Next Article in Journal
A Novel Ternary Catalyst PW4@MOF-808@SBA-15 for Deep Extraction Oxidation Desulfurization of Model Diesel
Previous Article in Journal
Microfiltration Membrane Pore Functionalization with Primary and Quaternary Amines for PFAS Remediation: Capture, Regeneration, and Reuse
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Research on Synthesis, Structure, and Catalytic Performance of Tetranuclear Copper(I) Clusters Supported by 2-Mercaptobenz-zole-Type Ligands

1
State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, University of Chinese Academy of Sciences, Fuzhou 350002, China
2
College of Chemistry and Materials Science, Fujian Normal University, Fuzhou 350007, China
3
Fujian College, University of Chinese Academy of Sciences, Fuzhou 350002, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(17), 4228; https://doi.org/10.3390/molecules29174228
Submission received: 5 August 2024 / Revised: 26 August 2024 / Accepted: 4 September 2024 / Published: 6 September 2024
(This article belongs to the Section Inorganic Chemistry)

Abstract

:
Tetrahedral copper(I) clusters [Cu4(MBIZ)4(PPh3)2] (2), [Cu4(MBOZ)4(PPh3)4] (6) (MBIZ = 2-mercaptobenzimidazole, MBOZ = 2-mercaptobenzoxazole) were prepared by regulation of the copper-thiolate clusters [Cu6(MBIZ)6] (1) and [Cu8(MBOZ)8I] (5) with PPh3. With the presence of iodide anion, the regulation provided the iodide-containing clusters [CuI4(MBIZ)3(PPh3)3I] (3) and [CuI4(MBOZ)3(PPh3)3I] (7). The cyclic voltammogram of 3 in MeCN (0.1 M nBu4NPF6, 298 K) at a scan rate of 100 mV s−1 shows two oxidation processes at Epa = +0.11 and +0.45 V with return waves observed at Epc = +0.25 V (vs. Fc+/Fc). Complex 3 has a higher capability to lose and gain electrons in the redox processes than complexes 2, 4, 4′, 6, and 7. Its thermal stability was confirmed by thermogravimetric analysis. The catalytic performance of 3 was demonstrated by the catalytic transformation of iodobenzenes to benzonitriles using AIBN as the cyanide source. The nitrile products show potential applications in the preparation of 1,3,5-triazine compounds for organic fluorescence materials.

Graphical Abstract

1. Introduction

Copper cluster compounds exhibit versatile properties and potential applications in the fields of catalysis [1,2,3,4,5,6], photoluminescent materials [7,8,9,10], sensing [11,12,13], biology and medicine [14,15,16,17], and electronics [18,19]. The tetranuclear copper(I) clusters have received great attention as a result of structural certainty for the convenience of study on the relationship between their configuration and performance. A number of clusters have been reported in the literature. They were mostly supported by small molecule ligands containing carbon [5,7,20], nitrogen [1,21,22], oxygen [23,24,25], halogen [6,8,9,10], phosphine [26,27,28], sulfur [2,3,4,29,30,31], and selenium [14,32,33,34] as coordination atoms. Among which, the [CuI4] clusters, such as copper-halide [8,9,10] and copper-selenolate [34] clusters, were usually studied for the sensing applications on account of their good performance in photoluminescence; copper-acetate clusters were shown to be potential precursors for the chemical vapor deposition of copper onto substrate materials in the study of the deep submicron integrated circuits [24]. The copper-acetylide [5], copper-alkylamide [1], and copper-thiolate clusters [2,3,4,29,30] were reported to show chemical catalytic reforming or have biological catalysis in enzymes or cells.
Tetranuclear [CuI4(SR)x] clusters show abundant structural diversity with various μx-SR bridging ligands, which play important roles in the configuration and stability of the skeleton of the clusters (Figure 1). The tetrahedron-type tetramers were usually formed by self-assembling of four Cu+ ions and six μ2-SR thiolates (or three dithiolates), with the Cu···Cu distances ranging from 2.6 to 2.9 Å [35,36,37,38,39]. The coordination of six μ2-SR thione ligands (e.g., thiourea) to four Cu+ ions with the support of four halide ions (Cl or I) results in the formation of the adamantane tetramers with a lengthening of Cu···Cu distances between 3.6 and 3.9 Å [40,41]. As electron-deficient ligands, thione ligands were also employed to produce hexagon tetramers [42,43,44,45,46]. Moreover, the μ2-SR bridging mode was applied in the construction of the quadrilateral-type [47,48,49,50,51] and the chair-type tetramers [52,53,54,55,56], in which the Cu+ ions were generally coordinated by electron-rich thiolates (e.g., BuS, AdS (Ad = adamantyl)) side by side. The wheel-type tetramers were obtained by bridging four [CuI(PR3)] units via tripodal [ArPS3]2− ligands [57,58,59]. The combination of the μ2-SR and μ3-SR coordination modes gave rise to the staircase tetramers with the assistance of four triphenylphosphines as the terminal ligands [60,61,62,63,64]. In contrast, the bonding of four thiolates to four Cu+ ions in the μ3-SR mode only afforded the cubane-type tetramers [4,30]. A type of pyramid [Cu4S] tetramers were synthesized by binding of four Cu+ ions with a μ4-sulfur bridge under the support of four chelate ligands at the base [65,66,67,68]. Though there have been a significant number of copper(I) tetramers reported in the literature, the study of tetranuclear [CuI] clusters with three or more types of ligands remains challenging and stimulating.
Aryl nitriles have wide applications not only in organic synthesis but also in materials sciences. The nitrile group can be easily transformed into a variety of functional groups, such as aldehydes, amines, amides, tetrazoles, triazoles, pyrimidines, triazine, etc. 1,3,5-triazine core was recognized as an electron acceptor that can make π-conjugated D-A molecules for application in organic electronic devices. The great significance has consistently stimulated the development of research methods for their preparation.
Encouraged by these reasons, we start to study the synthesis of tetranuclear copper(I)-thiolate clusters with nitrogen-containing mercaptan ligands (2-mercaptobenzimidazole (MBIZ), 2-mercaptobenzoxazole (MBOZ), and 2-mercaptobenzothiazole (MBTZ). The clusters are studied to catalyze the transformation of aryl iodides to aryl nitriles. The latters are used to produce octupolar π-conjugated molecules for the study of fluorescent materials.

2. Results and Discussions

We characterized all tetranuclear copper clusters by SC-XRD, XRD, and NMR (showed in Supplementary Materials). In air, the stirring of [Cu(CH3CN)4]BF4 with 2-mercaptobenzimidazole (MBIZ) in MeCN afforded a hexanuclear Cu(I) cluster [CuI6(MBIZ)6]·THF (1) (Scheme 1) [69]. The compound was regulated by the addition of PPh3 to generate a tetranuclear Cu(I) cluster [CuI4(MBIZ)4(PPh3)2]·3.5THF (2) with the Cu-S bond lengths and the Cu···Cu distances ranging from 2.26 to 2.85 Å and 2.68 to 4.26 Å, respectively. The regulation would lead to the formation of [CuI4(MBIZ)3(PPh3)3I]·3THF·CH3CN (3) with the presence of Et4NI [31,70,71,72,73,74]. The iodide anion is located on the top of the cluster, with the Cu-I bond lengths being between 2.78 and 3.12 Å. The Cu-S bond lengths and the Cu···Cu distances are presented as 2.250(1)/2.402(1) Å and 2.862(1)/4.069(1) Å under the condition of P-3 symmetry. Meanwhile, the reaction of [Cu(CH3CN)4]BF4 with MBIZ and PPh3 (or P(p-CH3-Ph)3) in the presence of HBF4 in MeCN provided the adamantane-type [CuI4(MBIZ)6(PPh3)4](BF4)4 (4) and [CuI4(MBIZ)6(P(p-CH3-Ph)3)4] (BF4)4 (4′) clusters, in which any two CuI ions are bridged by one MBIZ-thione molecule with the Cu-S bond lengths and the Cu···Cu distances varying from 2.35 to 2.37 Å (4′ = 2.35 to 2.42 Å) and 4.05 to 4.06 Å (4′ = 3.96 to 4.18 Å). In comparison, the reaction of 2-mercaptobenzoxazole (MBOZ) with CuI produced an octo-nuclear cylinder-type Cu(I) cluster [CuI8(MBOZ)8I](H3O)·CH3CN (5) [75]. An iodide anion is sited in the middle of this cylinder with the Cu-I bond lengths of 3.122(1) Å and 3.189(1) Å alternately. The cluster was regulated by PPh3 to form a tetranuclear Cu(I) cluster [CuI4(MBOZ)4(PPh3)4] (6) [61,76]. In comparison to complex 2, the top mercaptan ligand of complex 6 offers only the sulfur donor binding to the top three Cu+ ions, in accompany with an additional PPh3 binding to the bottom Cu+ ion. The change in the coordination mode leads to a lengthening of the Stop···Cubottom distances from 2.847(2) to 3.695(1) Å. With the addition of Et4NI, the regulation produced the iodide-containing cluster [CuI4(MBOZ)3(PPh3)3I]·CHCl3 (7) with a structure similar to complex 3. The crystal structures of complexes 2, 3, 4′, 6, and 7 are shown in Figure 2. Unexpectedly, the reaction of the analogous ligand 2-mercaptobenzothiazole (MBTZ) with either [Cu(CH3CN)4]BF4 or CuI only afforded the dimer complex 8. No corresponding cluster was obtained under the similar conditions.
The cyclic voltammogram of complex 3 in MeCN (0.1 M nBu4NPF6, 298 K) at a scan rate of 100 mV s−1 shows two oxidation processes at Epa = +0.11 and +0.45 V with return waves observed at Epc = +0.25 V (vs. Fc+/Fc) (Figure 3). It has inferior comparability with the fully irreversible process, with no return wave observed at Epa = +0.15 and +0.21 V for complex 7 under the same experimental conditions. This suggests a degree of decreased kinetic activity for the [CuI4] cluster with the ligand MBOZ. Without the support of iodide, the [CuI4] cluster 2 shows two electrochemical oxidation processes at Epa = −0.03 and +0.36 V without a reduction process. This contrasts with the oxidation process observed at Epa = −0.03 and +0.22 V for complex 6. As for complex 4 and 4′, their oxidation potentials are similar; they show two electrochemical oxidation processes at Epa = +0.36 and +0.49 V (Epa of 4′ = +0.35 and +0.48 V) without a reduction process. (The oxidation-reduction potentials of complex 2, 3, 4, 4′, 6, and 7 are shown in Table 1) It is not difficult to observe through a cyclic voltammogram that the oxidation and reduction peak area of complex 3 is larger than others, which means that when complex 3 reacts on the electrode surface, its electron transfer amount is maximized (Figure S11). And complex 3 is the only one that undergoes oxidation processes at both the ligand center and cluster core (Figures S7 and S10). The above result indicates that complex 3 has a higher capability to lose and gain electrons upon oxidation/reduction than complexes 2, 4, 4′, 6, and 7. The chemical redox stability of complex 3 was checked by electrochemical cycling tests (Figure S8).
The thermal stability of complex 3 was examined by the technique of thermogravimetric analysis (Figure 4). The first weight loss process occurs at 25–190 °C with a weight loss of about 6.38%. It is attributed to the escape of solvent molecules in the crystal sample. The second weight loss process occurs at 220–450 °C with a weight loss of about 48.62%. The significant weight loss is believed to result from the damage to the complex. According to the test results of differential scanning calorimetry, we can conclude that complex 3 did not undergo a significant phase transition below 180 °C (Figure S12). The result shows that complex 3 is stable for thermal catalytic reactions.
Research on organic catalysis by transition-metal clusters has attracted more and more interest, for example, the catalytic cyanation reaction. Excessive cyanide anion might be unfavorable for catalytic cyanation [77]. The cyanide source has always been a key factor. 2,2′-Azobis(2-methylpropionitrile) (AIBN) was considered as one of the low-toxicity cyanide sources for coupling reactions [78,79,80]. Thus, the stirring of iodobenzenes (0.2 mmol), AIBN (0.3 mmol), complex 3 (10 mol%), KI (0.6 mmol), and 1,8-diazabicyclo [5.4.0]undec-7-ene (DBU, 0.6 mmol) in CH3CN for 12 h at 120 °C provided the corresponding products 10 in normal yields of 40–56% (Figure 5). The low yield of 10l could be due to the variability of the bromide group on the substrate.
Aryl nitriles possess versatile utilities and are useful building blocks in the organic synthesis for organic optoelectronic materials. The trimerization of compound 10l using trifluoromethanesulfonic acid as catalyst provided 2,4,6-tris(5-bromothiophen-2-yl)-1,3,5-triazine (compound 11). The reaction of compound 11 with three equivalents of 4-octoxyphenylboronic acid afforded 2,4,6-tris(5-(4-(octyloxy)phenyl)thiophen-2-yl)-1,3,5-triazine (compound 12) in the presence of the catalyst Pd(PPh3)4 (Scheme 2). The CH2Cl2 solution of compound 12 (1.0 × 10−5 mol/L) emitted its characteristic fluorescence at 465 nm under the excitation at 391 nm (Figure 6), demonstrating its application potential in photo-electrochemistry.

3. Experimental Section

3.1. General

Chemicals. Unless otherwise stated, all inorganic reactions and manipulations are performed under air atmosphere. Complex [Cu(CH3CN)4]BF4, 2,4,6-tris(5-bromothiophen-2-yl)-1,3,5-triazine and (4-(octyloxy)phenyl)boronic acid were prepared according to the reported methods [81,82,83,84]. We characterized compounds by SC-XRD [85,86,87], XRD, and NMR; and some of these compounds were analyzed by CV (cyclic voltammetry) [88], TGA (thermogravimetric analysis), DSC (differential scanning calorimetry), elemental analysis, UV−vis spectra and fluorescence spectra (showed in Supplementary Materials).

3.2. Synthesis Procedures

3.2.1. Synthesis of Complex 2, 3, 4, 4′, 6, 7

The synthesis procedure of complex 1, 5, and 8 can be found in the Supporting Information.
[CuI4(MBIZ)4(PPh3)2]·3.5THF (2). The PPh3 (26.20 mg, 0.1 mmol) was added to a colorless solution of complex 1 (67.42 mg, 0.05 mmol) in CH3CN/THF (2/2 mL). The reaction mixture was stirred at 25 °C for 15 min. The solution was filtered and evaporated at room temperature, then the product was deposited as colorless crystals in three days (38.42 mg, 31.31%). 1H NMR (400 MHz, CD3CN) δ 9.88 (s, 4H), 7.42 (t, J = 8.7 Hz, 12H), 7.21–7.12 (m, 8H), 7.10–7.02 (m, 18H), 7.01–6.85 (m, 8H), 3.67 (s, 14H), 1.89–1.74 (m, 14H). 13C NMR (101 MHz, CD3CN) δ 134.29, 129.80, 128.67. Anal. Calcd. for C78H78Cu4N8O3.5P2S4: C, 57.27; H, 4.81; N, 6.85. Found: C, 57.75; H, 4.97; N, 6.73.
[CuI4(MBIZ)3(PPh3)3I]·3THF·CH3CN (3). The Et4NI (25.72 mg, 0.1 mmol) was added to a colorless solution of complex 1 (67.42 mg, 0.05 mmol) in CH3CN/THF (2/2 mL). The reaction mixture was stirred at 25 °C for 15 min to form a pale yellow solution. The resulting pale yellow solution was treated with PPh3 (26.20 mg, 0.1 mmol) and stirred for 15 min. The colorless solution was filtered and evaporated at room temperature, then the product was deposited as colorless crystals in three days (47.26 mg, 33.64%). 1H NMR (400 MHz, CDCl3) δ 8.23 (s, 3H), 7.07–6.32 (m, 57H). 13C NMR (101 MHz, CDCl3) δ 134.29, 129.80, 128.67. Anal. Calcd. for C89H87Cu4IN7O3P3S3: C, 57.08; H, 4.68; N, 5.24. Found: C, 57.85; H, 4.82; N, 5.31.
[CuI4(MBIZ)6(PPh3)4](BF4)4 (4). 2-mercaptobenzoxazole (MBIZ) (15.00 mg, 0.1 mmol) and HBF4 (13.17 mg, 0.15 mmol) were dissolved in THF (2 mL) by stirring for 15 min. The solution of [Cu(CH3CN)]BF4 (31.50 mg, 0.1 mmol) in MeCN (2 mL) was added to the solution of MBIZ dropwise. The mixture was stirred at 25 °C for 15 min. The resulting colorless solution was treated with PPh3 (26.20 mg, 0.1 mmol) and stirred for 15 min. The solution was filtered and evaporated at room temperature, then the product was deposited as colorless microcrystals in three days (28.52 mg, 44.00%). The microcrystals were dissolved in CH3CN + CHCl3 (4 mL) and sonicated for 30 min. The solution was filtered, and the filtrate was treated with Et2O (10 mL) to obtain some white precipitate as the product. 1H NMR (400 MHz, CD3CN/CDCl3) δ 11.69 (s, 12H), 7.40–7.26 (m, 60H), 7.25–7.12 (m, 24H). 13C NMR (101 MHz, CD3CN/CDCl3) δ 133.40, 130.10, 128.80. Anal. Calcd. for C114H96B4Cu4F16N12P4S6: C, 53.66; H, 3.79; N, 6.59. Found: C, 53.48; H, 3.84; N, 6.54.
[CuI4(MBIZ)6(P(p-CH3-Ph)3)4]·(BF4)4 (4′). 2-mercaptobenzoxazole (MBIZ) (15.00 mg, 0.1 mmol) and HBF4 (13.17 mg, 0.15 mmol) were dissolved in THF (2 mL) by stirring for 15 min. The solution of [Cu(CH3CN)]BF4 (31.50 mg, 0.1 mmol) in MeCN (2 mL) was added to the solution of MBIZ dropwise. The mixture was stirred at 25 °C for 15 min. The resulting colorless solution was treated with tri(p-tolyl)phosphine (30.40 mg, 0.1 mmol) and stirred for 15 min. The solution was filtered and evaporated at room temperature, then the product was deposited as colorless microcrystals in three days (30.36 mg, 43.71%). The microcrystals were dissolved in CH3CN + CHCl3 (4 mL) and sonicated for 30 min. The solution was filtered, and the filtrate was treated with Et2O (10 mL) to obtain some white precipitate as the product. 1H NMR (400 MHz, CDCl3) δ 11.07 (s, 12H), 7.08 (d, J = 15.4 Hz, 24H), 6.78–6.59 (m, 48H), 1.93 (s, 36H). 13C NMR (101 MHz, CDCl3) δ 156.61, 139.40, 132.04, 130.81, 129.09, 127.77, 123.83, 111.74, 21.24. Anal. Calcd. for C126H120B4Cu4F16N12P4S6: C, 55.64; H, 4.45; N, 6.18. Found: C, 55.37; H, 4.48; N, 6.14.
[CuI4(MBOZ)4(PPh3)4] (6). The PPh3 was added to the colorless solution of complex 5 (94.84 mg, 0.05 mmol) in CH3CN/THF (2/2 mL) (26.20 mg, 0.1 mmol). The reaction mixture was stirred at 25 °C for 15 min. The solution was filtered and evaporated at room temperature, then the product was deposited as colorless crystals in three days (46.15 mg, 32.31%). 1H NMR (400 MHz, CDCl3) δ 7.39 (t, J = 8.6 Hz, 24H), 7.20–7.02 (m, 36H), 6.97–6.55 (m, 16H). 13C NMR (101 MHz, CDCl3) δ 141.36, 133.88, 129.25, 128.33. Anal. Calcd. for C100H76Cu4N4O4P4S4: C, 63.08; H, 4.02; N, 2.94. Found: C, 62.79; H, 4.09; N, 2.91.
[CuI4(MBOZ)3(PPh3)3I]·CHCl3 (7). The Et4NI (25.72 mg, 0.1 mmol) was added to the colorless solution of complex 5 (94.84 mg, 0.05 mmol) in CH3CN/THF (2/2 mL). The reaction mixture was stirred at 25 °C for 15 min to form a pale yellow solution. The resulting pale yellow solution was treated with PPh3 (26.20 mg, 0.1 mmol) and stirred for 15 min. The colorless solution was filtered and evaporated at room temperature, then the product was deposited as colorless microcrystals in three days (52.32 mg, 40.13%). The microcrystals were recrystallized from CH3CN/CHCl3 to afford the product as some colorless crystals. 1H NMR (400 MHz, CDCl3) δ 7.60–7.35 (m, 18H), 7.06 (s, 27H), 6.84 (d, J = 4.7 Hz, 6H), 6.54–6.34 (m, 6H). 13C NMR (101 MHz, CDCl3) δ 151.02, 133.92, 129.35, 128.28, 122.65, 122.34, 116.94, 109.51. Anal. Calcd. for C76H58Cl3Cu4IN3O3P3S3: C, 52.53; H, 3.36; N, 2.42. Found: C, 53.05; H, 3.43; N, 2.44.

3.2.2. General Procedure for the Synthesis of Compound 10

Aromatic iodides 9 (0.20 mmol), Azodiisobutyronitrile (AIBN, 0.30 mmol, 49.26 mg), complex 3 (10% mmol, 32.00 mg), potassium iodide (0.60 mmol, 99.60 mg), 1,8-Diazabicyclo [5.4.0]undec-7-ene (DBU, 0.60 mmol, 91.34 mg), and CH3CN (2 mL) were mixed in a 50 mL Teflon screw-cap sealed tube. The mixture was vigorously stirred at 120 °C under N2 atmosphere for 12 h (oil bath). After cooling to room temperature, the reaction mixture was diluted with dichloromethane (20 mL), filtered through a pad of silica gel, and concentrated under reduced pressure. The crude product was purified on a silica gel column eluted with petroleum and ether/ethyl acetate (8:1 v/v) to yield the products 10. NMR spectra are presented in pages S16–S27 (SI).
4-Ethoxybenzonitrile (10a): yield, 43% (11.10 mg); white solid. M.p. 62–63 °C. 1H NMR (400 MHz, CDCl3) δ 7.56 (d, J = 8.8 Hz, 2H), 6.92 (d, J = 8.8 Hz, 2H), 4.06 (q, J = 7.0 Hz, 2H), 1.43 (t, J = 7.0 Hz, 3H). 13C NMR (101 MHz, CDCl3) δ 162.34, 134.04, 119.41, 115.23, 103.72, 64.01, 14.65.
4-Methoxybenzonitrile (10b): yield, 40% (10.63 mg); pale yellow solid. M.p. 59–60 °C. 1H NMR (400 MHz, CDCl3) δ 7.58 (d, J = 8.4 Hz, 2H), 6.94 (d, J = 8.5 Hz, 2H), 3.85 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 162.94, 134.09, 119.36, 114.85, 104.04, 55.66.
4-tert-Butylbenzonitrile (10c): yield, 44% (12.77 mg); colorless liquid; 1H NMR (400 MHz, CDCl3) δ 7.58 (d, J = 8.4 Hz, 2H), 7.47 (d, J = 8.4 Hz, 2H), 1.32 (s, 9H). 13C NMR (101 MHz, CDCl3) δ 156.73, 132.05, 126.26, 119.26, 109.37, 35.35, 31.03.
4-Isopropylbenzonitrile (10d): yield, 40% (10.66 mg); colorless liquid; 1H NMR (400 MHz, CDCl3) δ 7.57 (d, J = 8.0 Hz, 2H), 7.31 (d, J = 8.1 Hz, 2H), 3.01–2.91 (m, 1H), 1.25 (d, J = 6.8 Hz, 6H). 13C NMR (101 MHz, CDCl3) δ 154.48, 132.35, 127.42, 119.31, 109.69, 34.50, 23.66.
4-Methylbenzonitrile (10e): yield, 42% (9.80 mg); colorless liquid; 1H NMR (400 MHz, CDCl3) δ 7.52 (d, J = 8.1 Hz, 2H), 7.26 (d, J = 8.0 Hz, 2H), 2.41 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 143.72, 131.98, 129.83, 119.14, 109.22, 21.80.
p-Phenylbenzonitrile (10f): yield, 48% (17.20 mg); white solid; 1H NMR (400 MHz, CDCl3) δ 7.71 (q, J = 8.3 Hz, 4H), 7.59 (d, J = 7.6 Hz, 2H), 7.49 (t, J = 7.5 Hz, 2H), 7.43 (t, J = 7.2 Hz, 1H).13C NMR (101 MHz, CDCl3) δ 145.77, 139.27, 132.71, 129.23, 128.78, 127.84, 127.34, 119.07, 111.00.
4-Cyanobenzoicacidmethylester (10g): yield, 48% (14.40 mg); white solid; 1H NMR (400 MHz, CDCl3) δ 8.12 (d, J = 8.3 Hz, 2H), 7.73 (d, J = 8.3 Hz, 2H), 3.95 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 165.52, 134.00, 132.32, 130.19, 118.06, 116.47, 52.83.
1,4-Benzodinitrile (10h): yield, 54% (13.80 mg); white solid; 1H NMR (400 MHz, CDCl3) δ 7.80 (s, 4H). 13C NMR (101 MHz, CDCl3) δ 132.93, 117.14, 116.84.
2-Cyanonaphthylene (10i): yield, 50% (15.30 mg); white solid; 1H NMR (400 MHz, CDCl3) δ 8.22 (s, 1H), 7.90 (t, J = 8.7 Hz, 3H), 7.68–7.56 (m, 3H). 13C NMR (101 MHz, CDCl3) δ 134.74, 134.26, 132.33, 129.30, 129.15, 128.52, 128.16, 127.76, 126.44, 119.37, 109.45.
1-Cyanonaphthalene (10j): yield, 56% (17.20 mg); white solid; 1H NMR (400 MHz, CDCl3) δ 8.24 (d, J = 8.3 Hz, 1H), 8.08 (d, J = 8.3 Hz, 1H), 7.92 (t, J = 6.6 Hz, 2H), 7.69 (t, J = 7.6 Hz, 1H), 7.62 (t, J = 7.5 Hz, 1H), 7.52 (t, J = 7.7 Hz, 1H). 13C NMR (101 MHz, CDCl3) δ 133.40, 133.02, 132.75, 132.45, 128.77, 128.71, 127.66, 125.25, 125.04, 117.95, 110.27.
Quinoline-6-carbonitrile (10k): yield, 55% (16.90 mg); white solid; 1H NMR (400 MHz, CDCl3) δ 9.05 (s, 1H), 8.20 (dd, J = 13.6, 8.3 Hz, 3H), 7.84 (d, J = 8.7 Hz, 1H), 7.54 (dd, J = 8.2, 4.1 Hz, 1H). 13C NMR (101 MHz, CDCl3) δ 153.32, 149.17, 136.53, 134.23, 131.14, 130.25, 127.66, 122.85, 118.59, 110.50.
5-Bromothiophene-2-acetonitrile (10l): yield, 20% (7.52 mg); yellow liquid; 1H NMR (400 MHz, CDCl3) δ 7.37 (d, J = 4.0 Hz, 1H), 7.08 (d, J = 4.0 Hz, 1H). 13C NMR (101 MHz, CDCl3) δ 138.01, 130.70, 120.08, 113.06, 111.17.

3.2.3. Experimental Procedures for the Synthesis of Compounds 11 and 12

2,4,6-tris(5-bromothiophen-2-yl)-1,3,5-triazine (11). Trifluoromethanesulfonic acid was added to a stirred solution of 10l (7 mmol, 1.32 g) in dry CHCl3 (20 mL) (14 mmol, 2.10 g) at 0 °C. The mixture was stirred for 36 h at room temperature. After removal of the solvent under reduced pressure, the residue was neutralized with an aqueous NaHCO3. The formed precipitate was collected by filtration, washed with water, methanol, acetone, and hexane in this order, and then dried under vacuum to afford 11 as an off-white solid (yield = 2.56 g, 65%). 1H NMR (400 MHz, CDCl3) δ 7.97 (d, J = 4.0 Hz, 1H), 7.17 (d, J = 4.0 Hz, 1H). 13C NMR (101 MHz, CDCl3) δ 166.86, 142.30, 132.17, 131.81, 120.85.
2,4,6-tris(5-(4-(octyloxy)phenyl)thiophen-2-yl)-1,3,5-triazine (12). The Pd(PPh3)4 (0.05 mmol, 0.06 g) and aqueous K2CO3 (2.0 M, 8 mL; N2 bubbled before use) were added to a solution of (4-(octyloxy)phenyl)boronic acid (2 mmol, 0.5 g) and 2,4,6-tris(5-bromothiophen-2-yl)-1,3,5-triazine (0.5 mmol, 0.28 g) in dry THF (20 mL). The mixture was vigorously stirred for 20 h at 60 °C under N2 atmosphere. The mixture was cooled and poured into a separatory funnel together with water. The aqueous solution was extracted with CH2Cl2 (3 × 50 mL), and the combined organic layers were dried over anhydrous Na2SO4 overnight. The solvent was removed and the crude product was purified on a silica gel column eluted with petroleum and CH2Cl2 (2:1 v/v) to afford 2,4,6-tris(5-(4-(octyloxy)phenyl)thiophen-2-yl)-1,3,5-triazine as a light yellow solid in yield 54% (0.25 g). 1H NMR (400 MHz, CDCl3) δ 8.04 (d, J = 3.8 Hz, 1H), 7.52 (d, J = 8.6 Hz, 2H), 7.15 (t, 1H), 6.81 (d, J = 8.7 Hz, 2H), 3.87 (t, J = 6.5 Hz, 2H), 1.81–1.61 (m, 2H), 1.42–1.29 (m, 2H), 1.29–1.08 (m, 8H), 0.80 (t, J = 6.6 Hz, 3H). 13C NMR (101 MHz, CDCl3) δ 167.24, 159.68, 151.17, 139.28, 132.75, 127.48, 126.66, 123.26, 114.99, 68.27, 31.98, 29.55, 29.41, 26.20, 22.82, 14.28.

4. Conclusions

In summary, we described the synthesis of a series of tetrahedral copper-thiolate tetramers by the regulation of multi-nuclear copper clusters with PPh3. The insertion of iodide anion to the cluster improved the redox activity and stability, which was supported by the study of electrochemistry and thermogravimetric analysis. The reaction performance was demonstrated by the catalytic cyanation of iodobenzenes using AIBN as the cyanide source. The products show potential applications as luminescent layer materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29174228/s1. Accession codes: CCDC 2336741, CCDC 2336735, CCDC 2374006, CCDC 2374007, CCDC 2374008, and CCDC 2336717 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif (accessed on 3 September 2024), or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: +44-1223-336033.

Author Contributions

Conceptualization, T.Z. and X.Z.; methodology, T.Z.; software, W.F.; validation, T.Z. and W.F.; formal analysis, X.Z.; investigation W.Z.; resources, X.Z.; data curation, T.Z. and X.Z.; writing-original draft preparation, T.Z.; writing-review and editing, X.Z.; visualization, W.Z.; supervision, X.Z.; project administration, X.Z.; funding acquisition, X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Fujian Province (grant no. 2020J01114).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data underlying this study are available in the published article and its online Supplementary Material.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sung, S.; Braddock, D.C.; Armstrong, A.; Brennan, C.; Sale, D.; White, A.J.P.; Davies, R.P. Synthesis, characterisation and reactivity of copper(I) amide complexes and studies on their role in the modified Ullmann amination reaction. Chem. Eur. J. 2015, 21, 7179–7192. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, M.-J.; Li, H.-X.; Lia, H.-Y.; Lang, J.-P. Copper(I) 5-phenylpyrimidine-2-thiolate complexes showing unique optical properties and high visible light-directed catalytic performance. Dalton Trans. 2016, 45, 17759–17769. [Google Scholar] [CrossRef] [PubMed]
  3. Johnston, E.M.; Carreira, C.; Dell’Acqua, S.; Dey, S.G.; Pauleta, S.R.; Moura, I.; Solomon, E.I. Spectroscopic definition of the CuZ° intermediate in turnover of nitrous oxide reductase and molecular insight into the catalytic mechanism. J. Am. Chem. Soc. 2017, 139, 4462–4476. [Google Scholar] [CrossRef]
  4. Nayak, M.; Joshi, D.K.; Kumar, K.; Singh, A.S.; Tiwari, V.K.; Bhattacharya, S. Synthesis of a series of a few hydrosulfide complexes of Cu(I). A μ3-SH-Bridged rare cubane-like tetramer showing efficient catalytic activity toward azide-alkyne cycloaddition. Inorg. Chem. 2021, 60, 8075–8084. [Google Scholar] [CrossRef]
  5. Liu, Y.; Blanchard, V.; Danoun, G.; Zhang, Z.; Tlili, A.; Zhang, W.; Monnier, F.; Lee, A.V.D.; Mao, J.; Taillefer, M. Copper-catalyzed sonogashira reaction in water. ChemistrySelect 2017, 2, 11599–11602. [Google Scholar] [CrossRef]
  6. Yuan, Q.; Huang, Z.; Wu, W.; Hong, D.; Zhu, S.; Wei, J.; Zhou, S.; Wang, S. Halide-bridged tetranuclear organocopper(I) clusters supported by indolyl-based NCN pincer ligands and their catalytic activities towards the hydrophosphination of alkenes. Dalton Trans. 2022, 51, 17795–17803. [Google Scholar] [CrossRef] [PubMed]
  7. Doshi, A.; Sundararaman, A.; Venkatasubbaiah, K.; Zakharov, L.N.; Rheingold, A.L.; Myahkostupov, M.; Piotrowiak, P.; Jakle, F. Pentafluorophenyl copper-pyridine complexes: Synthesis, supramolecular structures via cuprophilic and π-Stacking interactions, and solid-state luminescence. Organometallics 2012, 31, 1546–1558. [Google Scholar]
  8. Galimova, M.F.; Zueva, E.M.; Dobrynin, A.B.; Kolesnikov, I.E.; Musin, R.R.; Musina, E.I.; Karasik, A.A. Luminescent Cu4I4-cubane clusters based on N-methyl-5,10-dihydrophenarsazines. Dalton Trans. 2021, 50, 13421–13429. [Google Scholar] [CrossRef] [PubMed]
  9. Enikeeva, K.R.; Shamsieva, A.V.; Strelnik, A.G.; Fayzullin, R.R.; Zakharychev, D.V.; Kolesnikov, I.E.; Dayanova, I.R.; Gerasimova, T.P.; Strelnik, I.D.; Musina, E.I.; et al. Green emissive copper(I) coordination polymer supported by the diethylpyridylphosphine ligand as a luminescent sensor for overheating processes. Molecules 2023, 28, 706. [Google Scholar] [CrossRef]
  10. Dong, C.; Song, X.; Hasanov, B.E.; Yuan, Y.; Gutiérrez-Arzaluz, L.; Yuan, P.; Nematulloev, S.; Bayindir, M.; Mohammed, O.F.; Bakr, O.M. Organic−Inorganic hybrid glasses of atomically precise nanoclusters. J. Am. Chem. Soc. 2024, 146, 7373–7385. [Google Scholar] [CrossRef]
  11. Wang, B.; Gui, R.; Jin, H.; He, W.; Wang, Z. Red-emitting BSA-stabilized copper nanoclusters acted as a sensitive probe for fluorescence sensing and visual imaging detection of rutin. Talanta. 2018, 178, 1006–1010. [Google Scholar] [CrossRef] [PubMed]
  12. An, Y.; Ren, Y.; Bick, M.; Dudek, A.; Waworuntuc, E.H.-W.; Tang, J.; Chen, J.; Chang, B. Highly fluorescent copper nanoclusters for sensing and bioimaging. Biosens. Bioelectron. 2020, 154, 112078. [Google Scholar] [CrossRef]
  13. Shao, C.; Li, C.; Zhang, C.; Ni, Z.; Liu, X.; Wang, Y. Novel synthesis of orange-red emitting copper nanoclusters stabilized by methionine as a fluorescent probe for norfloxacin sensing. Spectrochim. Acta Part A 2020, 236, 118334. [Google Scholar] [CrossRef] [PubMed]
  14. Dias, R.d.S.; Cervo, R.; Siqueira, F.d.S.; Campos, M.M.A.d.; Lang, E.S.; Tirloni, B.; Schumacher, R.F.; Cargnelutti, R. Synthesis and antimicrobial evaluation of coordination compounds containing 2,20-bis(3-aminopyridyl) diselenide as ligand. Appl. Organomet. Chem. 2020, 34, e5750. [Google Scholar] [CrossRef]
  15. Ramaduraia, M.; Rajendranb, G.; Bamaa, T.S.; Prabhua, P.; Kathiravan, K. Biocompatible thiolate protected copper nanoclusters for an efficient imaging of lung cancer cells. J. Photochem. Photobiol. B 2020, 205, 111845. [Google Scholar] [CrossRef]
  16. Gourdon-Grünewaldt, L.; Blacque, O.; Gasser, G.; Cariou, K. Towards copper(I) clusters for photo-induced oxidation of biological thiols in living cells. ChemBioChem 2023, 24, e202300496. [Google Scholar] [CrossRef] [PubMed]
  17. Murugan, R.; Gunasekar, K.R.; Ali, S.; Sivalingam, C.; Subramanian, R.; Venkatesa, K.; Selvaraj, M.; Malakondaiah, S. Tenacious copper nanoclusters shielded by aromatic thiol for use in antibacterial, bioimaging, and dye degradation applications. J. Ind. Eng. Chem. 2024, 137, 448–454. [Google Scholar] [CrossRef]
  18. Liu, M.; Liang, J.; Liu, Z. Modulating the ferroelectric performance by altering halogen anions in the crystals of tetranuclear copper-clusters. New J. Chem. 2021, 45, 12091–12096. [Google Scholar] [CrossRef]
  19. Oh, J.; Yang, S.Y.; Kim, S.; Lee, C.; Cha, J.-H.; Jang, B.C.; Im, S.G.; Choi, S.-Y. Imidazole-based artificial synapses for neuromorphic computing: A cluster-type conductive filament via controllable nanocluster nucleation. Mater. Horiz. 2023, 10, 2035–2046. [Google Scholar] [CrossRef]
  20. Körte, L.A.; Schwabedissen, J.; Soffner, M.; Blomeyer, S.; Reuter, C.G.; Vishnevskiy, Y.V.; Neumann, B.; Stammler, H.-G.; Mitzel, N.W. Tris(perfluorotolyl)borane—A boron lewis superacid. Angew. Chem. Int. Ed. 2017, 56, 8578–8582. [Google Scholar] [CrossRef]
  21. Robinson, T.P.; Price, R.D.; Davidson, M.G.; Foxb, M.A.; Johnson, A.L. Why are the {Cu4N4} rings in copper(I) phosphinimide clusters [Cu{μ-N=PR3}]4 (R = NMe3 or Ph) planar? Dalton Trans. 2015, 44, 5611–5619. [Google Scholar] [CrossRef]
  22. Quintana, L.M.A.; Jiang, N.; Bacsa, J.; Pierre, H.S.L. Coinage metal tris(dialkylamido)imidophosphorane complexes as transmetallation reagents for cerium complexes. Dalton Trans. 2020, 49, 5420–5423. [Google Scholar] [CrossRef] [PubMed]
  23. Oguadinma, P.O.; Schaper, F. Bis(2-phenylethyl)-nacnac: A chiral diketiminate ligand and its copper complexes. Organometallics 2009, 28, 4089–4097. [Google Scholar] [CrossRef]
  24. Mothes, R.; Ruffer, T.; Shen, Y.; Jakob, A.; Walfort, B.; Petzold, H.; Schulz, S.E.; Ecke, R.; Gessnerb, T.; Lang, H. Phosphite copper(I) trifluoroacetates [((RO)3P)mCuO2CCF3] (m = 1, 2, 3): Synthesis, solid state structures and their potential use as CVD precursors. Dalton Trans. 2010, 39, 11235–11247. [Google Scholar] [CrossRef]
  25. Bellow, J.A.; Yousif, M.; Fang, D.; Kratz, E.G.; Cisneros, G.A.; Groysman, S. Synthesis and reactions of 3d metal complexes with the bulky alkoxide ligand [OCtBu2Ph]. Inorg. Chem. 2015, 54, 5624–5633. [Google Scholar] [CrossRef]
  26. Wolf, R.; Hey-Hawkins, E. Synthesis and molecular structure of the Cu4P8 cage compound [Cu4(P4Ph4)2(PCyp3)3]. Angew. Chem. Int. Ed. 2005, 44, 6241–6244. [Google Scholar] [CrossRef] [PubMed]
  27. Harford, P.J.; Haywood, J.; Smith, M.R.; Bhawal, B.N.; Raithby, P.R.; Uchiyamac, M.; Wheatley, A.E.H. Expanding the tools available for direct ortho cupration-targeting lithium phosphidocuprates. Dalton Trans. 2012, 41, 6148–6154. [Google Scholar] [CrossRef]
  28. Rathnayaka, S.C.; Lindeman, S.V.; Mankad, N.P. Multinuclear Cu(I) clusters featuring a new triply bridging coordination mode of phosphaamidinate ligands. Inorg. Chem. 2018, 57, 9439–9445. [Google Scholar] [CrossRef]
  29. Baslé, A.; Platsaki, S.; Dennison, C. Visualizing Biological Copper Storage: The importance of thiolate Coordinated tetranuclear clusters. Angew. Chem. 2017, 129, 8823–8826. [Google Scholar] [CrossRef]
  30. Trivedi, M.; Kumar, A.; Husain, A.; Rath, N.P. Copper(I) complexes containing PCP ligand catalyzed hydrogenation of carbon dioxide to formate under ambient conditions. Inorg. Chem. 2021, 60, 4385–4396. [Google Scholar] [CrossRef]
  31. Liang, X.-Q.; Gupta, R.K.; Li, Y.-W.; Ma, H.-Y.; Gao, L.-N.; Tung, C.-H.; Sun, D. Structural diversity of copper(I) cluster-based coordination polymers with pyrazine-2-thiol ligand. Inorg. Chem. 2020, 59, 2680–2688. [Google Scholar] [CrossRef] [PubMed]
  32. Dhayal, R.S.; Liao, J.-H.; Hou, H.-N.; Ervilita, R.; Liao, P.-K.; Liua, C.W. Copper(I) diselenocarbamate clusters: Synthesis, structures and single-source precursors for Cu and Se composite materials. Dalton Trans. 2015, 44, 5898–5908. [Google Scholar] [CrossRef]
  33. Hanau, K.; Rinn, N.; Dehnen, S. Variations in the interplay of intermetallic and metal chalcogenide units in organotin−copper selenide clusters. Inorg. Chem. 2020, 59, 198–202. [Google Scholar] [CrossRef] [PubMed]
  34. Karmakar, G.; Tyagi, A.; Wadawale, A.P.; Shah, A.Y.; Kedarnath, G.; Srivastava, A.P.; Singh, V. Accessing photoresponsive copper selenide nanomaterials and thin films through tetranuclear Cu(I) pyridylselenolate cluster. J. Mater. Sci. 2020, 55, 15439–15453. [Google Scholar] [CrossRef]
  35. Liu, H.; Bandeira, N.A.G.; Calhorda, M.J.; Drew, M.G.B.; Felix, V.; Novosad, J.; de Biani, F.F.; Zanello, P. Cu(I) and Ag(I) complexes of chalcogenide derivatives of the organometallic ligand dppf and the dppa analogue. J. Organomet. Chem. 2004, 689, 2808–2819. [Google Scholar] [CrossRef]
  36. Belo, D.; Figueira, M.J.; Mendonca, J.; Santos, I.C.; Almeida, M.; Henriques, R.T.; Duarte, M.T.; Rovira, C.; Veciana, J. Copper, cobalt and platinum complexes with dithiothiophene-based ligands. Eur. J. Inorg. Chem. 2005, 2005, 3337–3345. [Google Scholar] [CrossRef]
  37. Chen, C.; Weng, Z.; Hartwig, J.F. Synthesis of Copper(I) Thiolate Complexes in the thioetherification of aryl halides. Organometallics 2012, 31, 8031–8037. [Google Scholar] [CrossRef]
  38. Lane, A.C.; Vollmer, M.V.; Laber, C.H.; Melgarejo, D.Y.; Chiarella, G.M.; Fackler, J.P., Jr.; Yang, X.; Baker, G.A.; Walensky, J.R. Multinuclear Copper(I) and silver(I) amidinate complexes: Synthesis, luminescence, and CS2 insertion reactivity. Inorg. Chem. 2014, 53, 11357–11366. [Google Scholar] [CrossRef]
  39. Royappaa, A.T.; Papoular, R.J.; Gembicky, M.; Shepard, W.; Ross, A.D.; Stemen, A.G.; Bobbitt, J.J.; Doan, D.T.; Lapidus, S.H.; Johnston, D.H.; et al. Crystalline phase transitions and water-soluble complexes of copper(I) 2-hydroxyethanethiolate. Polyhedron 2022, 222, 115873. [Google Scholar] [CrossRef]
  40. Griffith, E.H.; Hunt, G.W.; Amma, E.L. The adamantane structure in polynuclear Cu4S6 cores: The crystal and molecular structures of Cu4[SC(NH2)2]6(NO3)4·4H2O and Cu4[SC(NH2)2]9(NO3)4·4H2O. J. Chem. Soc. Chem. Commun. 1976, 432–433. [Google Scholar] [CrossRef]
  41. Bowmaker, G.A.; Hanna, J.V.; Pakawatchai, C.; Skelton, B.W.; Thanyasirikul, Y.; White, A.H. Crystal structures and vibrational spectroscopy of copper(I) thiourea complexes. Inorg. Chem. 2009, 48, 350–368. [Google Scholar] [CrossRef]
  42. Raper, E.S.; Creighton, J.R.; Wilson, J.D.; Clegg, W.; Milne, A. Preparation, spectroscopy and crystal structure of a novel tetranuclear copper(I) benzimidazoline-2-thione complex: Cyclo-hexakis-μ2-(benzimidazoline-2-thione) tetrakis[benzimidazoline-2-thione copper(I)] perchlorate quatrodeca hydrate. Inorganica Chim. Acta 1988, 149, 265–271. [Google Scholar] [CrossRef]
  43. Christlieb, M.; Cowley, A.R.; Dilworth, J.R.; Donnelly, P.S.; Paterson, B.M.; Struthers, H.S.R.; White, J.M. New bimetallic compounds based on the bis(thiosemicarbazonato) motif. Dalton Trans. 2007, 327–331. [Google Scholar] [CrossRef] [PubMed]
  44. Bianketti, S.; Blake, A.J.; Wilson, C.; Hubberstey, P.; Champness, N.R.; Schroder, M. Second-sphere hydrogen-bonding in heteroditopic mercaptopyridinium copper(I) frameworks. CrystEngComm 2009, 11, 763–769. [Google Scholar] [CrossRef]
  45. Singh, S.; Bhattacharya, S. Studies of structural diversity due to inter-/intra-molecular hydrogen bonding and photoluminescent properties in thiocarboxylate Cu(i) and Ag(i) complexes. RSC Adv. 2014, 4, 49491–49500. [Google Scholar] [CrossRef]
  46. Limberg, C.; Frank, N.; Dallmann, A.; Braun-Cula, B.; Herwig, C. Mercaptothiacalixarenes steer 24 copper(I) centers to form a Hollow-Sphere structure featuring Cu2S2 motifs with exceptionally short Cu···Cu distances. Angew. Chem. Int. Ed. 2020, 59, 6735–6739. [Google Scholar]
  47. Pichota, A.; Pregosin, P.S.; Valentini, M.; Worle, M.; Seebach, D. X-Ray, molecular diffusion, and NOESY NMR studies of chiral, tetranuclear CuI catalysts based on monodentate thiol analogues of TADDOL. Angew. Chem. Int. Ed. 2000, 39, 153–156. [Google Scholar] [CrossRef]
  48. Komuro, T.; Kawaguchi, H.; Tatsumi, K. Synthesis and Reactions of triphenylsilanethiolato complexes of manganese(II), iron(II), cobalt(II), and nickel(II). Inorg. Chem. 2002, 41, 5083–5090. [Google Scholar] [CrossRef]
  49. Kuckmann, T.I.; Hermsen, M.; Bolte, M.; Wagner, M.; Lerner, H.-W. Silylchalcogenolates MESiRtBu2 (M = Na, Cu, Zn, Fe; E = S, Se, Te; R = tBu, Ph) and disilyldichalcogenides tBu2RSiE−ESiRtBu2 (E = S, Se, Te; R = tBu, Ph):  synthesis, properties, and structures. Inorg. Chem. 2005, 44, 3449–3458. [Google Scholar] [CrossRef]
  50. Melzer, M.M.; Mossin, S.; Cardenas, A.J.P.; Williams, K.D.; Zhang, S.; Meyer, K.; Warren, T.H. A Copper(II) Thiolate from Reductive Cleavage of an S-Nitrosothiol. Inorg. Chem. 2012, 51, 8658–8660. [Google Scholar] [CrossRef]
  51. Jana, A.; Jash, M.; Dar, W.A.; Roy, J.; Chakraborty, P.; Paramasivam, G.; Lebedkin, S.; Kirakci, K.; Manna, S.; Antharjanam, S.; et al. Carborane-thiol protected copper nanoclusters: Stimuli-responsive materials with tunable phosphorescence. Chem. Sci. 2023, 14, 1613–1626. [Google Scholar]
  52. Dance, I.G.; Fitzpatrick, L.J.; Craig, D.C.; Scudder, M.L. Monocyclic copper and silver tertiary alkanethiolates: Formation and molecular structures of (CuSBu-tert)4(Ph3P)2, (AgSCMeEt2)8(Ph3P)2 and (AgSBu-tert)14(Ph3P)4 and structural principles. Inorg. Chem. 1989, 28, 1853–1861. [Google Scholar] [CrossRef]
  53. Schneider, S.; Dzudza, A.; Raudasch-Sieber, G.; Marks, T.J. Copper(I) tert-Butylthiolato clusters as single-source precursors for high-quality chalcocite thin films:  precursor chemistry in solution and the solid state. Chem. Mater. 2007, 19, 2768–2779. [Google Scholar]
  54. Li, Z.; Li, X.; Lin, P.; Du, S. Self-assembly of 1D coordination polymers containing copper(I) tert-butylthiolato clusters: Structural characterization and properties. J. Cluster Sci. 2008, 19, 357–366. [Google Scholar] [CrossRef]
  55. Groysman, S.; Holm, R.H. A series of mononuclear quasi-two-coordinate copper(I) complexes employing a sterically demanding thiolate ligand. Inorg. Chem. 2009, 48, 621–627. [Google Scholar] [CrossRef] [PubMed]
  56. Najafabadi, B.K.; Corrigan, J.F. N-heterocyclic carbenes as effective ligands for the preparation of stabilized copper- and silver-t-butylthiolate clusters. Dalton Trans. 2014, 43, 2104–2111. [Google Scholar] [CrossRef]
  57. Shi, W.; Shafaei-Fallah, M.; Anson, C.E.; Rothenberger, A. A strategy for the build-up of transition-metal complexes containing tripodal [ArPOS2]2− and [ArPS3]2−ligands (Ar = 4-anisyl). Dalton Trans. 2005, 3909–3912. [Google Scholar] [CrossRef]
  58. Shi, W.; Shafaei-Fallah, M.; Anson, C.E.; Rothenberger, A. Metal thiophosphonates and related compounds: An emerging area of supramolecular coordination chemistry. Dalton Trans. 2006, 3257–3262. [Google Scholar] [CrossRef]
  59. Shafaei-Fallah, M.; Shi, W.; Fenske, D.; Rothenberger, A. Syntheses and structures of transition metal complexes with dithiophosphinato and trithiophosphinato ligands. Z. Anorg. Allg. Chem. 2006, 632, 1091–1096. [Google Scholar] [CrossRef]
  60. Dance, I.G.; Scudder, M.L.; Fitzpatrick, L.J. Preparation, distorted step structure, and topological analysis of (.mu.3-SPh)2(.mu.-SPh)2(CuPPh3)4(tol)2 (tol = toluene). Inorg. Chem. 1985, 24, 2547–2550. [Google Scholar] [CrossRef]
  61. Deivaraj, T.C.; Lai, G.X.; Vittal, J.J. Chemistry of thiocarboxylates:  synthesis and structures of neutral copper(I) thiocarboxylates with triphenylphosphine. Inorg. Chem. 2000, 39, 1028–1034. [Google Scholar] [CrossRef]
  62. Fernandez-Recio, L.; Fenske, D.; Fuhr, O. Copper Chalcogenide Cluster Compounds with Nitro-functionalized Ligand Shell. Z. Anorg. Allg. Chem. 2008, 634, 2853–2857. [Google Scholar] [CrossRef]
  63. Langer, R.; Wunsche, L.; Fenske, D.; Fuhr, O. Copper chalcogenide cluster compounds with bromo-functionalized ligand shell. Z. Anorg. Allg. Chem. 2009, 635, 2488–2494. [Google Scholar] [CrossRef]
  64. Langer, R.; Yadav, M.; Weinert, B.; Fenske, D.; Fuhr, O. Luminescence in functionalized copper thiolate clusters-synthesis and structural effects. Eur. J. Inorg. Chem. 2013, 2013, 3623–3631. [Google Scholar] [CrossRef]
  65. Johnson, B.J.; Antholine, W.E.; Lindeman, S.V.; Mankad, N.P. A Cu4S model for the nitrous oxide reductase active sites supported only by nitrogen ligands. Chem. Commun. 2015, 51, 11860–11863. [Google Scholar] [CrossRef] [PubMed]
  66. Johnson, B.J.; Antholine, W.E.; Lindeman, S.V.; Graham, M.J.; Mankad, N.P. A one-hole Cu4S cluster with N2O reductase activity: A structural and functional model for CuZ*. J. Am. Chem. Soc. 2016, 138, 13107–13110. [Google Scholar] [CrossRef]
  67. Hsu, C.-W.; Rathnayaka, S.C.; Islam, S.M.; MacMillan, S.N.; Mankad, N.P. N2O reductase activity of a [Cu4S] cluster in the 4CuI redox state modulated by hydrogen bond donors and proton relays in the secondary coordination sphere. Angew. Chem. Int. Ed. 2020, 132, 637–641. [Google Scholar] [CrossRef]
  68. Mankad, N.P. Triazenide-supported [Cu4S] structural mimics of CuZ that mediate N2O disproportionation rather than reduction. Chem. Sci. 2024, 15, 1820–1828. [Google Scholar] [CrossRef]
  69. Singh, A.; Singh, A.; Singh, N.; Jang, D.O. A 2-mercaptobenzimidazole-based emissive Cu(I) complex for selective determination of iodide with large Stokes shift. Sens. Actuators B 2017, 243, 372–379. [Google Scholar] [CrossRef]
  70. Wei, W.; Wu, M.; Gao, Q.; Zhang, Q.; Huang, Y.; Jiang, F.; Hong, M. A novel supramolecular tetrahedron assembled from tetranuclear copper(I) cluster molecules via aryl embrace interactions. Inorg. Chem. 2009, 48, 420–422. [Google Scholar] [CrossRef]
  71. Hao, Z.-M.; Guo, C.-H.; Wu, H.-S.; Zhang, X.-M. Luminescent boracite-like metal–organic frameworks constructed by Cu-centered CuCu4 tetrahedra and CuCu3 triangles with an acentric cubic superlarge cell. Cryst. Eng. Comm. 2010, 12, 55–58. [Google Scholar] [CrossRef]
  72. Hu, S.; Yu, F.-Y.; Zhang, P.; Zhou, A.-J. An acs-type metal–organic framework with an unprecedented undecanuclear copper cluster as secondary building unit. Eur. J. Inorg. Chem. 2012, 2012, 3669–3673. [Google Scholar] [CrossRef]
  73. Zhang, X.-Y.; Su, C.-H.; Pang, H.-Y.; Huang, F.-P.; Yu, Q.; Bian, H.-D. 1-(2-Hydroxyethyl)-5-mercapto-1H-tetrazole as ligand in Cu(I) complexes with or without X-ions (X = Cl, Br, I): Synthesis, crystal structures and properties. J. Coord. Chem. 2017, 70, 3650–3659. [Google Scholar] [CrossRef]
  74. Huang, D.-Y.; Hao, H.-M.; Yao, P.-F.; Qin, X.-H.; Huang, F.-P.; Yu, Q.; Bian, H.-D. CuIX (X = Cl, Br, I) inorganic networks separated and stabilized by a mercaptotetrazole ligand. Polyhedron 2015, 97, 260–267. [Google Scholar] [CrossRef]
  75. Yue, C.-Y.; Liu, F.-L.; Deng, W.-T.; Tao, J.; Hong, M.-C. Iodide-centered cuprous octatomic Ring: A luminescent molecular thermometer exhibiting dual-Emission character. Cryst. Growth Des. 2018, 18, 22–26. [Google Scholar] [CrossRef]
  76. Beswick, M.A.; Brasse, C.; Halcrow, M.A.; Raithby, P.R.; Russell, C.A.; Steiner, A.; Snaith, R.; Wright, D.S. Role of bifunctional N/S bridging in the association of metals; syntheses and structures of heterobimetallic [Ph3P)2CuL2Li(thf)2]·thf and tetrameric [{CuL(PPh3)}4](L = 1,3-benzoxazoline-2-thionate, thf = tetrahydrofuran). Dalton Trans. 1996, 3793–3797. [Google Scholar] [CrossRef]
  77. Zhang, X.; Zhang, Z.; Xiang, S.; Zhu, Y.; Chen, C.; Huang, D. Base induced C–CN bond cleavage at room temperature: A convenient method for the activation of acetonitrile. Inorg. Chem. Front. 2019, 6, 1135–1140. [Google Scholar] [CrossRef]
  78. Xu, H.; Liu, P.-T.; Li, Y.-H.; Han, F.-S. Copper-mediated direct aryl C–H cyanation with azobisisobutyronitrile via a free-radical pathway. Org. Lett. 2013, 15, 3354–3357. [Google Scholar] [CrossRef]
  79. Rong, G.; Mao, J.; Zheng, Y.; Yao, R.; Xu, X. Cu-Catalyzed direct cyanation of terminal alkynes with AMBN or AIBN as the cyanation reagent. Chem. Commun. 2015, 51, 13822–13825. [Google Scholar] [CrossRef]
  80. Elliott, Q.; Alabugin, I.V. AIBN as an electrophilic reagent for cyano group transfer. J. Org. Chem. 2023, 88, 2648–2654. [Google Scholar] [CrossRef]
  81. Li, C.; Li, W.; Henwood, A.F.; Hall, D.; Cordes, D.B.; Slawin, A.M.Z.; Lemaur, V.; Olivier, Y.; Samuel, I.D.W.; Zysman-Colman, E. Luminescent Dinuclear Copper(I) Complexes Bearing an Imidazolylpyrimidine Bridging Ligand. Inorg. Chem. 2020, 59, 14772–14784. [Google Scholar] [CrossRef] [PubMed]
  82. Yasuda, T.; Shimizu, T.; Liu, F.; Ungar, G.; Kato, T. Electro-functional octupolar π-conjugated columnar liquid crystals. J. Am. Chem. Soc. 2011, 133, 13437–13444. [Google Scholar] [CrossRef] [PubMed]
  83. Maeda, H.; Eifuku, N. Alkoxy-substituted Derivatives of π-Conjugated Acyclic Anion Receptors: Effects of Substituted Positions. Chem. Lett. 2009, 38, 208–209. [Google Scholar] [CrossRef]
  84. Percec, V.; Wang, S.; Huang, N.; Partridge, B.E.; Wang, X.; Sahoo, D.; Hoffman, D.J.; Malineni, J.; Peterca, M.; Jezorek, R.L.; et al. An Accelerated Modular-Orthogonal Ni-Catalyzed Methodology to Symmetric and Nonsymmetric Constitutional Isomeric AB2 to AB9 Dendrons Exhibiting Unprecedented Self-Organizing Principles. J. Am. Chem. Soc. 2021, 143, 17724–17743. [Google Scholar] [CrossRef]
  85. CrysAlis Pro, Oxford Diffraction: Yarnton, UK, 2010.
  86. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. A 2008, 64, 112–122. [Google Scholar] [CrossRef]
  87. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. C 2015, C71, 3–8. [Google Scholar] [CrossRef]
  88. Gagne, R.R.; Koval, C.A.; Lisensky, G.C. Ferrocene as an internal standard for electrochemical measurements. Inorg. Chem. 1980, 19, 2854–2855. [Google Scholar] [CrossRef]
Figure 1. Various tetranuclear [CuI4] clusters bridging by μx-SR ligands (x = 2, 3, or 4).
Figure 1. Various tetranuclear [CuI4] clusters bridging by μx-SR ligands (x = 2, 3, or 4).
Molecules 29 04228 g001
Scheme 1. Synthetic routes for the preparation of Cu(I) clusters 17.
Scheme 1. Synthetic routes for the preparation of Cu(I) clusters 17.
Molecules 29 04228 sch001
Figure 2. Crystal structures of complexes 2, 3, 4′, 6, and 7.
Figure 2. Crystal structures of complexes 2, 3, 4′, 6, and 7.
Molecules 29 04228 g002
Figure 3. (a) Cyclic voltammogram of complex 3 in MeCN (0.1 M nBu4NPF6) at 298 K; (b) Comparison of the oxidation processes of complexes 2, 3, 4, 4′, 6, and 7.
Figure 3. (a) Cyclic voltammogram of complex 3 in MeCN (0.1 M nBu4NPF6) at 298 K; (b) Comparison of the oxidation processes of complexes 2, 3, 4, 4′, 6, and 7.
Molecules 29 04228 g003
Figure 4. Thermogravimetric analysis curve of complex 3.
Figure 4. Thermogravimetric analysis curve of complex 3.
Molecules 29 04228 g004
Figure 5. Scope of iodobenzenes with respect to compound 10 a. a Reaction conditions: 9 (0.2 mmol), AIBN (0.3 mmol), KI (0.6 mmol), DBU (0.6 mmol), complex 3 (10 mol%), CH3CN (2.0 mL), N2, 120 °C, 12 h, isolated yields.
Figure 5. Scope of iodobenzenes with respect to compound 10 a. a Reaction conditions: 9 (0.2 mmol), AIBN (0.3 mmol), KI (0.6 mmol), DBU (0.6 mmol), complex 3 (10 mol%), CH3CN (2.0 mL), N2, 120 °C, 12 h, isolated yields.
Molecules 29 04228 g005
Scheme 2. Synthetic route for 2,4,6-tris(5-(4-(octyloxy)phenyl)thiophen-2-yl)-1,3,5-triazine.
Scheme 2. Synthetic route for 2,4,6-tris(5-(4-(octyloxy)phenyl)thiophen-2-yl)-1,3,5-triazine.
Molecules 29 04228 sch002
Figure 6. Studies of 2,4,6-tris(5-(4-(octyloxy)phenyl)thiophen-2-yl)-1,3,5-triazine for its fluorescence property.
Figure 6. Studies of 2,4,6-tris(5-(4-(octyloxy)phenyl)thiophen-2-yl)-1,3,5-triazine for its fluorescence property.
Molecules 29 04228 g006
Table 1. Oxidation–reduction potential (vs. Fc+/Fc) of complex 2, 3, 4, 4′, 6, and 7.
Table 1. Oxidation–reduction potential (vs. Fc+/Fc) of complex 2, 3, 4, 4′, 6, and 7.
ComplexEpaEpcComplexEpaEpc
2−0.03 V, +0.36 V-4+0.35 V, +0.48 V-
3+0.11 V, +0.45 V+0.25 V6−0.03 V, +0.22 V-
4+0.36 V, +0.49 V-7+0.15 V, +0.21 V-
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhu, T.; Zhan, W.; Fan, W.; Zhang, X. Research on Synthesis, Structure, and Catalytic Performance of Tetranuclear Copper(I) Clusters Supported by 2-Mercaptobenz-zole-Type Ligands. Molecules 2024, 29, 4228. https://doi.org/10.3390/molecules29174228

AMA Style

Zhu T, Zhan W, Fan W, Zhang X. Research on Synthesis, Structure, and Catalytic Performance of Tetranuclear Copper(I) Clusters Supported by 2-Mercaptobenz-zole-Type Ligands. Molecules. 2024; 29(17):4228. https://doi.org/10.3390/molecules29174228

Chicago/Turabian Style

Zhu, Tingyu, Wangyuan Zhan, Weibin Fan, and Xiaofeng Zhang. 2024. "Research on Synthesis, Structure, and Catalytic Performance of Tetranuclear Copper(I) Clusters Supported by 2-Mercaptobenz-zole-Type Ligands" Molecules 29, no. 17: 4228. https://doi.org/10.3390/molecules29174228

Article Metrics

Back to TopTop