Next Article in Journal
Susceptibility of the Formate Hydrogenlyase Reaction to the Protonophore CCCP Depends on the Total Hydrogenase Composition
Previous Article in Journal
Cyclometalated Ir(III) Complexes with Curcuminoid Ligands as Active Second-Order NLO Chromophores and Building Blocks for SHG Polymeric Films
Previous Article in Special Issue
Structures and Properties of Dinitrosyl Iron and Cobalt Complexes Ligated by Bis(3,5-diisopropyl-1-pyrazolyl)methane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of Catecholato Copper(II) Complexes with Sterically Hindered Neutral and Anionic N3 Type Ligands: Tris(3,5-diisopropyl-1-pyrazolyl)methane and Hydrotris(3,5-diisopropyl-1-pyrazolyl)borate

1
Department of Chemistry, Ibaraki University, Mito, Ibaraki 310-8512, Japan
2
Department of Chemistry, Graduate School of Pure and Applied Sciences, University of Tsukuba, Tsukuba 305-8571, Japan
*
Author to whom correspondence should be addressed.
Submission received: 17 April 2020 / Revised: 7 May 2020 / Accepted: 11 May 2020 / Published: 18 May 2020
(This article belongs to the Special Issue Scorpionates: Recent Advances)

Abstract

:
Three catecholato copper(II) complexes, [Cu(catCl4)(L1′)], [Cu(catBr4)(L1′)], and [Cu(catCl4)(L1H)], supported by sterically hindered neutral and anionic N3 type ligands: tris(3,5-diisopropyl-1-pyrazolyl)methane (referred to as L1′) and hydrotris(3,5-diisopropyl-1-pyrazolyl)borate (referred to as L1), are synthesized and characterized in detail. Their X-ray structures reveal that both [Cu(catCl4)(L1′)] and [Cu(catBr4)(L1′)] complexes have a five-coordinate square-pyramidal geometry and [Cu(catCl4)(L1H)] complex has a four-coordinate square-planar geometry. The L1H is unusual protonated ligand that controls its overall charge. For the three catecholato copper(II) complexes, the oxidation state of copper is divalent, and catechol exists in catecholate as two minus anion. This difference in coordination geometry affects their d-d and CT transitions energy and ESR parameters.

Graphical Abstract

1. Introduction

Transition metal complexes ligated by the hydrotris(pyrazolyl)borate as an anionic nitrogen-containing tripod ligand, as first prepared by Professor S. Trofimenko in 1966, are widely studied compounds [1]. An important advance in this chemistry is the introduction of alkyl substitutions of the pyrazolyl rings at the 3 (and 5) position(s) to prevent the formation of an inert hexa-coordinate compound [2]. The coordination behavior of transition metal complexes can easily be changed by introducing substituents with different electronic and steric properties on the pyrazolyl rings. Therefore, transition metal complexes based on these ligands have attracted a great deal of interest, and are still undergoing many investigations [3,4].
On the other hand, isoelectronic tris(pyrazolyl)methane ligands have received less attention. This ligand is formally derived from hydrotris(pyrazolyl)borate ligand, in which the central boron atom is replaced by a carbon atom. This tris(pyrazolyl)methane was also prepared by Professor S. Trofimenko in 1970 as a neutral nitrogen-containing tripod ligand [5]. Some researchers have improved the synthetic method of this ligand, including Elguero and co-workers [6], Reger and co-workers [7], and us [8,9]. We found that tris(pyrazolyl)methane could be synthesized in high yields using an autoclave [8,9].
By using these ligands of L (hydrotris(pyrazolyl)borate) and L’ (tris(pyrazolyl)methane), structural differences in chlorido and nitrato copper(II) complexes were observed: neutral four-coordinate [CuCl(L)] versus neutral five-coordinate [CuCl2(L’)] [8], neutral four-coordinate [CuCl(L)] versus cationic four-coordinate [CuCl(L’)](ClO4) [10], mononuclear [CuCl(L)] versus binuclear [Cu(μ-Cl)(L’)]2(ClO4)2 [11] and bidentate [Cu(κ2-O2NO)(L)] versus monodentate [Cu(κ-ONO2)2(L’)] [8,12], and their different electron donation properties in copper(I) carbonyl complexes: [Cu(CO)(L)] (ν(CO), 2056 cm−1) versus [Cu(CO)(L’)](PF6) (ν(CO), 2107 cm−1) [9].
In this research, we expanded our copper(II) coordination chemistry toward catecholato complex, since the interaction between copper and catechol has been reported in copper-containing proteins such as catechol oxidase and tyrosinase [13,14,15,16]. Moreover, catechol itself is known as “non-innocent” ligand [17,18,19,20,21,22,23,24] and is a redox active ligand that reversibly undergoes one-electron sequences illustrated in Scheme 1.
We report herein our catecholato copper(II) chemistry with sterically hindered neutral tris(3,5-diisopropyl-1-pyrazolyl)methane (L1′) and anionic hydrotris(3,5-diisopropyl-1-pyrazolyl)borate (L1), as shown in Scheme 2. The structures of catecholato and o-semiquinonato copper(II) complexes have been determined and deposited in the Cambridge Structural Database [25]. Their structural parameters are listed in Scheme 3 and Table 1 [26,27,28,29,30,31,32,33,34,35,36,37]. The reported catechol-copper(II) coordination can be divided into two groups: catecholato copper(II) complex and o-semiquinonato copper(II) complex. Regarding the structural parameters of catechol-copper(II) coordination, it can be characterized by the bond lengths of d (C–O) and d (C1–C2) in catechol: > 1.3 Å and < ~1.45 Å for catecholato copper(II) complex and < 1.3 Å and > ~1.45 Å for o-semiquinonato copper(II) complex, respectively. The magnetic properties in catechol-copper(II) coordination may also provide good evidence for defining the redox isomers of catechol. With the foregoing in mind and motivated to obtain greater insight into the structural and spectroscopic differences between hindered neutral L1′ and anionic L1, we prepared three catecholato copper(II) complexes, namely [Cu(catCl4)(L1′)], [Cu(catBr4)(L1′)], and [Cu(catCl4)(L1H)]. All complexes were characterized by single crystal X-ray crystallography and spectroscopic techniques, viz. IR/far-IR, UV-Vis, and ESR spectroscopy. Several decisive differences in their structures and physicochemical properties of catecholato copper(II) complexes were observed and are discussed in detail.

2. Results and Discussion

2.1. Synthesis

All of the four complexes prepared as shown in Scheme 4 gave a satisfactory elemental analysis. The reaction of mononuclear chlorido copper(II) [CuCl2(L1′)] with suitable catechols, catH2X4 (X = Cl and Br) and NEt3 at −50 °C yielded green colored mononuclear catecholato complexes, [Cu(CatCl4)(L1′)] and [Cu(CatBr4)(L1′)], respectively. On the other hand, purple colored mononuclear catecholato complex [Cu(CatCl4)(L1H)] was synthesized using the binuclear hydroxido complex [{Cu(L1)}(μ-OH)]2.

2.2. Structure

Successful single-crystal X-ray structural analyses were performed on compounds [Cu(catCl4)(L1′)], [Cu(catBr4)(L1′)], and [Cu(catCl4)(L1H)]. The perspective drawings of all three complexes are shown in Figure 1 and Figure 2. The selected bond distances and bond angles of the obtained complexes are summarized in their figure captions.
The coordination geometry around the copper(II) ion in [Cu(catCl4)(L1′)] and [Cu(catBr4)(L1′)] is essentially a five-coordinate square-pyramidal geometry with the basal plane comprising two nitrogen atoms (N11 and N31) from tris(pyrazolyl)methane and two oxygen atoms (O40 and O45) from catechol and whose axial site is occupied by the remaining nitrogen atom (N21) (Figure 1). This coordination geometry is supported by the structural parameter τ5 values (0.02 in [Cu(catCl4)(L1′)] and 0.03 in [Cu(catBr4)(L1′)]) [38]. The apical deviation of the copper(II) ion from the corresponding least-squares N2O2 basal plane is 0.15 Å in [Cu(catCl4)(L1′)] and 0.17 Å in [Cu(catBr4)(L1′)]. The averaged Cu–O distances of 1.919(2) Å in [Cu(catCl4)(L1′)] and 1.917(3) Å in [Cu(catBr4)(L1′)] and the averaged C–O distances of 1.317(3) Å in [Cu(catCl4)(L1′)] and 1.318(4) Å in [Cu(catBr4)(L1′)]. In general, the C–O bond distances in catCl42− (3,4,5,6-tetrachlorocatecholate) and sqCl4●− (3,4,5,6-tetrachloro-o-semiquinate) fall into the range of 1.30–1.36 Å and 1.26–1.31 Å, respectively (Table 1). Therefore, the coordinated catechol group in [Cu(catCl4)(L1′)] and [Cu(catBr4)(L1′)] is best described as catX42−, rather than either sqX4●− or qX4 ((3,4,5,6-tetrahalogeno-o-quinone) (Scheme 1), and both copper are divalent. This oxidation assignment is consistent with the data of d-d transitions in UV-Vis spectra and ESR parameters in ESR spectra (vide infra). The dihedral angle between apical pyrazole and catechol ring is 99.3° in [Cu(catCl4)(L1′)] and 90.0° in [Cu(catBr4)(L1′)]. Therefore, the apical pyrazole locates in a nearly vertical position.
In contrast to the L1′ complexes, the L1 complex is something different. Its structure revealed that hydrotris(pyrazolyl)borate adopts an unusual bidentate mode of coordination mode with one dangling pyrazole ring. The apical nitrogen (N21) is located away from the copper(II) center; the distance between Cu1 and N21 is 2.991(6) Å. The apical deviation of the copper(II) ion from the corresponding least-squares N2O2 basal plane is 0.11 Å, indicating that the coordination geometry around the copper(II) ion in [Cu(catCl4)(L1H)] is essentially a four-coordinate square-planar geometry whose basal plane consists of two nitrogen atoms (N11 and N31) from hydrotris(pyrazolyl)borate and two oxygen atoms (O40 and O45) from catechol. The dihedral angle between apical pyrazole and catechol ring is 71.2°. The distance of N21···O40 is 4.112(7) Å and that of N21···O45 is 2.903(7) Å. Therefore, the dangling pyrazole ring is tilted toward N31 pyrazole ring. From the averaged Cu–O distances of 1.921(4) Å and the averaged C–O distances of 1.335(8) Å, the coordinated catechol group is best described as catCl42− rather than either sqCl4●− or qCl4 (Scheme 1 and Table 1) as well as the L1′ complexes and this copper is also divalent. This oxidation assignment is also consistent with the UV-Vis and ESR data (vide infra). The charge consideration of the catechol and the copper oxidation state indicates the anionic hydrotris(pyrazolyl)borate should be neutral. This unusual behavior suggests the dangling pyrazole must be protonated at the apical nitrogen (N21). This behavior is supported by ν(N–H) in its IR spectrum (vide infra) and intramolecular hydrogen bond between N21 and O45 (2.903(7) Å). This protonated hydrotris(pyrazolyl)borate is very rare. The first example of copper(II) complex was reported by Professor S. Trofimenko in 1994 [39]. Other reported examples include V(IV) complex [40], Mn(II) complex [41], Pt(II) complexes [42,43,44], and Pt(IV) complex [45].

2.3. IR Spectroscopy

IR spectra of the three catecholato complexes were measured using KBr pellets as shown in Figure 3. One typical C=N stretching vibration in both [Cu(catCl4)(L1′)] and [Cu(catBr4)(L1′)] complexes was shifted to 1557 and 1556 cm−1 from the corresponding L1′ ligand at 1553 cm−1, respectively [9]. Coordinated catechols typically show two strong stretching bands attributed to the ring stretching (ν(C–C)) and the CO group stretching (ν(C–O)) around ~1500 cm−1 and ~1350 cm−1, respectively [46,47,48,49,50]. In the L1′ complexes [Cu(catCl4)(L1′)] and [Cu(catBr4)(L1′)], two broad characteristic bands were observed at 1466 and 1400 cm−1, corresponding to the ring stretching (ν(C–C)) and the CO group stretching (ν(C–O)), respectively. In the L1 complex [Cu(catCl4)(L1H)], the ν(C=N) band split to 1569 and 1537 cm−1 due to different pyrazole ring environment. Intensity and stretching energy indicate that former is derived from the protonated pyrazole ring. Catechol ring stretching was also observed at 1453 and 1377 cm−1. Moreover, the ν(N–H) band was observed at 3142 cm−1. This is the first observation of the protonated pyrazole by IR spectroscopy.
Far-IR spectra of the three catecholato complexes were measured using CsI pellets as shown in Figure 4. From the literature, the M–O stretching bands of the coordinated catechol with transition metals were observed between 500 and 600 cm−1 [48,49,50]. However, the M–O stretching band values for the three catecholato complexes shown in Figure 4 are so broad and complicated that these assignments are very difficult at this stage and require more experiments and DFT calculations to make reliable assignments.

2.4. UV-Vis Spectroscopy

UV-Vis absorption spectra of the catecholato complexes [Cu(catCl4)(L1′)], [Cu(catBr4)(L1′)], and [Cu(catCl4)(L1H)] are shown in Figure 5. As mentioned in Section 2.1, the structures of the L1′ complexes [Cu(catCl4)(L1′)] and [Cu(catBr4)(L1′)] contain a five-coordinate square-pyramidal geometry. This indicates that the d-d transition energies of the L1′ complexes are almost identical at 659 and 971 nm for [Cu(catCl4)(L1′)] and 654 and 972 nm for [Cu(catBr4)(L1′)]. However, in [Cu(catCl4)(L1H)], a different energy was observed for its d–d transition at 830 nm. This energy gap is consistent with other copper(II) coordination chemistry [14,51,52,53]. In the visible region, the L1′ complexes [Cu(catCl4)(L1′)] and [Cu(catBr4)(L1′)] exhibit characteristic bands at 505 nm and 503 nm. This band attributes to catecholate2− to Cu(II) CT transition. For the latter, an additional peak also appears at 380 nm. One the other hand, the L1 complex [Cu(catCl4)(L1H)] has a moderately intense band at 494 nm. The differences in these CT bands indicate the colors of the complexes are different: L1′ complexes [Cu(catCl4)(L1′)] and [Cu(catBr4)(L1′)] are green and L1 complex [Cu(catCl4)(L1H)] is purple.

2.5. ESR Spectroscopy

The frozen glass ESR spectra of the complexes at 137 K are presented in Figure 6. The ESR spectra show that these complexes have an S = 1/2 ground state. The order of g|| > g⊥ > 2.0023 is satisfied in all the complexes, confirming the presence of unpaired electron of the copper(II) ion in dx2−y2 orbital. These observations are consistent with the above consideration that catechol is coordinated as catecholate two minus anion. Some differences in the ESR parameters in all the complexes are caused by different coordination geometries: five-coordinate square-pyramidal and four-coordinate square-planar. The ESR parameters of the reported Cu(II)-catBu2 or Cu(II)-catCl4 complexes are also consistent with an S = 1/2 ground state [27,28,29,30,33]. On the other hand, the reported magnetism of Cu(II)-sqBu2 or sqCl4 complexes are diamagnetism (ESR silent) [27,33,34,36,37] or ferromagnetism [35].

3. Materials and Methods

3.1. Material and General Techniques

Preparation and handling of all the complexes was performed under an argon atmosphere using standard Schlenk tube techniques or in a VAC inert atmosphere glovebox containing argon gas. Dichloromethane and acetonitrile were carefully purified by refluxing and distilling under an argon atmosphere over phosphorous pentoxide and calcium hydride prior to use, respectively [54]. Other reagents are commercially available and were used without further purification. [CuCl2(L1′)] [8,11] and [{Cu(L1)}(μ-OH)]2 [55] were prepared using the published methods.

3.2. Instrumentation

IR spectra (4000–400 cm−1) and far-IR (650–150 cm−1) spectra were recorded on KBr pellets and on CsI pellets, respectively, using a JASCO FT/IR-550 spectrophotometer (JASCO, Tokyo, Japan). Abbreviations used in the description of the vibration data are as follows: vs, very strong; s, strong; m, medium; w, weak. UV-Vis spectra at low temperature were measured on an Otsuka Electronics MCPD-2000 system (Otsuka Electronics, Tokyo, Japan) with an optical fiber attachment (300–1100 nm). ESR spectra as frozen solutions (dichloromethane/1,2-dichloroethane) were recorded on a Bruker EMX-T ESR spectrometer (Bruker Japan, Yokohama, Japan) at 137 K in quartz tubes (diameter 5 mm) with a liquid nitrogen temperature controller BVT 3000. The elemental analyses (C, H, N) were performed by the Chemical Analysis Center at the University of Tsukuba.

3.3. Preparation of Complexes

3.3.1. [Cu(catCl4)(L1′)]

To a solution of [CuCl2(L1′)] (160 mg, 0.266 mmol) in dichloromethane (40 cm3) was added tetrachlorocatechol (81.5 mg, 0.329 mmol) and triethylamine (72.6 mg, 0.717 mmol) dissolved in dichloromethane (10 cm3) at −50 °C and the solution was stirred at −50 °C for 30 min. During the reaction, the color of the solution gradually turned from yellow-green to green. After it was stirred at 0 °C for 30 min, the solvent was evaporated under vacuum. The resulting solid was extracted with acetonitrile (40 mL). The filtrate was evaporated under vacuum and green powder was obtained. The green crystals were crystallized from acetonitrile at −30 °C. Single crystals were obtained from acetonitrile at −30 °C.
Yield: 75% (155 mg, 0.200 mmol). Calcd for C34H46Cl4CuN6O2: C, 52.62; H, 5.97; N, 10.83. Found: C, 52.48; H, 6.19; N, 10.81. IR (KBr, ν/cm−1): 2966 s, 2930 m, 2871 m, 1557 m, 1469 vs, 1400 m, 1385 m, 1292 s, 1237 m, 1181 m, 1110 w, 1064 m, 975 s, 803 s, 725 w, 669 m. Far-IR (CsI, ν/cm−1): 635 w, 585 vs, 570 vs, 530 m, 470 s, 439 m, 409 w, 366 m, 329 s, 282 vs, 237 w, 211 vs. EPR (137 K, dichloromethane/1,2-dichloroethane) g|| 2.30, A|| 161 G, g⊥ 2.07, A⊥ 15 G. 161 G. UV-Vis (dichloromethane, λmax/nm (ε/cm−1 mol−1 dm3)) at 273 K: 505 (100), 659 (80), 971 (30).

3.3.2. [Cu(catBr4)(L1′)]

To a solution of [CuCl2(L1′)] (108 mg, 0.180 mmol) in dichloromethane (25 cm3) was added tetrabromocatechol (87.0 mg, 0.204 mmol) and triethylamine (43.6 mg, 0.431 mmol) dissolved in dichloromethane (10 cm3) at −50 °C, and the solution was stirred at −50 °C for 30 min. The color of the solution gradually turned from yellow-green to green. After it was stirred at 0 °C for 30 min, the solvent was evaporated under vacuum. The resulting solid was extracted with acetonitrile (50 mL). The filtrate was evaporated under vacuum and green powder was obtained. The green crystals were crystallized from acetonitrile at −30 °C. Single crystals were obtained from acetonitrile at −30 °C.
Yield: 75% (129 mg, 0.135 mmol). Calcd for C34H46Br4CuN6O2: C, 42.81; H, 4.86; N, 8.81. Found: C, 42.60; H, 4.70; N, 8.74. IR (KBr, ν/cm−1): 2967 s, 1556 m, 1455 vs, 1400 m, 1268 s, 1238 s, 1181 w, 1062 m, 927 m, 823 s, 739 m, 669 w. Far-IR (CsI, ν/cm−1): 637 w, 616 s, 569 s, 526 m, 499 w, 483 w, 410 vs, 358 s, 304 m, 255 vs, 202 s. EPR (137 K, dichloromethane/1,2-dichloroethane) g|| 2.28, A|| 163 G, g⊥ 2.06, A⊥ 15 G. UV-Vis (dichloromethane, λmax/nm (ε/cm−1 mol−1 dm3)) at 273 K: 380 (380), 503 (100), 659 (80), 985 (30).

3.3.3. [Cu(catCl4)(L1H)]

To a solution of [{Cu(L1)}(μ-OH)]2 (221 mg, 0.203 mmol) in dichloromethane (20 cm3) was added tetrachlorocatechol (116 mg, 0.468 mmol) dissolved in dichloromethane (15 cm3) at −50 °C, and the solution was stirred at −50 °C for 30 min. The color of the solution gradually turned purple. After it was stirred at 0 °C for 30 min, the solution was concentrated under vacuum. After this, the solution was cooled at −50 °C to give purple crystals. Single crystals were obtained from dichloromethane at −50 °C.
Yield: 59% (186 mg, 0.240 mmol). Anal. Calcd for C33H47BCl4CuN6O2: C, 51.08; H, 6.11; N, 10.83. Found: C, 50.87; H, 6.03; N, 10.56. IR (KBr, ν/cm−1): 3142 w, 2969 s, 2871 m, 2537 w, 1569 m, 1537 m, 1453 vs, 1377 s, 1308 m, 1258 s, 1181 m, 1059 s, 972 s, 805 s, 741 m, 634 w. Far-IR (KBr, ν/cm−1): 640 s, 578 s, 529 m, 478 s, 438 w, 397 s, 329 m, 285 vs, 220 s, 175 w. EPR (137 K, dichloromethane/1,2- dichloroethane) g|| 2.27, A|| 156 G, g⊥ 2.06, A⊥ 16 G. UV-Vis (dichloromethane, λmax/nm (ε/cm−1 mol−1 dm3)) at 273 K: 494 (380), 830 (120); at 223 K: 490 (400), 821 (120); at 195 K: 488 (440), 821 (130).

3.4. X-Ray Crystal Structure Determination

Crystal data and refinement parameters for the three catecholato copper(II) complexes [Cu(catCl4)(L1′)]·2.5(CH3CN), [Cu(catBr4)(L1′)]·2.5(CH3CN), and [Cu(catCl4)(L1H)] are given in Table 2. All crystallographic data have been deposited at the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK and copies can be obtained on request, free of charge, by quoting the publication citation and the deposition numbers. CCDC numbers: 620467 for [Cu(catCl4)(L1′)]·2.5(CH3CN), 620468 for [Cu(catBr4)(L1′)]·2.5(CH3CN), and 620469 for [Cu(catCl4)(L1H)] (CIF and the checkCIF output files can be found at Supplementary Materials).
The diffraction data were measured on a Rigaku/MSC Mercury CCD system (Rigaku, Tokyo, Japan) with graphite monochromated Mo Kα (λ = 0.71070 Å) radiation at low temperature. The unit cell parameters of each crystal were determined using CrystalClear [56] from 6 images. The crystal to detector distance was ca. 45 mm. Data were collected using 0.5° intervals in φ and ω to a maximum 2θ value of 55.0°. A total of 744 oscillation images were collected. The highly redundant data sets were reduced using CrystalClear and corrected for Lorentz and polarization effects [56]. An empirical absorption correction was applied for each complex. Structures were solved by direct methods (SIR92 and SIR97) [57,58] and heavy-atom Patterson methods [59]. The position of the copper ions and their first coordination sphere were located from a direct method E-map; other non-hydrogen atoms were found in alternating difference Fourier syntheses, and least squares refinement cycles. During the final refinement cycles the temperature factors were refined anisotropically. Refinement was carried out by a full matrix least-squares method on F2. All calculations were performed with the CrystalStructure [60] crystallographic software package except for refinement, which was performed using SHELXL 2013 [61]. Hydrogen atoms were placed in calculated positions. Sheldrick weighting scheme was used. Crystallographic data and structure refinement parameters including the final discrepancies (R and Rw) are listed in Table 2. The crystals of [Cu(catCl4)(L1H)] show a slightly lower quality of diffraction and some carbon atoms were disordered. Moreover, the solvent molecules in this crystal were highly disordered. Therefore, PLATON SQUEEZE was used to account for severely disordered solvent molecules [62].

4. Conclusions

The three catecholato copper(II) complexes [Cu(catCl4)(L1′)], [Cu(catBr4)(L1′)], and [Cu(catCl4)(L1H)] ligated by sterically hindered neutral tris(3,5-diisopropyl-1-pyrazolyl)methane (L1′) and anionic hydrotris(3,5-diisopropyl-1-pyrazolyl)borate (L1) were synthesized and characterized. The structures of the synthesized complexes were as follows: the L1′ complexes, [Cu(catCl4)(L1′)] and [Cu(catBr4)(L1′)], adopted a five-coordinate square-pyramidal geometry, whereas the L1 complex [Cu(catCl4)(L1H)] adopted a four-coordinate square-planar geometry. In [Cu(catCl4)(L1H)], a protonated pyrazole was found to be present in the apical position to neutralize its charge. This anomalous N–H bond was observed in its IR spectrum. To our knowledge, this is the first direct proof for this protonated nitrogen in hydrotris(pyrazolyl)borate. These geometrical changes in catecholato copper(II) complexes were their d–d and CT transitions in UV-Vis absorption spectra and ESR parameters in ESR spectra. The direct structural and physicochemical changes due to different ligand charges were observed. Therefore, several tris(pyrazolyl)methane ligands were extensively explored [63,64]. We are also underway in determining how the structure of transition metals is affected by ligand charges and environments. Moreover, the anti-oxidant activity of our redox catecholato copper(II) complexes is also of interest.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2304-6740/8/5/37/s1, the CIF and the checkCIF output files of [Cu(catCl4)(L1′)]·2.5(CH3CN), [Cu(catBr4)(L1′)]·2.5(CH3CN) and [Cu(catCl4)(L1H)].

Author Contributions

K.F. conceived and designed the project. T.O. and M.O. performed the experiments. M.O. and K.F. analyzed the data. K.F. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

We are grateful for the support from the Japan Society for the Promotion of Science (JSPS) and an Ibaraki University Priority Research Grant. We also acknowledge K. Okamoto and Y. Miyashita (University of Tsukuba) for their helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Trofimenko, S. Boron-pyrazole chemistry. J. Am. Chem. Soc. 1966, 88, 1842–1844. [Google Scholar] [CrossRef]
  2. Trofimenko, S.; Calabrese, J.C.; Thompson, J.S. Novel polypyrazolylborate ligands: Coordination control through 3-substituents of the pyrazole ring. Inorg. Chem. 1987, 26, 1507–1514. [Google Scholar] [CrossRef]
  3. Trofimenko, S. Scorpionates: The Coordination Chemistry of Polypyrazolylborate Ligands; Imperial College Press: London, UK, 1999. [Google Scholar]
  4. Pettinari, C. Scorpionates II: Chelating Borate Ligands; Imperial College Press: London, UK, 2008. [Google Scholar]
  5. Trofimenko, S. Geminal poly(1-pyrazolyl)alkanes and their coordination chemistry. J. Am. Chem. Soc. 1970, 92, 5118–5126. [Google Scholar] [CrossRef]
  6. 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]
  7. 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]
  8. Fujisawa, K.; Ono, T.; Aoki, H.; Ishikawa, Y.; Miyashita, Y.; Okamoto, K.; Nakazawa, H.; Higashimura, H. Copper(II) complexes with a novel tris(3,5-diisopropyl-1-pyrazolyl)methane ligand, [Cu(X2){HC(3,5-iPr2pz)3}] (X = Cl and NO3). Inorg. Chem. Commun. 2004, 7, 330–332. [Google Scholar] [CrossRef]
  9. Fujisawa, K.; Ono, T.; Ishikawa, Y.; Amir, N.; Miyashita, Y.; Okamoto, K.; Lehnert, N. Structural and electronic differences of copper(I) complexes with tris(pyrazolyl)methane and hydrotris(pyrazolyl)borate ligands. Inorg. Chem. 2006, 45, 1698–1713. [Google Scholar] [CrossRef]
  10. Fujisawa, K.; Iwamoto, H.; Tobita, K.; Miyashita, Y.; Okamoto, K. Copper(II) nitrato and chloro complexes with sterically hindered tridentate ligands: Influence of ligand framework and charge on their structure and physicochemical properties. Inorg. Chim. Acta 2009, 362, 4500–4509. [Google Scholar] [CrossRef]
  11. Fujisawa, K.; Tobita, K.; Sakuma, S.; Savard, D.; Leznoff, D.B. Binuclear and mononuclear copper(II) chlorido complexes with hindered neutral N3 type ligands: Influence of ligand framework and charge on their structure and physicochemical properties. Inorg. Chim. Acta 2019, 486, 582–588. [Google Scholar] [CrossRef]
  12. Lehnert, N.; Cornelissen, U.; Neese, F.; Ono, T.; Noguchi, Y.; Okamoto, K.; Fujisawa, K. Synthesis and spectroscopic characterization of copper(II)–nitrito complexes with hydrotris(pyrazolyl)borate and related coligands. Inorg. Chem. 2007, 46, 3916–3933. [Google Scholar] [CrossRef]
  13. Quist, D.A.; Diaz, D.E.; Liu, J.L.; Karlin, K.D. Activation of dioxygen by copper metalloproteins and insights from model complexes. J. Biol. Inorg. Chem. 2017, 22, 253–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Solomon, E.I.; Heppner, D.E.; Johnston, E.M.; Ginsbach, J.W.; Cirera, J.; Qayyum, M.; Kieber-Emmons, M.T.; Kjaergaard, C.H.; Hadt, R.G.; Tian, L. Copper active sites in biology. Chem. Rev. 2014, 114, 3659–3853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Qayyum, M.F.; Sarangi, R.; Fujisawa, K.; Stack, T.D.P.; Karlin, K.D.; Hodgson, K.O.; Hedman, B.; Solomon, E.I. L-Edge X-ray absorption spectroscopy and DFT calculations on Cu2O2 species: Direct electrophilic aromatic attack by side-on peroxo bridged dicopper(II) complexes. J. Am. Chem. Soc. 2013, 135, 17417–17431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Mirica, L.M.; Vance, M.; Rudd, D.J.; Hedman, B.; Hodgson, K.O.; Solomon, E.I.; Stack, T.D.P. Tyrosinase reactivity in a model complex: An alternative hydroxylation mechanism. Science 2005, 308, 1890–1892. [Google Scholar] [CrossRef] [Green Version]
  17. Tezgerevska, T.; Alley, K.G.; Boskovic, C. Valence tautomerism in metal complexes: Stimulated and reversible intramolecular electron transfer between metal centers and organic ligands. Coord. Chem. Rev. 2014, 268, 23–40. [Google Scholar] [CrossRef]
  18. Luca, O.R.; Crabtree, R.H. Redox-active ligands in catalysis. Chem. Soc. Rev. 2013, 42, 1440–1459. [Google Scholar] [CrossRef]
  19. Kaim, W.; Schwederski, B. Non-innocent ligands in bioinorganic chemistry—An overview. Coord. Chem. Rev. 2010, 254, 1580–1588. [Google Scholar] [CrossRef]
  20. Ray, K.; Petrenko, T.; Wieghardt, K.; Neese, F. Joint spectroscopic and theoretical investigations of transition metal complexes involving non-innocent ligands. Dalton Trans. 2007, 1552–1566. [Google Scholar] [CrossRef]
  21. Zanello, P.; Corsini, M. Homoleptic, mononuclear transition metal complexes of 1,2-dioxolenes: Updating their electrochemical-to-structural (X-ray) properties. Coord. Chem. Rev. 2006, 250, 2000–2022. [Google Scholar] [CrossRef]
  22. Ward, M.D.; McCleverty, J.A. Non-innocent behaviour in mononuclear and polynuclear complexes: Consequences for redox and electronic spectroscopic properties. Dalton Trans. 2002, 275–288. [Google Scholar] [CrossRef]
  23. Pierpont, C.G. Unique properties of transition metal quinone complexes of the MQ3 series. Coord. Chem. Rev. 2001, 219–221, 415–433. [Google Scholar] [CrossRef]
  24. Pierpont, C.G. Studies on charge distribution and valence tautomerism in transition metal complexes of catecholate and semiquinonate ligands. Coord. Chem. Rev. 2001, 216–217, 99–125. [Google Scholar] [CrossRef]
  25. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The Cambridge structural database. Acta Cryst. 2016, B72, 171–179. [Google Scholar] [CrossRef] [PubMed]
  26. Davidson, R.A.; Hao, J.; Rheingold, A.L.; Miller, J.S. High spin ground state copper(II) and nickel(II) complexes possessing the 3,5-di-tert-butyl-1,2-semiquinonate radical anion. Polyhedron 2017, 133, 348–357. [Google Scholar] [CrossRef]
  27. Verma, P.; Weir, J.; Mirica, L.; Stack, T.D.P. Tale of a twist: Magnetic and optical switching in copper(II) semiquinone complexes. Inorg. Chem. 2011, 50, 9816–9825. [Google Scholar] [CrossRef]
  28. Rall, J.; Wanner, M.; Albrecht, M.; Hornung, F.; Kaim, W. Sensitive valence tautomer equilibrium of paramagnetic complexes [(L)Cun+(Qn)] (n = 1 or 2; Q = quinones) related to amine oxidase enzymes. Chem. Eur. J. 1999, 5, 2802–2809. [Google Scholar] [CrossRef]
  29. Speier, G.; Tisza, S.; Tyeklár, Z.; Lange, C.W.; Pierpont, C.G. Coligand-dependent shifts in charge distribution for copper complexes containing 3,5-di-tert-butylcatecholate and 3,5-di-tert-butylsemiquinonate ligands. Inorg. Chem. 1994, 33, 2041–2045. [Google Scholar] [CrossRef]
  30. Buchanan, R.M.; Wilson-Blumenberg, C.; Trapp, C.; Larsen, S.K.; Greene, D.L.; Pierpont, C.G. Counterligand 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]
  31. Ackermann, J.; Meyer, F.; Kaifer, E.; Pritzkow, H. Tuning the activity of catechol oxidase model complexes by geometric changes of the dicopper core. Chem. Eur. J. 2002, 8, 247–258. [Google Scholar] [CrossRef]
  32. Börzel, H.; Comba, P.; Pritzkow, H. Structural studies on dicopper(II) compounds with catechol oxidase activity. Chem. Commun. 2001, 97–98. [Google Scholar] [CrossRef]
  33. Berreau, L.M.; Mahapatra, S.; Halfen, J.A.; Houser, R.P.; Young, V.G., Jr.; Tolman, W.B. Reactivity of peroxo- and bis(μ-oxo)dicopper complexes with catechols. Angew. Chem. Int. Ed. 1999, 38, 207–210. [Google Scholar] [CrossRef]
  34. Bencini, A.; Dei, A.; Sangregorio, C.; Totti, F.; Vaz, M.G.F. Crystal and molecular structure and magnetic exchange properties of bis(di-μ-ethoxo-bis(3,5-di-tert-butylsemiquinonato)dicopper(II)) complex. A synergy between DFT and experimental magnetochemistry. Inorg. Chem. 2003, 42, 8065–8071. [Google Scholar] [CrossRef] [PubMed]
  35. Ruf, M.; Noll, B.C.; Groner, M.D.; Yee, G.T.; Pierpont, C.G. Pocket semiquinonate complexes of cobalt(II), copper(II), and zinc(II) prepared with the hydrotris(cumenylmethylpyrazolyl)borate ligand. Inorg. Chem. 1997, 36, 4860–4865. [Google Scholar] [CrossRef] [PubMed]
  36. 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]
  37. Thompson, J.S.; Calabrese, J.C. Synthesis, spectroscopy, and structures of copper(II)-3,5-di-tert-butyl-o-semiquinone complexes. Inorg. Chem. 1985, 24, 3167–3171. [Google Scholar] [CrossRef]
  38. 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, 1349–1356. [Google Scholar] [CrossRef]
  39. Rheingold, A.L.; Haggerty, B.S.; Trofimenko, S. The first structurally characterized metal complex involving the free acid of a new tris(pyrazoly1)borate as ligand. Angew. Chem. Int. Ed. Engl. 1994, 33, 1983–1985. [Google Scholar] [CrossRef]
  40. Kime-Hunt, E.; Spartalian, K.; DeRusha, M.; Nunn, C.M.; Carrano, C.J. Synthesis, characterization, and molecular structures of a series of [(3,5-dimethylpyrazolyl)borato]vanadium(III) and -(IV) complexes. Inorg. Chem. 1989, 28, 4392–4399. [Google Scholar] [CrossRef]
  41. Singh, U.P.; Singh, R.; Hikichi, S.; Akita, M.; Moro-oka, Y. Characterization of a dinuclear Mn(II) tri(μ-carboxylato) complex with the hindered hydrotris(3,5-diisopropyl-1-pyrazolyl)borate (=TpiPr2) ligand: Intramolecular hydrogen bonding interaction between the protonated TpiPr2 and Mn-coordinating carboxylate ligands. Inorg. Chim. Acta 2000, 310, 273–278. [Google Scholar]
  42. Reinartz, S.; White, P.S.; Brookhart, M.; Templeton, J.L. Structural characterization of an intermediate in arene C–H bond activation and measurement of the barrier to C–H oxidative addition: A platinum(II) η2-benzene adduct. J. Am. Chem. Soc. 2001, 123, 12724–12725. [Google Scholar] [CrossRef]
  43. Norris, C.M.; Reinartz, S.; White, P.S.; Templeton, J.L. Barriers for arene C–H bond activation in platinum(II) η2-arene intermediates. Organometallics 2002, 21, 5649–5656. [Google Scholar] [CrossRef]
  44. Kostelansky, C.N.; MacDonald, M.G.; White, P.S.; Templeton, J.L. Stoichiometric alkane dehydrogenation with Tp’PtMe2H to form Tp’Pt(η2-olefin)(H) complexes. Organometallics 2006, 25, 2993–2998. [Google Scholar] [CrossRef]
  45. Reinartz, S.; White, P.S.; Brookhart, M.; Templeton, J.L. Five-coordinate trispyrazolylborate dihydridosilyl platinum(IV) complexes. J. Am. Chem. Soc. 2001, 123, 6425–6426. [Google Scholar] [CrossRef] [PubMed]
  46. Cuevas, A.; Geis, L.; Pintos, V.; Chiozzone, R.; Sanchíz, J.; Hummert, M.; Schumann, H.; Kremer, C. Synthesis, molecular structure and magnetic properties of a rhenium(IV) compound with catechol. J. Mol. Struct. 2009, 921, 80–84. [Google Scholar] [CrossRef]
  47. Horsman, G.P.; Jirasek, A.; Vaillancourt, F.H.; Barbosa, C.J.; Jarzecki, A.A.; Xu, C.; Mekmouche, Y.; Spiro, T.G.; Lipscomb, J.D.; Blades, M.W.; et al. Spectroscopic studies of the anaerobic enzyme-substrate complex of catechol 1,2-dioxygenase. J. Am. Chem. Soc. 2005, 127, 16882–16891. [Google Scholar] [CrossRef] [Green Version]
  48. Öhrström, L.; Michaud-Soret, I. Fe–catecholate and Fe–oxalate vibrations and isotopic substitution shifts from DFT quantum chemistry. J. Phys. Chem. A 1999, 103, 256–264. [Google Scholar] [CrossRef]
  49. Edwards, C.F.; Griffith, W.P.; White, A.J.P.; Williams, D.J. New catecholato oxorhenium(V) complexes. J. Chem. Soc. Dalton Trans. 1992, 957–962. [Google Scholar] [CrossRef]
  50. Michaud-Soret, I.; Andersson, K.K.; Que, L., Jr. Resonance Raman studies of catecholate and phenolate complexes of recombinant human tyrosine hydroxylase. Biochemistry 1995, 34, 5504–5510. [Google Scholar] [CrossRef]
  51. Solomon, E.I.; Szilagyi, R.K.; George, S.D.; Basumallick, L. Electronic structures of metal sites in proteins and models: Contributions to function in blue copper proteins. Chem. Rev. 2004, 104, 419–458. [Google Scholar] [CrossRef]
  52. Hathaway, B.J. A new look at the stereochemistry and electronic properties of complexes of the copper(II) ion. Struct. Bond. 1984, 57, 55–118. [Google Scholar]
  53. Hathaway, B.J.; Billing, D.E. The electronic properties and stereochemistry of mono-nuclear complexes of the copper(II) ion. Coord. Chem. Rev. 1970, 5, 143–207. [Google Scholar] [CrossRef]
  54. Armarego, W.L.F.; Chai, C.L.L. Purification of Laboratory Chemicals, 7th ed.; Butterworth-Heinemann: Oxford, UK, 2013. [Google Scholar]
  55. Kitajima, N.; Fujisawa, K.; Fujimoto, C.; Moro-oka, Y.; Hashimoto, S.; Kitagawa, T.; Toriumi, T.; Tatsumi, K.; Nakamura, A. A new model for dioxygen binding in hemocyanin. Synthesis, characterization, and molecular structure of the μ-η2:η2 peroxo dinuclear copper(II) complexes, [Cu(HB(3,5-R2pz)3)]2(O2) (R = i-Pr and Ph). J. Am. Chem. Soc. 1992, 114, 1277–1291. [Google Scholar] [CrossRef]
  56. CrystalClear: Data Collection and Processing Software; Rigaku Corporation: Tokyo, Japan, 2015.
  57. Altomare, A.; Cascarano, G.; Giacovazzo, C.; Guagliardi, A. Completion and refinement of crystal structures with SIR92. J. Appl. Cryst. 1993, 26, 343–350. [Google Scholar] [CrossRef]
  58. Altomare, A.; Burla, M.C.; Camalli, M.; Cascarano, G.L.; Giacovazzo, C.; Guagliardi, A.; Moliterni, A.G.G.; Polidori, G.; Spagna, R. SIR97: A new tool for crystal structure determination and refinement. J. Appl. Cryst. 1999, 32, 115–119. [Google Scholar] [CrossRef]
  59. Beurskens, P.T.; Beurskens, G.; de Gelder, R.; García-Granda, S.; Israel, R.; Gould, R.O.; Smits, J.M.M. The DIRDIF99 Program System, Technical Report of the Crystallography Laboratory; University of Nijmegen: Nijmegen, The Netherlands, 1999. [Google Scholar]
  60. Crystal Structure 4.3: Crystal Structure Analysis Package; Rigaku Corporation: Tokyo, Japan, 2018.
  61. Sheldrick, G.M. SHELXL Version 2014/3: A short history of SHELX. Acta Cryst. 2008, A64, 112–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Spek, A.L. PLATON SQUEEZE: A tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Cryst. 2015, 71, 9–18. [Google Scholar]
  63. Martins, L.M.D.R.S.; Pombeiro, A.J.L. Tris(pyrazol-1-yl)methane metal complexes for catalytic mild oxidative functionalizations of alkanes, alkenes and ketones. Coord. Chem. Rev. 2014, 265, 74–88. [Google Scholar] [CrossRef]
  64. Martins, L.M.D.R.S.; Pombeiro, A.J.L. Water-soluble C-scorpionate complexes—Catalytic and biological applications. Eur. J. Inorg. Chem. 2016, 15–16, 2236–2252. [Google Scholar] [CrossRef]
Scheme 1. Redox isomers of catechol.
Scheme 1. Redox isomers of catechol.
Inorganics 08 00037 sch001
Scheme 2. Ligands in this research.
Scheme 2. Ligands in this research.
Inorganics 08 00037 sch002
Scheme 3. Schematic drawing of two types of copper(II) complexes in Table 1.
Scheme 3. Schematic drawing of two types of copper(II) complexes in Table 1.
Inorganics 08 00037 sch003
Scheme 4. Schematic drawing of the synthetic schemes of complexes.
Scheme 4. Schematic drawing of the synthetic schemes of complexes.
Inorganics 08 00037 sch004
Figure 1. Crystal structures of [Cu(catCl4)(L1′)] (left) and [Cu(catBr4)(L1′)] (right) showing 50% displacement ellipsoids and the atom labeling scheme. Hydrogen atoms and solvents were omitted for reasons of clarity. Relevant bond lengths (Å) and angles (°): Cu1–N11, 2.036(2); Cu1–N21, 2.313(2); Cu1–N31, 2.044(2); Cu1–O40, 1.928(1); Cu1–O45, 1.911(2); C40–O40, 1.316(3); C45–O45, 1.318(2); C40–C45, 1.425(3); N11–Cu1–N21, 83.90(7); N11–Cu1–N31, 86.95(7); N21–Cu1–N31, 86.54(7); N11–Cu1–O40, 94.10(6); N21–Cu1–O40, 102.93(6); N31–Cu1–O40, 170.53(8); N11–Cu1–O45, 171.90(7); N21–Cu1–O45, 103.86(7); N31–Cu1–O45, 91.16(7); O40–Cu1–O45, 86.49(7). Cu1–N11, 2.023(3); Cu1–N21, 2.304(3); Cu1–N31, 2.053(2); Cu1–O40, 1.929(2); Cu1–O45, 1.905(3); C40–O40, 1.318(5); C45–O45, 1.318(4); C40–C45, 1.421(5); N11–Cu1–N21, 85.30(11); N11–Cu1–N31, 86.38(11); N21–Cu1–N31, 85.99(10); N11–Cu1–O40, 95.07(11); N21–Cu1–O40, 104.73(10); N31–Cu1–O40, 169.25(11); N11–Cu1–O45, 170.76(11); N21–Cu1–O45, 103.25(12); N31–Cu1–O45, 90.69(10); O40–Cu1–O45, 86.22(11).
Figure 1. Crystal structures of [Cu(catCl4)(L1′)] (left) and [Cu(catBr4)(L1′)] (right) showing 50% displacement ellipsoids and the atom labeling scheme. Hydrogen atoms and solvents were omitted for reasons of clarity. Relevant bond lengths (Å) and angles (°): Cu1–N11, 2.036(2); Cu1–N21, 2.313(2); Cu1–N31, 2.044(2); Cu1–O40, 1.928(1); Cu1–O45, 1.911(2); C40–O40, 1.316(3); C45–O45, 1.318(2); C40–C45, 1.425(3); N11–Cu1–N21, 83.90(7); N11–Cu1–N31, 86.95(7); N21–Cu1–N31, 86.54(7); N11–Cu1–O40, 94.10(6); N21–Cu1–O40, 102.93(6); N31–Cu1–O40, 170.53(8); N11–Cu1–O45, 171.90(7); N21–Cu1–O45, 103.86(7); N31–Cu1–O45, 91.16(7); O40–Cu1–O45, 86.49(7). Cu1–N11, 2.023(3); Cu1–N21, 2.304(3); Cu1–N31, 2.053(2); Cu1–O40, 1.929(2); Cu1–O45, 1.905(3); C40–O40, 1.318(5); C45–O45, 1.318(4); C40–C45, 1.421(5); N11–Cu1–N21, 85.30(11); N11–Cu1–N31, 86.38(11); N21–Cu1–N31, 85.99(10); N11–Cu1–O40, 95.07(11); N21–Cu1–O40, 104.73(10); N31–Cu1–O40, 169.25(11); N11–Cu1–O45, 170.76(11); N21–Cu1–O45, 103.25(12); N31–Cu1–O45, 90.69(10); O40–Cu1–O45, 86.22(11).
Inorganics 08 00037 g001
Figure 2. Crystal structure of [Cu(catCl4)(L1H)] showing 50% displacement ellipsoids and the atom labeling scheme (left: top view, right: side view). Hydrogen atoms were omitted for reasons of clarity. Relevant bond lengths (Å) and angles (°): Cu1–N11, 1.976(5); Cu1–N31, 1.973(5); Cu1–O40, 1.890(4); Cu1–O45, 1.952(4); C40–O40, 1.336(8); C45–O45, 1.334(6); C40–C45, 1.412(8); N11–Cu1–N31, 91.2(2); N11–Cu1–O40, 96.1(2); N31–Cu1–O40, 169.3(2); N11–Cu1–O45, 159.1(2); N31–Cu1–O45, 89.36(19); O40–Cu1–O45, 86.61(17).
Figure 2. Crystal structure of [Cu(catCl4)(L1H)] showing 50% displacement ellipsoids and the atom labeling scheme (left: top view, right: side view). Hydrogen atoms were omitted for reasons of clarity. Relevant bond lengths (Å) and angles (°): Cu1–N11, 1.976(5); Cu1–N31, 1.973(5); Cu1–O40, 1.890(4); Cu1–O45, 1.952(4); C40–O40, 1.336(8); C45–O45, 1.334(6); C40–C45, 1.412(8); N11–Cu1–N31, 91.2(2); N11–Cu1–O40, 96.1(2); N31–Cu1–O40, 169.3(2); N11–Cu1–O45, 159.1(2); N31–Cu1–O45, 89.36(19); O40–Cu1–O45, 86.61(17).
Inorganics 08 00037 g002
Figure 3. IR spectra of [Cu(catCl4)(L1′)] (top), [Cu(catBr4)(L1′)] (middle), and [Cu(catCl4)(L1H)] (bottom) measured by KBr pellets.
Figure 3. IR spectra of [Cu(catCl4)(L1′)] (top), [Cu(catBr4)(L1′)] (middle), and [Cu(catCl4)(L1H)] (bottom) measured by KBr pellets.
Inorganics 08 00037 g003
Figure 4. Far-IR spectra of [Cu(catCl4)(L1′)] (top), [Cu(catBr4)(L1′)] (middle), and [Cu(catCl4)(L1H)] (bottom) measured by CsI pellets.
Figure 4. Far-IR spectra of [Cu(catCl4)(L1′)] (top), [Cu(catBr4)(L1′)] (middle), and [Cu(catCl4)(L1H)] (bottom) measured by CsI pellets.
Inorganics 08 00037 g004
Figure 5. UV-Vis spectra of [Cu(catCl4)(L1′)] (green), [Cu(catBr4)(L1′)] (brown), and [Cu(catCl4)(L1H)] (blue) measured by CH2Cl2 solution at 0 °C.
Figure 5. UV-Vis spectra of [Cu(catCl4)(L1′)] (green), [Cu(catBr4)(L1′)] (brown), and [Cu(catCl4)(L1H)] (blue) measured by CH2Cl2 solution at 0 °C.
Inorganics 08 00037 g005
Figure 6. ESR spectra of [Cu(catCl4)(L1′)] (top), [Cu(catBr4)(L1′)] (middle), and [Cu(catCl4)(L1H)] (bottom) measured by CH2Cl2/C2H4Cl2 solution at 137 K.
Figure 6. ESR spectra of [Cu(catCl4)(L1′)] (top), [Cu(catBr4)(L1′)] (middle), and [Cu(catCl4)(L1H)] (bottom) measured by CH2Cl2/C2H4Cl2 solution at 137 K.
Inorganics 08 00037 g006
Table 1. Structural parameters of selected catecholato and o-semiquinonato copper(II) complexes.
Table 1. Structural parameters of selected catecholato and o-semiquinonato copper(II) complexes.
Complexd (Cu–O)
/ Å
d (C–O)
/ Å
d (C1–C2)
/ Å
Reference
catecholato complex
[Cu(catCl4)(L1′)]1.9280(14)
1.9106(18)
1.316(3)
1.318(2)
1.425(3)this work
[Cu(catBr4)(L1′)]1.929(2)
1.905(3)
1.318(5)
1.318(4)
1.421(5)this work
[Cu(catCl4)(L1H)]1.890(4)
1.952(4)
1.336(8)
1.334(6)
1.412(8)this work
[Cu(catBu2)(bipy)(MeOH)] a1.929(2)
1.898(2)
1.344(4)
1.342(4)
1.422(4)[26]
[Cu(catBu2)(DBED)] a1.943(2)
1.924(2)
1.347(3)
1.349(3)
1.425(4)[27]
[Cu(catBu2)(Me3-tacn)] a,b1.911(3)
1.903(3)
1.343(5)
1.346(4)
1.421(5)[28]
1.932(2)
1.895(3)
1.341(5)
1.353(4)
1.420(5)
[Cu(catBu2)(py)2](BF4) b1.964(2)
1.919(2)
1.354(4)
1.335(4)
1.423(6)[29]
1.964(2)
1.908(2)
1.367(3)
1.336(4)
1.408(4)
[Cu(catBu2)(bpy)] 1.901(5)
1.870(5)
1.364(8)
1.338(8)
1.407(10)[30]
[Cu(catCl4)Cu(H2O)2(μ-py1)](ClO4)2 a 1.933(2)
1.934(2)
1.339(4)
1.344(4)
1.412(5)[31]
[Cu(catCl4)Cu(H2O)(μ-py2)](ClO4)2 a,b 1.918(3)
1.979(3)
1.322(6)
1.337(6)
1.401(7)[31]
1.939(3)
1.984(3)
1.336(6)
1.333(5)
1.407(7)
[Cu(catCl4)Cu(H2O)(μ-py3)](ClO4)2 a 1.9684(17).1.9465(17)1.333(3)
1.328(3)
1.421(3)[31]
[Cu(catCl4)(bispidine1)] a2.456(2)
1.909(2)
1.301(3)
1.313(3)
1.440(3)[32]
[{Cu(catCl4)}2(μ-bispidine2)] a,b1.947(4)
1.899(4)
1.340(7)
1.336(7)
1.416(9)[32]
1.914(3)
1.930(5)
1.316(8)
1.324(6)
1.429(9)
[Cu(catCl4)(Bn3-tacn)] a1.940(4)
1.915(4)
1.335(7)
1.317(6)
1.430(7)[33]
o-semiquinonato complex
[Cu(sqBu2)(DPyA)(thf)2](BF4) a1.977(3)
1.978(2)
1.297(4)
1.276(5)
1.461(5)[26]
[Cu(sqBu2)(bipy)](BF4) a1.933(3)
1.936(3)
1.293(6)
1.287(5)
1.452(7)[26]
[Cu(sqBu2)(DBED)](SbF6) a1.975(2)
1.924(2)
1.285(3)
1.269(3)
1.464(4)[27]
[Cu(sqBu2)(TMCD)](SbF6) a1.963(2)
1.949(2)
1.289(3)
1.291(3)
1.455(4)[27]
[Cu(sqCl4)(Bn3-tacn)] a1.988(5)
2.001(4)
1.272(6)
1.265(7)
1.435(8)[33]
[Cu(sqBu2)(EtO)]2 b1.952(4)
1.949(5)
1.261(7)
1.294(7)
1.477(9)[34]
1.934(5)
1.945(5)
1.309(8)
1.275(9)
1.444(9)
[Cu(sqBu2)(TpCum,Me)] a1.952(3)
1.971(3)
1.279(5)
1.265(6)
1.457(7)[35]
[Cu(sqBu2)2]2 b1.918(4)
1.955(4)
1.291(7)
1.296(6)
1.470(8)[36]
1.944(4)
1.941(4)
1.290(7)
1.296(6)
1.45(1)
[Cu(sqBu2){NH(Py)2}](ClO4)2 a,b1.962(5)
1.964(4)
1.293(7)
1.304(7)
1.446(9)[37]
1.934(5)
1.969(4)
1.289(7)
1.284(4)
1.45(1)
a bipy = 2,2′-bipyridine, DBED = N,N’-di-tert-butylethane-1,2-diamine, Me3-tacn = 1,4,7-trimethyl-1,4,7-triazacyclononane), py1 = μ2-3,5-bis{N,N-dimethylaminoethyl(methyl)aminomethyl}pyrazolate, py2 = μ2-3,5-bis{bis(2-(diethylamino)ethyl)aminomethyl}pyrazolate, py3 = μ2-3,5-bis{N,N-dimethylaminoethyl(methyl)aminomethyl}pyrazolate, bispidine1 = 9,9-dihydroxy-1,5-bis(methoxycarbonyl)-3,7-dimethyl-2,4-bis(2-pyridyl)-3,7-diazabicyclo(3.3.1)nonane, bispidine2 = 7,7′-propano-bis(3-methyl-1,5-bis(methoxycarbonyl)-2,4-bis(2-pyridyl)-3,7-diazabicyclo(3.3.1)nonan-9-one, Bn3-tacn = 1,4,7-tribenzyl-1,4,7-triazacyclononane, DPyA = N-(pyridin-2-yl)pyridin-2-amine, TMCD = N,N,N’,N’-tetramethylcyclohexane-1,2-diamine, TpCum,Me = hydrotris(3-cumenyl-5-methylpyrazol-1-yl)borate, NH(Py)2 = di-2-pyridylamine, b two crystallographically independent molecules.
Table 2. Crystal data and structure refinement of copper complexes.
Table 2. Crystal data and structure refinement of copper complexes.
Complex[Cu(catCl4)(L1′)]·
2.5(CH3CN)
[Cu(catBr4)(L1′)]·
2.5(CH3CN)
[Cu(catCl4)(L1H)]
CCDC number620467620468620469
Empirical FormulaC39H53.5Cl4CuN8.5O2C39H53.5Br4CuN8.5O2C33H47BCl4CuN6O2
Formula Weight878.771056.57775.94
Crystal SystemMonoclinicMonoclinicMonoclinic
Space GroupC2/c (#15)C2/c (#15)P2/c (#13)
a19.7412(11)19.834(7)15.405(7)
b15.9152(8)16.302(5) 13.137(6)
c29.637(2)29.749(10)21.004(10)
β/°103.9780(9)107.918(3)101.852(6)
V39035.8(10)9152(5)4160(3)
Z884
Dcalc/g cm−31.2921.5331.239
μ(MoKα)/cm17.61640.2228.160
Temperature/°C−71−69−61
2θ Range/°6–556–556–55
Reflections Collected 30283 35994 33408
Unique Reflections10131 10344 9475
Rint0.03660.04100.0710
Number of Variables504504424
Reflections/Parameter Ratio20.1620.5222.35
Residuals: R1 (I > 2 σ (I))0.04760.05810.1099
Residuals: R (All Reflections)0.05190.07680.1477
Residuals: wR2 (All Reflections)0.11870.11180.2762
Goodness of Fit Indicator1.1391.1451.110
Max/Min Peak/e Å−30.32/−0.361.07/−0.681.05/−0.73
R1 = Σ ||Fo| − |Fc||/Σ |Fo|, wR2 = [Σ (w(Fo2 − Fc2)2)/Σ w(Fo2)2]1/2.

Share and Cite

MDPI and ACS Style

Fujisawa, K.; Ono, T.; Okamura, M. Synthesis and Characterization of Catecholato Copper(II) Complexes with Sterically Hindered Neutral and Anionic N3 Type Ligands: Tris(3,5-diisopropyl-1-pyrazolyl)methane and Hydrotris(3,5-diisopropyl-1-pyrazolyl)borate. Inorganics 2020, 8, 37. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8050037

AMA Style

Fujisawa K, Ono T, Okamura M. Synthesis and Characterization of Catecholato Copper(II) Complexes with Sterically Hindered Neutral and Anionic N3 Type Ligands: Tris(3,5-diisopropyl-1-pyrazolyl)methane and Hydrotris(3,5-diisopropyl-1-pyrazolyl)borate. Inorganics. 2020; 8(5):37. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8050037

Chicago/Turabian Style

Fujisawa, Kiyoshi, Tetsuya Ono, and Moemi Okamura. 2020. "Synthesis and Characterization of Catecholato Copper(II) Complexes with Sterically Hindered Neutral and Anionic N3 Type Ligands: Tris(3,5-diisopropyl-1-pyrazolyl)methane and Hydrotris(3,5-diisopropyl-1-pyrazolyl)borate" Inorganics 8, no. 5: 37. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8050037

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