Next Article in Journal
Effect-Directed Profiling of Strawberry Varieties and Breeding Materials via Planar Chromatography and Chemometrics
Previous Article in Journal
Effects of Auricularia auricula Polysaccharides on Gut Microbiota Composition in Type 2 Diabetic Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tris(2-Pyridyl)Arsine as a New Platform for Design of Luminescent Cu(I) and Ag(I) Complexes

by
Yan V. Demyanov
1,
Evgeniy H. Sadykov
1,
Marianna I. Rakhmanova
1,
Alexander S. Novikov
2,3,
Irina Yu. Bagryanskaya
4 and
Alexander V. Artem’ev
1,*
1
Nikolaev Institute of Inorganic Chemistry, SB RAS, 3 Acad. Lavrentiev Ave., 630090 Novosibirsk, Russia
2
Saint Petersburg State University, Universitetskaya Nab. 7/9, 199034 Saint Petersburg, Russia
3
Peoples’ Friendship University of Russia (RUDN University), Miklukho-Maklaya Street 6, 117198 Moscow, Russia
4
N. N. Vorozhtsov Novosibirsk Institute of Organic Chemistry, SB RAS, 9 Acad. Lavrentiev Ave., 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(18), 6059; https://doi.org/10.3390/molecules27186059
Submission received: 23 August 2022 / Revised: 5 September 2022 / Accepted: 8 September 2022 / Published: 16 September 2022
(This article belongs to the Section Inorganic Chemistry)

Abstract

:
The coordination behavior of tris(2-pyridyl)arsine (Py3As) has been studied for the first time on the example of the reactions with CuI, CuBr and AgClO4. When treated with CuI in CH2Cl2 medium, Py3As unexpectedly affords the scorpionate complex [Cu(Py3As)I]∙CH2Cl2 only, while this reaction in MeCN selectively leads to the dimer [Cu2(Py3As)2I2]. At the same time, the interaction of CuBr with Py3As exclusively gives the dimer [Cu2(Py3As)2Br2]. It is interesting to note that the scorpionate [Cu(Py3As)I]∙CH2Cl2, upon fuming with a MeCN vapor (r.t., 1 h), undergoes quantitative dimerization into the dimer [Cu2(Py3As)2I2]. The reaction of Py3As with AgClO4 produces complex [Ag@Ag4(Py3As)4](CIO4)5 featuring a Ag-centered Ag4 tetrahedral kernel. At ambient temperature, the obtained Cu(I) complexes exhibit an unusually short-lived photoluminescence, which can be tentatively assigned to the thermally activated delayed fluorescence of (M + X) LCT type (M = Cu, L = Py3As; X = halogen). For the title Ag(I) complexes, QTAIM calculations reveal the pronounced argentophilic interactions for all short Ag∙∙∙Ag contacts (3.209–3.313 Å).

1. Introduction

Over the past decade, luminescent Cu(I) and Ag(I) complexes have attracted a considerable attention due to their intriguing structural diversity [1,2,3,4,5,6,7,8,9,10,11] and ability to exhibit efficient phosphorescence or thermally activated delayed fluorescence (TADF) [12,13,14,15,16,17,18,19], or even dual emission [20,21]. Owing to these features, such compounds are now considered as promising emitters for energy-efficient OLEDs of second and third generations (PHOLED and TADF OLED) [22,23], just as for light-emitting electrochemical cells (LEECs) [24]. Moreover, Cu(I) and Ag(I) complexes are reported to be perspective luminescent sensors and X-ray scintillators, as well as “smart” materials, the emission of which is sensitive to the external stimuli (temperature, pressure, chemicals) [25,26,27,28].
To design luminescent Cu(I) and Ag(I) complexes, various C-, N-, and P-donor ligands such as carbenes, azaheterocycles, and phosphines are commonly exploited [29,30,31,32,33,34]. At that, “heavy pnictine(III)”-based ligands, e.g., arsines, remain almost unexplored in this regard. Meanwhile, recent studies [35,36,37] have demonstrated that the arsine ligands can be preferable over similar phosphines to fabricate Cu(I)-based TADF materials [38]. Indeed, the higher spin-orbital coupling (SOC) strength of arsenic (ζl = 1202 cm−1) [39] against phosphorus (ζl = 230 cm−1) [39] makes the emission rate of Cu(I)-arsine complexes much faster compared to that of similar Cu(I)-phosphine derivatives [38]. Considering that the triplet or TADF emitters with short decay times (<2 μs) are essential for OLED related applications, the design of new Cu(I)-arsine complexes, just as their Ag(I) congeners, represents a daunting challenge.
Working in this field, we have paid attention to pyridylarsines, “heavy pnictogen” analogs of pyridylphosphines. The latter are a famous family of ligands which are widely used for the design of Cu(I) and Ag(I) complexes showing very efficient TADF or/and phosphorescence [40,41,42,43,44]. The survey of the literature reveals, however, that the azaheterocyclic-substituted arsines are very limited in number [45,46,47], and their Cu(I)/Ag(I) derivatives are even scarcer [38,46,47]. Herein, for the first time, we have employed tris(2-pyridyl)arsine as a stabilizing platform for Cu(I) and Ag(I) complexes. The latter have been studied in terms of solid-state luminescence, thermal stability and electronic structure.

2. Results and Discussion

2.1. Synthesis and Characterization

It was reported that tris(2-pyridyl)phosphine (Py3P) reacted with Cu(I) halides to irreversibly give air-stable dimeric complexes [Cu2(Py3P)2X2] (X = Cl, Br, I) in high yields [48]. Therefore, it would be reasonable to expect that Py3As could also give the related complexes. Meanwhile, we have found that the reactivity of Py3As differs from that of Py3P. Namely, CuI easily reacts with Py3As in CH2Cl2 to furnish the unexpected scorpionate [Cu(Py3As)I] (1) only, isolated as a solvate 1·CH2Cl2 in 89% yield (Scheme 1). By contrast, CuBr under similar conditions affords a dimeric complex [Cu2(Py3As)2Br2] (2) in 92% yield. Our numerous attempts to synthesize CuBr-based scorpionate under the varied conditions (different Cu/Py3As ratios, diverse solvents) failed: complex 2 was already formed in all the experiments. Of note, CuCl also easily reacts with Py3As, but a product formed was found to be easily oxidized in air to produce unidentified green Cu(II) complexes.
To explain the different reactivity of Py3As towards CuBr and CuI, the dimerization energies for the equilibriums 2[Cu(Py3As)X] ↔ [Cu2(Py3As)2X2] (X = Br, I) have been assessed at the PBE0/def2-TZVP level of the theory. The performed calculations reveal that the formation of a dimer is indeed thermodynamically more preferable for X = Br by 2.34 kcal∙mol–1, whereas scorpionate is favorable for X = I by 0.97 kcal∙mol–1 (Supplementary Materials, Figure S10). It should be noted, however, that an additional stabilization of scorpionate [Cu(Py3As)I] in 1·CH2Cl2 is possible through the solvate molecules (vide infra).
Our attempts to transform 2 into scorpionate [Cu(Py3As)Br] using recrystallization from different solvents failed. Meanwhile, we have found that the fuming of 1·CH2Cl2 with MeCN vapor (23 °C, 1 h) results in quantitative and irreversible dimerization into [Cu2(Py3As)2I2] (Figure 1a). The reaction is accompanied by the changing of the parent emission color of 1 (vide infra) to green, which is specific for complexes of [Cu2(Py3As)2X2] type (X = P or As). The powder X-ray diffraction (PXRD) pattern of the product [Cu2(Py3As)2I2], differs from that of [Cu2(Py3As)2Br2] (2), but closely resembles that of a similar phosphine derivative, [Cu2(Py3P)2Br2] (Figure 1b). Since the attempts to grow X-ray quality crystals of 1a met with no success, its structure has been proved by PXRD, mid-IR and 1H NMR data only. Eventually, we have found that complex 1a can be successfully synthesized in an almost quantitative yield by the straightforward reaction of Py3As with CuI in MeCN (r.t., 10 min) (Figure 1a). Thus, a noticeable effect of solvents on the reaction outcome has been established for the reaction of Py3As with CuI.
In the next step, we have examined the complexation of Py3As with Ag(I) using AgClO4 as a precursor. The reaction easily occurs in MeCN medium, and the following recrystallization of the product in water produces a fournuclear cluster [Ag@Ag4(Py3As)4](CIO4)5·3H2O (3·3H2O) in high yield (Scheme 2). Of note, the same product is also formed even when the AgClO4/Py3As molar ratio deviates from the stoichiometric one (5:4) being 1:1, 2:1, or 1:2. For comparison, the reaction of bis(2-pyridyl)phenylarsine (Py2PhAs) with AgClO4 under similar conditions has also been implemented to deliver a dinuclear complex [Ag2(Py2PhAs)2(MeCN)2](CIO4)2·CH3CN (4·CH3CN), the structure of which resembles that of 2. Again, the AgClO4/Py2PhAs ratio does not affect the reaction outcome: only complex 2 is formed.
The X-ray derived structures of the prepared complexes are displayed in Figure 2, and the selected interatomic distances are listed in Table 1. In the structure of 1·CH2Cl2 (Figure 2), the Cu atom is coordinated by Py3As in a N,N′,N″-tripodal manner, and the iodine atom completes a distorted tetrahedral environment of the metal (τ4 = 0.85) [49]. Note that the iodine atom of 1 deviates from the axis passing from As and Cu atoms by only 0.8°, unlike the similar complex with tris(2-pyridyl)arsine oxide, where such deviation reaches 9.4° [50]. The dihedral angles between averaged pyridine planes of 1 are ≈ 114.3°, 114.4° and 131.2°.
Complex 2 consists of two CuBr units bridged by two Py3As ligands (μ2-As,N,N’) in a head-to-tail manner so that an inversion center is located in the middle of the molecule (Figure 2). Therefore, each Cu atom adopts a distorted tetrahedral [Cu@AsN2Br] environment (τ4 = 0.90). The intramolecular Cu∙∙∙Cu distances (ca. 3.94 Å) are too long for the appearance of metallophilic interactions (cf. twice van der Waals radius of Cu is 2.80 Å [51]). The dihedral angle between planes of the coordinated pyridine rings of 2 is about 124.8°. Overall, the structure of 2 closely resembles that of similar Py3P-based complex [Cu2(Py3P)2Br2] [48].
Crystals of 3·3H2O contain three independent [Ag@Ag4(Py3As)4]5+ cations charge-balanced by non-coordinated ClO4 anions, as well as the lattice water molecules. The independent [Ag@Ag4(Py3As)4]5+ cations have a similar structure, which consists of a C3-symmetrical Ag@Ag4 tetrahedral kernel supported by four Py3As ligands. The central Ag atom adopts a slightly distorted tetrahedral geometry (τ4 = 0.98) constituted of four As atoms. Each vertex Ag atom of the Ag@Ag4 kernel is coordinated by three N atoms of three neighboring Py3As ligands, thus accepting a trigonal pyramidal geometry [3 + 1]. Considering that the average distance between central and vertex Ag atoms of Ag@Ag4 kernel (3.25 Å) is shorter than twice van der Waals Ag radius (3.44 Å) [51], argentophilic interactions could occur (vide infra). To sum up, Py3As ligands in 3 exhibit an As,N,N′,N″-coordination manner that is similar to that of Py3P in [Ag@Ag4(Py3P)4]5+ complexes [52].
The structure of the cationic part of 4·CH3CN, [Ag2(Py2PhAs)2(MeCN)2]2+, is formed by two Ag(I) cations bridged by two Py2AsPh ligands in a head-to-tail manner. Both metal cations are also ligated by MeCN, thereby adopting a distorted trigonal pyramidal geometry. Again, the Ag…Ag distance in 4 being 3.2206(4) Å implies argentophilic interaction, which is actually taking pace according to the theoretical calculations (vide infra).
The synthesized compounds are moderately soluble in dichloromethane and acetonitrile (1·CH2Cl2 and 2), or in water (3·3H2O and 4·CH3CN). All of them are air-stable, except for 4·CH3CN, which quickly loses the coordinated acetonitrile molecules upon storage in air. The 1H NMR spectra of 14 demonstrate one set of signals from the coordinated arsines and ancillary ligands (Figures S3–S7), indicating the existence of symmetrical species in a solution (closely uninvestigated). The mid-IR spectra of these complexes show characteristic bands of the coordinated pyridyl-containing arsenic ligands (Figure S8). Moreover, complexes 3·3H2O and 4·CH3CN display a broad band at 1097–1099 cm−1 belonged of νCl–O stretching vibrations of free ClO4 anions, and specific bands from H2O and MeCN molecules/ligands, respectively, i.e., νO–H = 3610 cm−1 and νC≡N = 2268 cm−1.
The thermal stability of the above complexes has been studied by TGA, DTG and DTA techniques under argon atmosphere (Figure S9). The solvate molecules of 1·CH2Cl2 are lost at the range of 100–127 °C, but complex 1 itself is stable up to ≈200 °C. A higher stability is inherent in 2, which begins to decompose at about 240 °C. Compounds 3·3H2O and 4·CH3CN lose the solvate molecules at 120–140 °C and 135–155 °C, respectively, after which they remain stable at least up to 270 °C.

2.2. Theoretical Consideration

The electronic structure of luminescent Cu(I) complexes 1 and 2 has been investigated at the PBE0/def2TZVP level of the theory (for details, see §6 in ESI) to understand electronic transitions responsible for the excitation. For both complexes, the highest occupied molecular orbital (HOMO) and nearby HOMO-n are largely contributed by metal’ d-orbitals and the lone pairs in halogen atoms (Figure 3, Figures S11, S12; Tables S2, S3). The lowest unoccupied molecular orbital (LUMO) and nearby LUMO+n are mainly π-orbitals on the pyridine rings (Figure 3, Figures S11, S12; Tables S2, S3). Interestingly, a lone pair in As atom does not contribute to the highest MOs (HOMO–HOMO-6). The fact that the frontier MOs of 1 and 2 are well separated in the space indicates a quite small energy gap between the lowest singlet (S1) and triplet (T1) exited states. Overall, the predicted HOMO/LUMO distribution is very typical for emitting TADF Cu(I) halide complexes.
A detailed consideration of HOMOs of 1 and 2 reveals small energy gaps separation between the HOMO and HOMO-1 levels which are, moreover, populated by d-orbitals with different spatial orientations (Figures S11 and S12). In particular, the HOMO-1/HOMO separation is 140 cm−1 for 2, and it is just 2 cm−1 for 1. According to the literature [12,13], such a scenario demonstrates a strong SOC mixing of the S1 and T1 states, which are originated from HOMO-1 → LUMO and HOMO → LUMO transitions, respectively. This, in turn, accelerates the rates of both S1 → T1 and T1 → S1 spin-forbidden processes, and hence, increases total rate of the luminescence. The experimental results fully confirm these predictions (vide infra). TD-DFT computations of 1 and 2 testify to the (M + X) LCT character (X = Br or I) of the low-energy absorptions (Tables S4, S5) that is specific for the related complexes. Furthermore, the (M + X) LCT nature of the lowest excited states is confirmed by a specific spin distribution in the computed T1 state of 1 (Figure S15). The lowest (LSOMO) and highest (HSOMO) single occupied molecular orbitals of the T1 state of 1 (Figure S15) closely resemble HOMO and LUMO in its S0 state (Figure 3). The calculated ∆E(T1–S0) energy gap for the optimized states of 1 being 1.81 eV reasonably agrees with the emission energy of 1·CH2Cl2 at a pure phosphorescence regime (77 K), i.e., 1.98 eV or 625 nm (vide infra). Therefore, one can expect that the emitting excited states of the compounds discussed should be of the (M + X) LCT type.
For Ag(I) complexes 3 and 4, which appear to be almost non-emissive, argentophilic interactions have been examined by QTAIM (quantum theory “atoms in molecule”) method at the ωB97XD/DZP-DKH level (for details, see §6.3 in ESI). The QTAIM analysis of model structures reveals the presence of the bond critical points (3, –1) for metallophilic interactions in 3 and 4 (Table 2). The low magnitude of the electron density (0.017–0.021 a.u.), positive values of the Laplacian of electron density (0.028–0.030 a.u.), and negative energy density (from –0.002 to –0.003 a.u.) in the bond critical points (3, –1) for Ag(I)···Ag(I) interactions in 3 and 4 are typical for metallophilic interactions in other metal complexes [53,54,55,56,57,58,59]. The balance between the Lagrangian kinetic energy G(r) and potential energy density V(r) at the bond critical points (3, –1) for Ag(I)···Ag(I) interactions in 3 and 4 [viz. –G(r)/V(r) < 1] shows some covalent contribution in these short contacts [60]. The Laplacian of electron density is typically decomposed into the sum of contributions along the three principal axes of maximal variation, giving three eigenvalues of the Hessian matrix (λ1, λ2 and λ3), and the sign of λ2 can be utilized to distinguish the bonding (attractive, λ2 < 0) weak interactions from the non-bonding ones (repulsive, λ2 > 0) [61,62]. Thus, the Ag(I)···Ag(I) interactions in 3 and 4 have an attractive character. For illustration, the calculated electron density distribution at the metal atoms of 4 is plotted in Figure 4.

2.3. Photoluminescence of 1 and 2

At ambient temperature, Cu(I) complexes 1, 1a and 2 emit pronounced solid state photoluminescence (PL), whereas Ag(I) derivatives 3 and 4 appear to be almost non-emissive. The recorded emission and excitation spectra of 1, 1a and 2 are shown in Figure 5, and the measured PL properties are given in Table 3. As follows from these data, scorpionate 1 manifests an orange PL, while dimers 1a and 2 emit in a green region. In the terms of PL quantum yields (PLQYs), the emission of the studied compounds is moderate at 298 K (10–14%). All the emission profiles are of broad and structureless shape that is inherent in PL of the charge transfer origin [14]. The corresponding excitation curves are displayed by typical bands extending from the UV-edge and are sharply falling close at ≈450 nm (1a, 2) or 590 nm (1). The fact that the emission and emission profiles of iodide 1a and bromide 2 are almost superimposable is not confusing; previously, the similar cases were documented for other halide complexes, including the related ones [38,48,63]. The PL lifetimes of 1, 1a and 2 at 298 K are remarkably short (0.8–1.9 μs) compared to the most known Cu(I) complexes. Accordingly, the radiative constants (kr = PLQY/τ) being (0.53–1.75)∙105 s–1 are relatively high, thereby indicating a strong SOC effect that is also predicted by DFT calculations (vide supra). For comparison, radiative rates (kr) of the similar Py3P-based complexes [Cu2(Py3P)2X2] at 298 K are much lower: 2.9∙104 and 2.6∙104 s–1 for X = Br and I, respectively. The acceleration of the emission rates observed in the arsine complexes is obviously attributed to much stronger SOC effect of arsenic compared to that of phosphorus. Previously, this effect was already demonstrated for Cu(I)-arsine complexes [38].
Temperature-dependent PL spectra of 1 and 2 (Figure 6) demonstrate the significant enhancement of PL intensity upon cooling that is accompanied by a slight bathochromic shift of the bands by 10–20 nm (Table 3). When passing from 298 to 77 K, the PL lifetimes of 1 and 2 increase by 13 and 25 times, thus amounting 25 and 23 μs at 77 K, respectively. Taken together, these observations suggest the TADF manifestation at ambient temperature and phosphorescence at 77 K. According to the DFT and TD-DFT computations, the 1(M + X) LCT emissive state is responsible for TADF (298 K), and 3(M + X) LCT state is active at the phosphorescence regime (77 K). It should be underlined that the same emission scheme was previously proved for the related dimeric complexes based on Py3P and PhPy2As ligands. The ∆EST energy gaps between the 1(M + X) LCT and 3(M + X) LCT states of 1 and 2 can be roughly estimated by the red-shifting of the left flank of emission bands at their half height (Figure S17). The estimated ∆EST gaps, being 340 and 890 cm–1 for 1 and 2, respectively, fall close to the range of such values (∆EST < 1500 cm–1) for TADF-active Cu(I) complexes [12,13,14].

3. Materials and Methods

3.1. General

All synthetic procedures were carried out under an argon atmosphere using the standard Schlenk technique. CuI (≥99%, Sigma, Gillingham, UK), AgClO4 (97%, Alfa Aesar, Heysham, UK), and MeCN (HPLC grade, Cryochrom, St. Petersburg, Russia) were used as purchased. Prior to use, CuBr was freshly synthesized by the treatment of CuBr2 with Cu powder in MeCN solution. Tris(2-pyridyl)arsine, [50] and bis(2-pyridyl)phenylarsine (Py2AsPh) [38] were prepared according to the literature procedures. 1H NMR spectra were recorded using a Bruker AV-500 spectrometer at 500.13 MHz. Chemical shifts were reported in δ (ppm) relative to residual peaks of protonated CDCl3, DMSO-d6, and CD3CN. FT-IR spectra were recorded on a Bruker Vertex 80 spectrometer. Powder X-ray diffraction patterns were recorded on a Shimadzu XRD-7000 diffractometer (Cu-Kα radiation, Ni – filter, 3–35° 2θ range, 0.03° 2θ step, 5s per point). Thermogravimetric analyses (TGA – c-DTA – DTG) were carried out in a closed Al2O3 pan under argon flow at a 10 °C/min–1 heating rate using a NETZSCH STA 449 F1 Jupiter STA. CHN microanalyses were performed on a MICRO cube analyzer.
Emission and excitation spectra were recorded on a Fluorolog 3 spectrometer (Horiba Jobin Yvon) equipped with a cooled PC177CE-010 photon detection module and an R2658 photomultiplier. The absolute PLQYs were determined at 298 K using a Fluorolog 3 Quanta-phi integrating sphere. Temperature-dependent excitation and emission spectra as well as emission decays were recorded using an Optistat DN optical cryostat (Oxford Instruments) integrated with the above spectrometer.

3.2. [Cu(Py3As)I]·CH2Cl2 (1·CH2Cl2)

A mixture of CuI (16.5 mg, 0.087 mmol) and Py3As (30 mg, 0.097 mmol) in CH2Cl2 (2 mL) was stirred at room temperature for 10 min. To the resulting solution, hexane (1 mL) was added dropwise and a precipitate formed was centrifuged and dried in air. Orange powder. Yield: 39 mg (89%). Single crystals of 1·CH2Cl2 were grown by slow evaporation of CH2Cl2 solution for overnight. 1H NMR (500.13 MHz, CDCl3, ppm), δ: 9.09 (ddd, J = 5.0 Hz, J = 1.9 Hz, J = 0.9 Hz, 3H, H6 in Py), 7.90 (dt, J = 7.5 Hz, J = 1.2 Hz, 3H, H4 in Py), 7.71 (dt, J = 7.6 Hz, J = 1.8 Hz, 3H, H5 in Py), 7.35 (ddd, J = 7.7 Hz, J = 5.0 Hz, J = 1.3 Hz, 3H, H3 in Py), 5.31 (s, 2H in CH2Cl2). FT-IR (KBr, cm–1): 409 (m), 457 (w), 482 (s), 617 (w), 638 (w), 669 (w), 702 (m), 733 (s), 758 (vs), 773 (m), 787 (m), 897 (vw), 1003 (m), 1045 (m), 1088 (w), 1103 (w), 1152 (m), 1227 (w), 1271 (m), 1422 (s), 1447 (vs), 1553 (m), 1572 (s), 1636 (vw), 2955 (w), 3028 (w), 3042 (w). Calculated for C16H14AsCuICl2N3 (584.58): C, 32.9; H, 2.4; N, 7.2. Found: C, 33.0; H, 2.5; N, 7.2.

3.3. [Cu2(Py3As)2I2] (1a)

Method 1: A solid sample of 1·CH2Cl2 (15 mg, 0.026 mmol) was placed in a 3 mL vial, which was then placed in a closed 50 mL weighing bottle containing MeCN (≈0.5 mL) on a bottom. Exposure of a solid sample 1·CH2Cl2 under MeCN vapor at ambient temperature for 1 h results in the formation of 1a as an off-white solid. Yield: 99% (12.5 mg).
Method 2: A mixture of CuI (8.5 mg, 0.045 mmol) and Py3As (15 mg, 0.049 mmol) in CH3CN (1 mL) was stirred at room temperature for 10 min. The formed precipitate was centrifuged and dried in vacuum. White powder. Yield: 40 mg (89%).
1H NMR (500.13 MHz, CD3CN, ppm), δ: 8.88 (d, J = 5.0 Hz, 6H, H6 in Py), 8.03 (d, J = 7.6 Hz, 6H, H4 in Py), 7.85 (t, J = 7.7 Hz, 6H, H5 in Py), 7.49-7.43 (m, 6H, H3 in Py). FT-IR (KBr, cm–1): 405 (w), 419 (m), 459 (w), 474 (m), 484 (s), 617 (w), 637 (m), 741 (w), 756 (vs), 779 (m), 988 (m), 1007 (m), 1049 (m), 1088 (w), 1103 (w), 1119 (w), 1155 (m), 1227 (w), 1275 (m), 1410 (s), 1423 (vs), 1447 (vs), 1558 (s), 1570 (s), 1578 (s), 1634 (w), 2974 (m), 3038 (m), 3059 (m). Calculated for C30H24As2Cu2I2N6 (999.29): C, 36.1; H, 2.4; N, 8.4. Found: C, 36.0; H, 2.4; N, 8.3.

3.4. [Cu2(Py3As)2Br2] (2)

A mixture of CuBr (12 mg, 0.083 mmol) and Py3As (26 mg, 0.084 mmol) in CH3CN (1 mL) was stirred at room temperature for 10 min. The formed precipitate was centrifuged and dried in vacuum. White powder. Yield: 69 mg (92%). Single crystals of 2 were grown by a diffusion of Et2O vapor into an CH3CN solution for overnight. 1H NMR (500.13 MHz, CDCl3, ppm), δ: 9.07 (d, J = 4.8 Hz, 6H, H6 in Py), 7.92–7.86 (m, 6H, H4 in Py), 7.71 (t, J = 7.9 Hz, 6H, H5 in Py), 7.38–7.32 (m, 6H, H3 in Py). FT-IR (KBr, cm–1): 405 (m), 417 (m), 459 (w), 474 (m), 490 (s), 619 (w), 637 (m), 671 (w), 696 (w), 733 (w), 743 (w), 766 (vs), 775 (vs), 889 (vw), 966 (w), 989 (m), 1007 (m), 1045 (m), 1082 (w), 1105 (w), 1121 (w), 1153 (m), 1233 (vw), 1277 (w), 1414 (s), 1427 (s), 1449 (vs), 1558 (m), 1578 (s), 1638 (vw), 2953 (w), 2978 (w), 3032 (m), 3057 (w). Calculated for C30H24As2Cu2Br2N6 (905.29): C, 39.8; H, 2.7; N, 9.3. Found: C, 39.7; H, 2.8; N, 9.3.

3.5. [Ag@Ag4(Py3As)4](CIO4)5·3H2O (3·3H2O)

A mixture of AgClO4 (20.5 mg, 0.099 mmol) and Py3As (25 mg, 0.081 mmol) in CH3CN (1 mL) was stirred at room temperature for 10 min. To the resulting solution, diethyl ether (1 mL) was then added, and the precipitate formed was centrifuged and dried in vacuum. White powder. Yield: 156 mg (83%). Single crystals of 3·3H2O were grown by slow evaporation of water solution for few days. 1H NMR (500.13 MHz, CD3CN, ppm), δ: 8.27–8.21 (m, 12H, H6 in Py), 7.89 (t, J = 7.6 Hz, 12H, H4 in Py), 7.40 (t, J = 6.5 Hz, 12H, H5 in Py), 7.02 (d, J = 7.8 Hz, 12H, H3 in Py). FT-IR (KBr, cm–1): 405 (m), 467 (s), 507 (m), 621 (vs), 665 (m), 702 (m), 758 (s), 903 (m), 928 (m), 1005 (s), 1047 (vs), 1082 (vs), 1097 (vs), 1165 (m), 1246 (vw), 1287 (w), 1427 (s), 1452 (s), 1558 (m), 1578 (s), 1630 (w), 2995 (w), 3084 (m), 3610 (w). Calculated for C60H54As4Ag5N12Cl5O23 (2327.43): C, 31.0; H, 2.3; N, 7.2. Found: C, 30.9; H, 2.5; N, 7.2.

3.6. [Ag2(Py2AsPh)2(MeCN)2](ClO4)2·CH3CN (4·CH3CN)

A mixture of AgClO4 (15 mg, 0.073 mmol) and bis(2-pyridyl)phenylarsine (25 mg, 0.081 mmol) in CH3CN (1 mL) was stirred at room temperature for 10 min. To a resulting solution, diethyl ether (1 mL) was then added, and the precipitate formed was centrifuged and dried in vacuum. White powder. Yield: 72 mg (85%). Single crystals of 4·CH3CN were grown by a diffusion of Et2O vapor into an CH3CN solution for overnight. 1H NMR (500.13 MHz, DMSO-d6, ppm), δ: 8.73 (d, J = 5.0 Hz, 4H, H6 in Py), 7.87 (t, J = 7.8 Hz, 4H, H4 in Py), 7.56–7.46 (m, 14H, H5 in Py, o-H, m-H and p-H in Ph), 7.43 (d, J = 7.9 Hz, 4H, H3 in Py). FT-IR (KBr, cm–1): 405 (w), 469 (m), 488 (m), 623 (s), 696 (m), 743 (s), 764 (m), 926 (w), 989 (m), 1001 (m), 1049 (s), 1099 (vs), 1161 (m), 1236 (vw), 1287 (w), 1425 (s), 1437 (m), 1454 (s), 1483 (w), 1560 (m), 1580 (m), 1636 (w), 1973 (vw), 2251 (w), 2268 (w), 3059 (w). Calculated for C38H35As2Ag2N7Cl2O8 (1154.21): C, 39.5; H, 3.1; N, 8.5. Found: C, 39.5; H, 3.0; N, 8.5.

3.7. X-ray Crystallography

The data were collected on a Bruker Kappa Apex II CCD diffractometer using φ, ω-scans of narrow (0.5°) frames with Mo Kα radiation (λ = 0.71073 Å) and a graphite monochromator. The structures were solved by direct methods SHELXL97 and refined by a full matrix least-squares anisotropic-isotropic (for H atoms) procedure using the SHELXL-2018/3 programs set [64]. Absorption corrections were applied using the empirical multiscan method with the SADABS program [65]. The positions of the hydrogen atoms were calculated with the riding model. Free solvent accessible volume in compound 3 derived from PLATON routine analysis was found to be 8.5% (1001.0 Å3). This volume is occupied by H2O molecules, but this structure is based on very weak data (a better dataset cannot be obtained). Therefore, we employed the PLATON/SQUEEZE procedure to calculate the contribution to the diffraction from H2O molecules and thereby produced a set of H2O-free diffraction intensities. The final formula of 3, C60H48Ag5As4N12Cl5O10, was derived from the SQUEEZE results. The crystallographic data and details of the structure refinements are summarized in Table S1.
CCDC 2090740, 2090741, 2126017 and 2126016 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Center at https://www.ccdc.cam.ac.uk/structures/ (accessed on 1 July 2022).

4. Conclusions

In conclusion, the reactions of tris(2-pyridyl)arsine (Py3As), an earlier unexplored ligand, with CuI, CuBr, and AgClO4 have been studied. The interaction with CuI features a remarkable solvent-directing effect on the product structure. This reaction in CH2Cl2 results in the crystallization of mononuclear scorpionate [Cu(Py3As)I]∙CH2Cl2, whilst MeCN favors the selective formation of the dimer [Cu2(Py3As)2I2]. Noteworthy, scorpionate [Cu(Py3As)I]∙CH2Cl2, when fumed with MeCN vapor, easily (r.t., 1 h) and quantitatively dimerizes into [Cu2(Py3As)2I2]. On the contrary, the treatment of Py3As with CuBr, regardless of solvent nature, affords the dimer [Cu2(Py3As)2Br2] only. At ambient temperature, the above Cu(I) complexes manifest visible photoluminescence with noticeably short decay times (0.8–1.9 μs) and moderate quantum yields (10–14%). Taking into account the results of DFT computations and temperature-dependent photophysical measurements, the emission observed has been tentatively assigned to the thermally activated delayed fluorescence of (M + X) LCT kind (M = Cu, L = Py3As; X = halogen).
Complexation of Py3As with AgClO4 results in the assembly of complex [Ag@Ag4(Py3As)4](CIO4)5, containing Ag-centered tetrahedron Ag@Ag4 supported by four Py3As ligands. According to QTAIM analysis of this cluster, argentophilic interactions between its central and peripheral Ag(I) cations are observed.
To sum up, our findings reveal that the coordination chemistry of Py3As in some cases, e.g., in the reactions with Cu halides, may differ from that of Py3P. Moreover, the Py3As-derived Cu(I) complexes demonstrate higher emission rates (kr) at 298 K compared to the similar Py3P-based analogs, thus highlighting the prospects of employing the arsine ligands for the design of short-lived TADF materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27186059/s1, Table S1. X-ray crystallographic data for 1·CH2Cl2, 2, 3·3H2O and 4·CH3CN; Figure S1. Experimental and simulated PXRD patterns of an as-synthesized sample of 1·CH2Cl2; Figure S2. Experimental and simulated PXRD patterns of an as-synthesized sample of 2; Figure S3. 1H NMR spectrum of 1·CH2Cl2 (CDCl3, 25 °C); Figure S4. 1H NMR spectrum of 1a (CD3CN, 25 °C); Figure S5. 1H NMR spectrum of 2 (CDCl3, 25 °C); Figure S6. 1H NMR spectrum of 3·3H2O (CD3CN, 25 °C); Figure S7. 1H NMR spectrum of 4·CH3CN (DMSO-d6, 25 °C); Figure S8. FT-IR spectra for the complexes 1·CH2Cl2, 1a, 2, 3·3H2O and 4·CH3CN in the 400–3250 cm–1 region; Figure S9. TGA&DTG curves for 1·CH2Cl2, 2, 3·3H2O and 4·CH3CN; Figure S10. Gibbs free energies calculated for the equilibria 2 [Cu(Py3As)X] ↔ [Cu2(Py3As)2X2] (X = Br, I) at PBE0/def2TZVP level; Figure S11. Selected frontier molecular orbitals (isovalue = 0.04) calculated for the optimized S0 state geometry of [Cu(AsPy3)I] (1) at PBE0/def2TZVP level; Figure S12. Selected frontier molecular orbitals (isovalue = 0.04) calculated for the optimized S0 state geometry of [Cu2(Py3As)2Br2] (2) at PBE0/def2TZVP level; Figure S13. The UV-Vis spectrum of [Cu(AsPy3)I] (1) (CH2Cl2, 298 K) and absorption patterns (vertical bars) calculated at the TD-PBE0/def2TZVP level; Figure S14. The UV-Vis spectrum of [Cu2(Py3As)2Br2] (2) (MeCN, 298 K) and absorption patterns (vertical bars) calculated at the TD-PBE0/def2TZVP level; Table S2. Atomic contributions to selected molecular orbitals of [Cu(AsPy3)I] (1) in the ground state (S0) geometry according Mulliken population analysis at PBE0/def2TZVP level; Table S3. Atomic contributions to selected molecular orbitals of [Cu2(Py3As)2Br2] (2) in the ground state (S0) geometry according Mulliken population analysis at PBE0/def2TZVP level; Table S4. Calculated (TD-PBE0/def2-TZVP) energies and characters of the main singlet excitations (f > 0.01) of [Cu(AsPy3)I] (1); Table S5. Calculated (TD-PBE0/def2-TZVP) energies and characters of the main singlet excitations (f > 0.01) of [Cu2(Py3As)2Br2] (2); Figure S15. LSOMO and HSOMO (isovalue = 0.04) calculated for the optimized gas phase T1 state geometry of [Cu(AsPy3)I] (1) at PBE0/def2TZVP level; Figure S16. Temperature dependent excitation spectra of 1·CH2Cl2 (a) and 2 (b) recorded at λreg = 595 and 520 nm, respectively; Figure S17. Red-shifting emission profile of 1·CH2Cl2 (left) and 2 (right) upon cooling from 298 to 77 K; Figure S18. PL decay kinetics for 1·CH2Cl2 (left) and 2 (right). Citation of reference [66,67,68,69,70,71,72,73,74,75,76,77,78].

Author Contributions

Investigation data curation, visualization, Y.V.D.; Photophysical measurements, M.I.R.; DFT calculations, E.H.S. and A.S.N.; Crystallography, I.Y.B.; Project conceptualization, administration, supervision, writing-review and editing, and funding acquisition, A.V.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Russian Science Foundation (Project No. 21-73-10110) and the Ministry of Science and Higher Education of the Russian Federation (projects No. 121031700321-3, No. 121031700313-8, and No. 1021051503141-0-1.4.1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The DFT calculations and topological analysis of the electron density distribution were supported by the RUDN University Strategic Academic Leadership Program. We thank Andrey Baranov (Nikolaev Institute of Inorganic Chemistry) for help in synthesis of 1. The authors would like to acknowledge the Multi-Access Chemical Research Center SB RAS for spectral and analytical measurements.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples used are available from the authors. Crystallographic details, 1H NMR and FT-IR spectra, TGA&DTG curves, and computation details.

References

  1. Lescop, C. Coordination-Driven Supramolecular Synthesis Based on Bimetallic Cu(I) Precursors: Adaptive Behavior and Luminescence. Chem. Rec. 2020, 21, 544–557. [Google Scholar] [CrossRef] [PubMed]
  2. Marchenko, R.D.; Sukhikh, T.S.; Ryadun, A.A.; Potapov, A.S. Synthesis, Crystal Structure, and Luminescence of Cadmium(II) and Silver(I) Coordination Polymers Based on 1,3-Bis(1,2,4-triazol-1-yl)adamantane. Molecules 2021, 26, 5400. [Google Scholar] [CrossRef] [PubMed]
  3. Moutier, F.; Schiller, J.; Calvez, G.; Lescop, C. Self-assembled luminescent Cu(I) tetranuclear metallacycles based on 3,3′-bipyridine ligands. Org. Chem. Front. 2021, 8, 2893–2902. [Google Scholar] [CrossRef]
  4. Marchenko, R.D.; Lysova, A.A.; Samsonenko, D.G.; Dybtsev, D.N.; Potapov, A.S. Synthesis, structural diversity, luminescent properties and antibacterial effects of cadmium(II) and silver(I) coordination compounds with bis(1,2,3-benzotriazol-1-yl)alkanes. Polyhedron 2020, 177, 114330. [Google Scholar] [CrossRef]
  5. Khisamov, R.; Sukhikh, T.; Bashirov, D.; Ryadun, A.; Konchenko, S. Structural and Photophysical Properties of 2,1,3-Benzothiadiazole-Based Phosph(III)azane and Its Complexes. Molecules 2020, 25, 2428. [Google Scholar] [CrossRef]
  6. Sukhikh, T.S.; Khisamov, R.M.; Konchenko, S.N. Unexpectedly Long Lifetime of the Excited State of Benzothiadiazole Derivative and Its Adducts with Lewis Acids. Molecules 2021, 26, 2030. [Google Scholar] [CrossRef]
  7. Abramov, P.A.; Komarov, V.Y.; Pischur, D.A.; Sulyaeva, V.S.; Benassi, E.; Sokolov, M.N. Solvatomorphs of (Bu4N)2[{Ag(N2-py)}2Mo8O26]: Structure, colouration and phase transition. CrystEngComm. 2021, 23, 8527–8537. [Google Scholar] [CrossRef]
  8. Chupina, A.V.; Shayapov, V.; Novikov, A.S.; Volchek, V.V.; Benassi, E.; Abramov, P.A.; Sokolov, M.N. [{AgL}2Mo8O26]n complexes: A combined experimental and theoretical study. Dalton Trans. 2020, 49, 1522–1530. [Google Scholar] [CrossRef]
  9. Chupina, A.V.; Mukhacheva, A.A.; Abramov, P.A.; Sokolov, M.N. Complexation and Isomerization of [β-Mo8O26]4− in the Presence of Ag+ and DMF. J. Struct. Chem. 2020, 61, 299–308. [Google Scholar] [CrossRef]
  10. Shmakova, A.A.; Berezin, A.S.; Abramov, P.A.; Sokolov, M.N. Self-Assembly of Ag+/[PW11NbO40]4– Complexes in Nonaqueous Solutions. Inorg. Chem. 2020, 59, 1853–1862. [Google Scholar] [CrossRef]
  11. Li, J.-J.; Liu, C.-Y.; Guan, Z.-J.; Lei, Z.; Wang, Q.-M. Anion-Directed Regulation of Structures and Luminescence of Heterometallic Clusters. Angew. Chem. Int. 2022, 61, e202201549. [Google Scholar]
  12. Leitl, M.J.; Zink, D.M.; Schinabeck, A.; Baumann, T.; Volz, D.; Yersin, H. Copper(I) Complexes for Thermally Activated Delayed Fluorescence: From Photophysical to Device Properties. Top Curr. Chem. 2016, 374, 141–174. [Google Scholar] [CrossRef]
  13. Yersin, H.; Rausch, A.F.; Czerwieniec, R.; Hofbeck, T.; Fischer, T. The triplet state of organo-transition metal compounds. Triplet harvesting and singlet harvesting for efficient OLEDs. Coord. Chem. Rev. 2011, 255, 2622–2652. [Google Scholar] [CrossRef]
  14. Yersin, H.; Czerwieniec, R.; Shafikov, M.Z.; Suleymanova, A.F. TADF Material Design: Photophysical Background and Case Studies Focusing on CuI and AgI Complexes. ChemPhysChem 2017, 18, 3508–3535. [Google Scholar] [CrossRef] [PubMed]
  15. Zhu, K.; Cheng, Z.; Rangan, S.; Cotlet, M.; Du, J.; Kasaei, L.; Kasaei, L.; Teat, S.J.; Liu, W.; Chen, Y.; et al. A New Type of Hybrid Copper Iodide as Nontoxic and Ultrastable LED Emissive Layer Material. ACS Energy Lett. 2021, 6, 2565–2574. [Google Scholar] [CrossRef]
  16. Hei, X.; Liu, W.; Zhu, K.; Teat, S.J.; Jensen, S.; Li, M.; O’Carroll, D.M.; Wei, K.; Tan, K.; Cotlet, M.; et al. Blending Ionic and Coordinate Bonds in Hybrid Semiconductor Materials: A General Approach toward Robust and Solution-Processable Covalent/Coordinate Network Structures. J. Am. Chem. Soc. 2020, 142, 4242–4253. [Google Scholar] [CrossRef]
  17. Vinogradova, K.A.; Plyusnin, V.F.; Kupryakov, A.S.; Rakhmanova, M.I.; Pervukhina, N.V.; Naumov, D.Y.; Sheludyakova, L.A.; Nikolaenkova, E.B.; Krivopalov, V.P.; Bushuev, M.B. Halide impact on emission of mononuclear copper(I) complexes with pyrazolylpyrimidine and triphenylphosphine. Dalton Trans. 2014, 43, 2953–2960. [Google Scholar] [CrossRef]
  18. Shekhovtsov, N.A.; Kokina, T.E.; Vinogradova, K.A.; Panarin, A.Y.; Rakhmanova, M.I.; Naumov, D.Y.; Pervukhina, N.V.; Nikolaenkova, E.B.; Krivopalov, V.P.; Czerwieniec, R.; et al. Near-infrared emitting copper(I) complexes with a pyrazolylpyrimidine ligand: Exploring relaxation pathways. Dalton Trans. 2022, 51, 2898–2911. [Google Scholar] [CrossRef] [PubMed]
  19. Evariste, S.; El Sayed Moussa, M.; Wong, H.-L.; Calvez, G.; Yam, V.W.-W.; Lescop, C. Straightforward Preparation of a Solid-state Luminescent Cu11 Polymetallic Assembly via Adaptive Coordination-driven Supramolecular Chemistry. Z. Anorg. Allg. Chem. 2020, 646, 754–760. [Google Scholar] [CrossRef]
  20. Shekhovtsov, N.A.; Vinogradova, K.A.; Berezin, A.S.; Sukhikh, T.S.; Krivopalov, V.P.; Nikolaenkova, E.B.; Bushuev, M.B. Excitation wavelength dependent emission of silver(I) complexes with a pyrimidine ligand. Inorg. Chem. Front. 2020, 7, 2212–2223. [Google Scholar] [CrossRef]
  21. Malakhova, Y.A.; Sukhikh, T.S.; Rakhmanova, M.I.; Vinogradova, K.A. Effect of polymorphism on the luminescent properties on silver(I) nitrate complexes with 2-amino-5-phenylpyrazine. J. Struct. Chem. 2022, 63, 485–500. [Google Scholar] [CrossRef]
  22. Dumur, F. Recent advances in organic light-emitting devices comprising copper complexes: A realistic approach for low-cost and highly emissive devices? Org. Electron. 2015, 21, 27–39. [Google Scholar] [CrossRef]
  23. Ravaro, L.P.; Zanoni, K.P.S.; de Camargo, A.S.S. Luminescent Copper(I) complexes as promising materials for the next generation of energy-saving OLED devices. Energy Rep. 2020, 6, 37–45. [Google Scholar] [CrossRef]
  24. Housecroft, C.E.; Constable, E.C. TADF: Enabling luminescent copper(I) coordination compounds for light-emitting electrochemical cells. J. Mater. Chem. C 2022, 10, 4456–4482. [Google Scholar] [CrossRef] [PubMed]
  25. Cariati, E.; Lucenti, E.; Botta, C.; Giovanella, U.; Marinotto, D.; Righetto, S. Cu(I) hybrid inorganic–organic materials with intriguing stimuli responsive and optoelectronic properties. Coord. Chem. Rev. 2016, 306, 566–614. [Google Scholar] [CrossRef]
  26. Kirakci, K.; Fejfarová, K.; Martinčík, J.; Nikl, M.; Lang, K. Tetranuclear Copper(I) Iodide Complexes: A New Class of X-ray Phosphors. Inorg. Chem. 2017, 56, 4609–4614. [Google Scholar] [CrossRef]
  27. Conesa-Egea, J.; Zamora, F.; Amo-Ochoa, P. Perspectives of the smart Cu-Iodine coordination polymers: A portage to the world of new nanomaterials and composites. Coord. Chem. Rev. 2019, 381, 65–78. [Google Scholar] [CrossRef]
  28. Evariste, S.; Khalil, A.M.; Kerneis, S.; Xu, C.; Calvez, G.; Costuas, K.; Lescop, C. Luminescent vapochromic single crystal to single crystal transition in one-dimensional coordination polymer featuring the first Cu(I) dimer bridged by an aqua ligand. Inorg. Chem. Front. 2020, 7, 3402–3411. [Google Scholar] [CrossRef]
  29. Paderina, A.V.; Koshevoy, I.O.; Grachova, E.V. Keep it tight: A crucial role of bridging phosphine ligands in the design and optical properties of multinuclear coinage metal complexes. Dalton Trans. 2021, 50, 6003–6033. [Google Scholar] [CrossRef]
  30. Conaghan, P.J.; Matthews, C.S.B.; Chotard, F.; Jones, S.T.E.; Greenham, N.C.; Bochmann, M.; Credgington, D.; Romanov, A.S. Highly efficient blue organic light-emitting diodes based on carbene-metal-amides. Nat. Commun. 2020, 11, 1758. [Google Scholar] [CrossRef]
  31. Kobayashi, A.; Ehara, T.; Yoshida, M.; Kato, M. Quantitative Thermal Synthesis of Cu(I) Coordination Polymers That Exhibit Thermally Activated Delayed Fluorescence. Inorg. Chem. 2020, 59, 9511–9520. [Google Scholar] [CrossRef]
  32. Kirst, C.; Tietze, J.; Mayer, P.; Böttcher, H.-C.; Karaghiosoff, K. Coinage Metal Complexes of Bis(quinoline-2-ylmethyl)phenylphosphine-Simple Reactions Can Lead to Unprecedented Results. ChemistryOpen 2022, 11, e202100224. [Google Scholar] [CrossRef] [PubMed]
  33. Kirst, C.; Zoller, F.; Bräuniger, T.; Mayer, P.; Fattakhova-Rohlfing, D.; Karaghiosoff, K. Investigation of Structural Changes of Cu(I) and Ag(I) Complexes Utilizing a Flexible, Yet Sterically Demanding Multidentate Phosphine Oxide Ligand. Inorg. Chem. 2021, 60, 2437–2445. [Google Scholar] [CrossRef] [PubMed]
  34. Khalil, A.M.; Xu, C.; Delmas, V.; Calvez, G.; Costuas, K.; Haouas, M.; Lescop, C. Coordination-driven supramolecular syntheses of new homo- and hetero-polymetallic Cu(I) assemblies: Solid-state and solution characterization. Inorg. Chem. Front. 2021, 8, 4887–4895. [Google Scholar] [CrossRef]
  35. Galimova, M.F.; Zueva, E.M.; Dobrynin, A.B.; Samigullina, A.I.; Musin, R.R.; Musina, E.I.; Karasik, A.A. Cu4I4-cubane clusters based on 10-(aryl)phenoxarsines and their luminescence. Dalton Trans. 2020, 49, 482–491. [Google Scholar] [CrossRef] [PubMed]
  36. 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]
  37. Kobayashi, R.; Kihara, H.; Kusukawa, T.; Imoto, H.; Naka, K. Dinuclear Rhombic Copper(I) Iodide Complexes with Rigid Bidentate Arsenic Ligands. Chem. Lett. 2021, 50, 382–385. [Google Scholar] [CrossRef]
  38. Artem’ev, A.V.; Demyanov, Y.V.; Rakhmanova, M.I.; Bagryanskaya, I.Y. Pyridylarsine-based Cu(I) complexes showing TADF mixed with fast phosphorescence: A speeding-up emission rate using arsine ligands. Dalton Trans. 2022, 51, 1048–1055. [Google Scholar] [CrossRef]
  39. Montalti, M.; Credi, A.; Prodi, L.; Gandolfi, M.T. Handbook of Photochemistry, 3rd ed.; CRC Press: Boca Raton, FL, USA, 2006; p. 633. [Google Scholar]
  40. Wallesch, M.; Volz, D.; Zink, D.M.; Schepers, U.; Nieger, M.; Baumann, T.; Bräse, S. Bright Coppertunities: Multinuclear CuI Complexes with N–P Ligands and Their Applications. Chem. Eur. J. 2014, 20, 6578–6590. [Google Scholar] [CrossRef]
  41. Zink, D.M.; Bächle, M.; Baumann, T.; Nieger, M.; Kühn, M.; Wang, C.; Klopper, W.; Monkowius, U.; Hofbeck, T.; Yersin, H.; et al. Synthesis, Structure, and Characterization of Dinuclear Copper(I) Halide Complexes with P^N Ligands Featuring Exciting Photoluminescence Properties. Inorg. Chem. 2013, 52, 2292–2305. [Google Scholar] [CrossRef]
  42. Cheng, G.; Zhou, D.; Monkowius, U.; Yersin, H. Fabrication of a Solution-Processed White Light Emitting Diode Containing a Single Dimeric Copper(I) Emitter Featuring Combined TADF and Phosphorescence. Micromachines 2021, 12, 1500. [Google Scholar] [CrossRef] [PubMed]
  43. Hofbeck, T.; Niehaus, T.A.; Fleck, M.; Monkowius, U.; Yersin, H. P∩N Bridged Cu(I) Dimers Featuring Both TADF and Phosphorescence. From Overview towards Detailed Case Study of the Excited Singlet and Triplet States. Molecules 2021, 26, 3415. [Google Scholar] [CrossRef] [PubMed]
  44. Hofbeck, T.; Monkowius, U.; Yersin, H. Highly Efficient Luminescence of Cu(I) Compounds: Thermally Activated Delayed Fluorescence Combined with Short-Lived Phosphorescence. J. Am. Chem. Soc. 2015, 137, 399–404. [Google Scholar] [CrossRef]
  45. Kobayashi, R.; Fujii, T.; Imoto, H.; Naka, K. Dinuclear Gold(I) Chloride Complexes with Diarsine Ligands. Eur. J. Inorg. Chem. 2021, 217–222. [Google Scholar] [CrossRef]
  46. Plajer, A.J.; Crusius, D.; Jethwa, R.B.; García-Romero, Á.; Bond, A.D.; García-Rodríguez, R.; Wright, D.S. Coordination chemistry of the bench-stable tris-2-pyridyl pnictogen ligands [E(6-Me-2-py)3] (E = As, As=O, Sb). Dalton Trans. 2021, 50, 2393–2402. [Google Scholar] [CrossRef] [PubMed]
  47. Gneuß, T.; Leitl, M.J.; Finger, L.H.; Yersin, H.; Sundermeyer, J. A new class of deep-blue emitting Cu(I) compounds—effects of counter ions on the emission behavior. Dalton Trans. 2015, 44, 20045–20055. [Google Scholar] [CrossRef]
  48. Baranov, A.; Berezin, A.S.; Samsonenko, D.G.; Mazur, A.; Tolstoy, P.; Plyusnin, V.F.; Kolesnikov, I.E.; Artem’ev, A. New Cu(I) halide complexes showing TADF combined with room temperature phosphorescence: The balance tuned by halogens. Dalton Trans. 2020, 49, 3155–3163. [Google Scholar] [CrossRef]
  49. Yang, L.; Powell, D.R.; Houser, R.P. Structural variation in copper(I) complexes with pyridylmethylamide ligands: Structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, 955–964. [Google Scholar] [CrossRef]
  50. Gneuß, T.; Leitl, M.J.; Finger, L.H.; Rau, N.; Yersin, H.; Sundermeyer, J. A new class of luminescent Cu(I) complexes with tripodal ligands—TADF emitters for the yellow to red color range. Dalton Trans. 2015, 44, 8506–8520. [Google Scholar] [CrossRef]
  51. Bondi, A. van der Waals Volumes and Radii. J. Phys. Chem. 1964, 68, 441–451. [Google Scholar] [CrossRef]
  52. Artem’ev, A.V.; Bagryanskaya, I.Y.; Doronina, E.P.; Tolstoy, P.M.; Gushchin, A.L.; Rakhmanova, M.I.; Ivanov, A.Y.; Suturina, A.O. A new family of clusters containing a silver-centered tetracapped [Ag@Ag43-P)4] tetrahedron, inscribed within a N12 icosahedron. Dalton Trans. 2017, 46, 12425–12429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Kolari, K.; Sahamies, J.; Kalenius, E.; Novikov, A.S.; Kukushkin, V.Y.; Haukka, M. Metallophilic interactions in polymeric group 11 thiols. Solid State Sci. 2016, 60, 92–98. [Google Scholar] [CrossRef]
  54. Andrusenko, E.V.; Kabin, E.V.; Novikov, A.S.; Bokach, N.A.; Starova, G.L.; Kukushkin, V.Y. Metal-mediated Generation of Triazapentadienate-terminated Di- and Trinuclear μ2-Pyrazolate NiII Species and Control of their Nuclearity. New J. Chem. 2017, 41, 316–325. [Google Scholar] [CrossRef]
  55. Bikbaeva, Z.M.; Novikov, A.S.; Suslonov, V.V.; Bokach, N.A.; Kukushkin, V.Y. Metal-mediated reactions between dialkylcyanamides and acetamidoxime generate unusual (nitrosoguanidinate)nickel(II) complexes. Dalton Trans. 2017, 46, 10090–10101. [Google Scholar] [CrossRef]
  56. Novikov, A.S. Strong metallophilic interactions in nickel coordination compounds. Inorg. Chim. Acta 2018, 483, 21–25. [Google Scholar] [CrossRef]
  57. Shmelev, N.Y.; Okubazghi, T.H.; Abramov, P.A.; Komarov, V.Y.; Rakhmanova, M.I.; Novikov, A.S.; Gushchin, A.L. Intramolecular aurophilic interactions in dinuclear gold(i) complexes with twisted bridging 2,2′-bipyridine ligands. Dalton Trans. 2021, 50, 12448–12456. [Google Scholar] [CrossRef]
  58. Grudova, M.V.; Novikov, A.S.; Kubasov, A.S.; Khrustalev, V.N.; Kirichuk, A.A.; Nenajdenko, V.G.; Tskhovrebov, A.G. Aurophilic Interactions in Cationic Three-Coordinate Gold(I) Bipyridyl/Isocyanide Complex. Crystals 2022, 12, 613. [Google Scholar] [CrossRef]
  59. Shmelev, N.Y.; Okubazghi, T.H.; Abramov, P.A.; Rakhmanova, M.I.; Novikov, A.S.; Sokolov, M.N.; Gushchin, A.L. Asymmetric Coordination Mode of Phenanthroline-like Ligands in Gold(I) Complexes: A Case of the Antichelate Effect. Cryst. Growth Des. 2022, 22, 3882–3895. [Google Scholar] [CrossRef]
  60. Espinosa, E.; Alkorta, I.; Elguero, J.; Molins, E. From weak to strong interactions: A comprehensive analysis of the topological and energetic properties of the electron density distribution involving X–HF–Y systems. J. Chem. Phys. 2002, 117, 5529–5542. [Google Scholar] [CrossRef]
  61. Johnson, E.R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A.J.; Yang, W. Revealing Noncovalent Interactions. J. Am. Chem. Soc. 2010, 132, 6498–6506. [Google Scholar] [CrossRef]
  62. Contreras-García, J.; Johnson, E.R.; Keinan, S.; Chaudret, R.; Piquemal, J.-P.; Beratan, D.N.; Yang, W. NCIPLOT: A Program for Plotting Noncovalent Interaction Regions. J. Chem. Theory Comput. 2011, 7, 625–632. [Google Scholar] [CrossRef] [PubMed]
  63. Kobayashi, A.; Hasegawa, T.; Yoshida, M.; Kato, M. Environmentally Friendly Mechanochemical Syntheses and Conversions of Highly Luminescent Cu(I) Dinuclear Complexes. Inorg. Chem. 2016, 55, 1978–1985. [Google Scholar] [CrossRef] [PubMed]
  64. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  65. Bruker Apex3 Software Suite: Apex3, SADABS-2016/2 and SAINT, version 2018.7-2; Bruker AXS Inc.: Madison, WI, USA, 2017.
  66. Frisch, M.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  67. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  68. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  69. Pritchard, B.P.; Altarawy, D.; Didier, B.; Gibson, T.D.; Windus, T.L. New basis set exchange: An open, up-to-date resource for the molecular sciences community. J. Chem. Inf. Model 2019, 59, 4814–4820. [Google Scholar] [CrossRef]
  70. Bauernschmitt, R.; Ahlrichs, R. Treatment of electronic excitations within the adiabatic approximation of time dependent density functional theory. Chem. Phys. Lett. 1996, 256, 454–464. [Google Scholar] [CrossRef]
  71. Van Caillie, C.; Amos, R.D. Geometric derivatives of excitation energies using SCF and DFT. Chem. Phys. Lett. 1999, 308, 249–255. [Google Scholar] [CrossRef]
  72. Scalmani, G.; Frisch, M.J.; Mennucci, B.; Tomasi, J.; Cammi, R.; Barone, V. Geometries and properties of excited states in the gas phase and in solution: Theory and application of a time-dependent density functional theory polarizable continuum model. J. Chem. Phys. 2006, 124, 094107. [Google Scholar] [CrossRef]
  73. Chai, J.D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom–atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [Google Scholar] [CrossRef]
  74. Barros, C.L.; De Oliveira, P.J.P.; Jorge, F.E.; Canal Neto, A.; Campos, M. Gaussian basis set of double zeta quality for atoms Rb through Xe: Application in non-relativistic and relativistic calculations of atomic and molecular properties. Mol. Phys. 2010, 108, 1965–1972. [Google Scholar] [CrossRef]
  75. Jorge, F.E.; Canal Neto, A.; Camiletti, G.G.; Machado, S.F. Contracted Gaussian basis sets for Douglas–Kroll–Hess calculations: Estimating scalar relativistic effects of some atomic and molecular properties. J. Chem. Phys. 2009, 130, 064108. [Google Scholar] [CrossRef] [PubMed]
  76. Neto, A.C.; Jorge, F.E. All-electron double zeta basis sets for the most fifth-row atoms: Application in DFT spectroscopic constant calculations. Chem. Phys. Lett. 2013, 582, 158–162. [Google Scholar] [CrossRef]
  77. De Berrêdo, R.C.; Jorge, F.E. All-electron double zeta basis sets for platinum: Estimating scalar relativistic effects on platinum (II) anticancer drugs. J. Mol. Struct. Theochem. 2010, 961, 107–112. [Google Scholar] [CrossRef]
  78. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Reactions of Py3As with CuBr and CuI.
Scheme 1. Reactions of Py3As with CuBr and CuI.
Molecules 27 06059 sch001
Figure 1. (a) MeCN-driven dimerization of 1, and direct interaction of Py3As with CuI in MeCN; (b) associated changing in PXRD patterns: experimental diffractograms of starting complex 1 (down curve), product 1a (middle curve), and simulated PXRD pattern for [Cu2(Py3P)2I2] (top curve).
Figure 1. (a) MeCN-driven dimerization of 1, and direct interaction of Py3As with CuI in MeCN; (b) associated changing in PXRD patterns: experimental diffractograms of starting complex 1 (down curve), product 1a (middle curve), and simulated PXRD pattern for [Cu2(Py3P)2I2] (top curve).
Molecules 27 06059 g001
Scheme 2. Reactivity of Py3As and Py2PhAs in reactions with AgClO4.
Scheme 2. Reactivity of Py3As and Py2PhAs in reactions with AgClO4.
Molecules 27 06059 sch002
Figure 2. X-ray derived structures of 1·CH2Cl2, 2, 3·3H2O (one of three independent parts) and 4·CH3CN. The aromatic H atoms, solvate molecules and counterions are omitted for clarity.
Figure 2. X-ray derived structures of 1·CH2Cl2, 2, 3·3H2O (one of three independent parts) and 4·CH3CN. The aromatic H atoms, solvate molecules and counterions are omitted for clarity.
Molecules 27 06059 g002
Figure 3. Frontier orbitals of 1 and 2 calculated at PBE0/def2TZVP level (isosurface = 0.045).
Figure 3. Frontier orbitals of 1 and 2 calculated at PBE0/def2TZVP level (isosurface = 0.045).
Molecules 27 06059 g003
Figure 4. Contour line diagram of the Laplacian of electron density distribution ∇2ρ(r), bond paths, and selected zero-flux surfaces (left panel), visualization of electron localization function (ELF, center panel) and reduced density gradient (RDG, right panel) analyses for metallophilic interactions in complex 4. Bond critical points (3, –1) are shown in blue, nuclear critical points (3, –3)—in pale brown, ring critical points (3, +1)—in orange, bond paths are shown as pale brown lines, length units – Å, and the color scale for the ELF and RDG maps is presented in a.u.
Figure 4. Contour line diagram of the Laplacian of electron density distribution ∇2ρ(r), bond paths, and selected zero-flux surfaces (left panel), visualization of electron localization function (ELF, center panel) and reduced density gradient (RDG, right panel) analyses for metallophilic interactions in complex 4. Bond critical points (3, –1) are shown in blue, nuclear critical points (3, –3)—in pale brown, ring critical points (3, +1)—in orange, bond paths are shown as pale brown lines, length units – Å, and the color scale for the ELF and RDG maps is presented in a.u.
Molecules 27 06059 g004
Figure 5. Emission (a) and excitation (b) spectra of 1·CH2Cl2, 2 and 3·3H2O at 300 K. The emission spectra were recorded at λex = 390 nm (for 1a, 2) and 500 nm (1·CH2Cl2).
Figure 5. Emission (a) and excitation (b) spectra of 1·CH2Cl2, 2 and 3·3H2O at 300 K. The emission spectra were recorded at λex = 390 nm (for 1a, 2) and 500 nm (1·CH2Cl2).
Molecules 27 06059 g005
Figure 6. Temperature dependent emission spectra of 1·CH2Cl2 (a) and 2 (b) recorded at λex = 500 and 380 nm, respectively.
Figure 6. Temperature dependent emission spectra of 1·CH2Cl2 (a) and 2 (b) recorded at λex = 500 and 380 nm, respectively.
Molecules 27 06059 g006
Table 1. Selected bond lengths (Å) and angles (°) for 1·CH2Cl2, 2, 3·3H2O and 4·CH3CN.
Table 1. Selected bond lengths (Å) and angles (°) for 1·CH2Cl2, 2, 3·3H2O and 4·CH3CN.
1·CH2Cl22
Cu–N22.046(3)Cu–As2.3207(5)
Cu–N12.042(2)Cu–N1′2.051(3)
Cu–N1′2.042(2)Cu–N2′2.066(2)
Cu–I2.5003(6)Cu–Br2.4242(6)
Symmetry code: (′) x, y, z–1/2.Symmetry code: (′) 1–x, 1–y, 1–z.
3·3H2O4·CH3CN
Ag1∙∙∙Ag23.2413(14)Ag∙∙∙Ag′3.2206(4)
Ag1∙∙∙Ag33.2360(14)Ag–As′2.4746(3)
Ag1∙∙∙Ag43.2264(14)Ag–N12.347(2)
Ag1∙∙∙Ag53.3016(14)Ag–N22.361(2)
Ag1–As12.5780(15)Ag–N32.371(3)
Ag1–As22.5869(16)Symmetry code: (′) 1–x, 2–y, 1–z.
Ag1–As32.5900(16)
Ag1–As42.6041(15)
Ag–N2.260(13)–2.343(11)
Table 2. Values of the density of all electrons – ρ(r), Laplacian of electron density – ∇2ρ(r) and appropriate λ2 eigenvalues, energy density – Hb, potential energy density – V(r), Lagrangian kinetic energy – G(r), and electron localization function – ELF at the bond critical points (3, −1), corresponding to Ag(I)···Ag(I) interactions in 3 and 4.
Table 2. Values of the density of all electrons – ρ(r), Laplacian of electron density – ∇2ρ(r) and appropriate λ2 eigenvalues, energy density – Hb, potential energy density – V(r), Lagrangian kinetic energy – G(r), and electron localization function – ELF at the bond critical points (3, −1), corresponding to Ag(I)···Ag(I) interactions in 3 and 4.
Ag···Ag Contactρ(r)2ρ(r)λ2HbV(r)G(r)ELF
Complex 3
3.241 Å0.019 a.u.0.030 a.u.−0.019 a.u.−0.003 a.u.−0.013 a.u.0.010 a.u.0.125 a.u.
3.236 Å0.019 a.u.0.029 a.u.−0.019 a.u.−0.003 a.u.−0.013 a.u.0.010 a.u.0.129 a.u.
3.226 Å0.020 a.u.0.030 a.u.−0.020 a.u.−0.003 a.u.−0.014 a.u.0.011 a.u.0.129 a.u.
3.302 Å0.017 a.u.0.029 a.u.−0.017 a.u.−0.003 a.u.−0.012 a.u.0.009 a.u.0.112 a.u.
3.249 Å0.019 a.u.0.029 a.u.−0.019 a.u.−0.003 a.u.−0.013 a.u.0.010 a.u.0.124 a.u.
3.252 Å0.019 a.u.0.029 a.u.−0.019 a.u.−0.003 a.u.−0.013 a.u.0.010 a.u.0.124 a.u.
3.281 Å0.018 a.u.0.029 a.u.−0.018 a.u.−0.002 a.u.−0.012 a.u.0.010 a.u.0.117 a.u.
3.209 Å0.020 a.u.0.030 a.u.−0.020 a.u.−0.003 a.u.−0.014 a.u.0.011 a.u.0.136 a.u.
3.222 Å0.020 a.u.0.029 a.u.−0.020 a.u.−0.003 a.u.−0.014 a.u.0.011 a.u.0.134 a.u.
3.214 Å0.020 a.u.0.030 a.u.−0.020 a.u.−0.003 a.u.−0.014 a.u.0.011 a.u.0.134 a.u.
3.247 Å0.019 a.u.0.030 a.u.−0.019 a.u.−0.003 a.u.−0.013 a.u.0.010 a.u.0.125 a.u.
3.313 Å0.017 a.u.0.029 a.u.−0.017 a.u.−0.002 a.u.−0.011 a.u.0.009 a.u.0.110 a.u.
Complex 4
3.221 Å0.021 a.u.0.028 a.u.−0.021 a.u.−0.003 a.u.-0.014 a.u.0.011 a.u.0.156 a.u.
Table 3. Luminescent characteristics of Cu(I) complexes 1, 1a and 2.
Table 3. Luminescent characteristics of Cu(I) complexes 1, 1a and 2.
Complexλem, nmPL Lifetime, μsPLQY, %
[298 K]
298 K77 K298 K77 K
1·CH2Cl26056251.92510 a
1a511-0.8-14 b
25105200.92312 b
a λex = 500 nm, b λex = 390 nm.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Demyanov, Y.V.; Sadykov, E.H.; Rakhmanova, M.I.; Novikov, A.S.; Bagryanskaya, I.Y.; Artem’ev, A.V. Tris(2-Pyridyl)Arsine as a New Platform for Design of Luminescent Cu(I) and Ag(I) Complexes. Molecules 2022, 27, 6059. https://doi.org/10.3390/molecules27186059

AMA Style

Demyanov YV, Sadykov EH, Rakhmanova MI, Novikov AS, Bagryanskaya IY, Artem’ev AV. Tris(2-Pyridyl)Arsine as a New Platform for Design of Luminescent Cu(I) and Ag(I) Complexes. Molecules. 2022; 27(18):6059. https://doi.org/10.3390/molecules27186059

Chicago/Turabian Style

Demyanov, Yan V., Evgeniy H. Sadykov, Marianna I. Rakhmanova, Alexander S. Novikov, Irina Yu. Bagryanskaya, and Alexander V. Artem’ev. 2022. "Tris(2-Pyridyl)Arsine as a New Platform for Design of Luminescent Cu(I) and Ag(I) Complexes" Molecules 27, no. 18: 6059. https://doi.org/10.3390/molecules27186059

Article Metrics

Back to TopTop