)3}]1.79(−0.06)−1.50[52]
[ReO33-SO3C(pz)3}]-−0.83-[52]
[ReO{κ3-SO3C(pz)3}(HMT)] b(0.86)−0.83-[21]
[ReOCl{κ3-SO3C(pz)3}(PPh3)]Cl1.45(−0.94)(−1.41)[52]
[ReO32-HC(pz)3}(PTA)][ReO4] b-(−0.62)-[21]
[ReO3(Hpz)(HMT)][ReO4] b-(−0.33)-[21]
[FeCl23-CH3SO2OCH2C(pz)3}] c(1.06)−0.38-[49]
[FeCl33-HC(pz)3}] d(−0.11)--[92]
[FeCl33-HC(3,5-Me2pz)3}] d(−0.20)--[92]
[FeCl33-HC(3-iPrpz)3}] d(−0.04)--[92]
[Ru(p-cymene){κ3-SO3C(pz)3}]Cl(0.95)(−0.97)-[54]
[Ru(p-cymene){κ3-SO3C(pz)3}][BF4](0.96)(−0.97)-[54]
[Ru(p-cymene){κ3-SO3C(3-Phpz)3}]Cl1.02(−1.00)-[54]
[Ru(benzene){κ3-SO3C(pz)3}]Cl(1.07)(−0.87)-[54]
[Ru(benzene){κ3-SO3C(3-Phpz)3}]Cl(1.37)(−0.92)-[54]
[Ru(HMB){κ3-SO3C(pz)3}]Cl0.95(−1.11)-[54]
[Ru(cod)Cl{κ3-SO3C(pz)3}](0.96)(−1.10)-[54]
[Ru(cod)Cl{κ3-SO3C(3-Phpz)3}](0.99)(−1.27)-[54]
[RuCl{κ3-HC(pz)3}(bqdi)][ClO4] c(0.82)(−0.79)−1.39[89]
[Ru(H2O){κ3-HC(pz)3}(bqdi)][ClO4]2 c(0.44)--[89]
[Ru{κ3-HC(3,5-Me2pz)3}(NCCH3)3][BF4]2 c(0.42)--[93]
[Ru{κ3-HC(3,5-Ph2pz)3}(NCCH3)3][BF4]2 c(0.71)--[93]
[Co(OSO3H)(OCH3)(HOCH3){κ3-HC(pz)3}] b1.03−0.40 [12]
[Co{κ3-HOCH2C(pz)3}2](NO3)2(0.58)−0.68 [12]
[Co{κ3-HOCH2C(pz)3}2]·[Co{κ3-HOCH2C(pz)3} (H2O)3]2(Cl)6·6H2O(0.60)−0.67−1.21[12]
[CoCl2(H2O){κ3-PyCH2OCH2C(pz)3}]1.28−0.60-[12]
[CoCl2(H2O){κ3-CH3SO2OCH2C(pz)3}] c1.10−0.64-[12]
[CuCl23-CH3SO2OCH2C(pz)3}] c-−0.70-[49]
[AuCl22-HC(pz)3}]Cl c-−0.02−0.60[57]
[AuCl22-HOCH2C(pz)3}]Cl c-−0.01−0.58[57]
[AuCl22-HC(3,5-Me2pz)3}]Cl c-−0.11−0.69[57]
a Values in V ± 0.02 relative to SCE; in CH2Cl2; scan rate of 200 mV·s−1. Values for reversible waves are given in brackets. bqdi = o-benzoquinonediimine; 3-iPr = iso-propyl group; b In dimethyl sulfoxide (DMSO); c In acetonitrile (NCMe); d In dimethylformamide (DMF). SCE = saturated calomel electrode.
Table 2. Electrochemical EL Lever ligand parameter for C-homoscorpionate ligands.
Table 2. Electrochemical EL Lever ligand parameter for C-homoscorpionate ligands.
Tris(pyrazol-1-yl)methaneEL/V vs. SHE a
HC(pz)30.14
{SO3C(pz)3}−0.09
{SO3C(3-Phpz)3}−0.05
a for each coordinated pyrazol-1-yl group.

Share and Cite

MDPI and ACS Style

Martins, L.M.D.R.S. C-Homoscorpionate Oxidation Catalysts—Electrochemical and Catalytic Activity. Catalysts 2017, 7, 12. https://doi.org/10.3390/catal7010012

AMA Style

Martins LMDRS. C-Homoscorpionate Oxidation Catalysts—Electrochemical and Catalytic Activity. Catalysts. 2017; 7(1):12. https://doi.org/10.3390/catal7010012

Chicago/Turabian Style

Martins, Luísa M. D. R. S. 2017. "C-Homoscorpionate Oxidation Catalysts—Electrochemical and Catalytic Activity" Catalysts 7, no. 1: 12. https://doi.org/10.3390/catal7010012

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

Article Metrics

Back to TopTop