Next Article in Journal
Unusual Phase Behaviour for Organo-Halide Perovskite Nanoparticles Synthesized via Reverse Micelle Templating
Previous Article in Journal
Progress and Perspectives of Conducting Metal–Organic Frameworks for Electrochemical Energy Storage and Conversion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization and Photophysical Properties of Mixed Ligand (η3-Allyl)palladium(II) Complexes with N,N’Aromatic Diimines

by
Antonia Garypidou
1,
Konstantinos Ypsilantis
1,
Evaggelia Sifnaiou
1,
Maria Manthou
1,
Dimitris Thomos
1,
John C. Plakatouras
1,2,
Theodoros Tsolis
1 and
Achilleas Garoufis
1,2,*
1
Department of Chemistry, University of Ioannina, GR-45110 Ioannina, Greece
2
Institute of Materials Science and Computing, University Research Center of Ioannina (URCI), GR-45110 Ioannina, Greece
*
Author to whom correspondence should be addressed.
Chemistry 2023, 5(4), 2476-2489; https://doi.org/10.3390/chemistry5040162
Submission received: 23 October 2023 / Revised: 9 November 2023 / Accepted: 9 November 2023 / Published: 12 November 2023
(This article belongs to the Section Inorganic and Solid State Chemistry)

Abstract

:
Palladium(II) complexes of the general formula [Pd(η3-C3H5)(L)](PF6), where L is 4,7-diphenyl-1,10-phenanthroline (1), 2,9-dimethyl-1,10-phenanthroline (2), 2,9-dimethyl-4,7-diphenyl-1,10-phenanthroline (3), 5-methyl-1,10-phenanthroline (4), 3,4,7,8-tetramethyl-1,10-phenanthroline (5), and 2-(2′-pyridyl) quinoxaline (6), were synthesized and characterized using high-resolution ESI-MS, NMR techniques and, in the case of (6), single-crystal X-ray diffraction methods. In addition, their photophysical properties were investigated. Complexes (1)–(6) were emitted in the greenish-blue region, with those containing methyl-substituted phenanthrolines having the higher quantum yield (≈14%) in the solid state.

1. Introduction

In recent years, much research has focused on the synthesis of luminescent compounds, owing to their application in various fields, such as lightemitting devices (OLEDs) [1,2,3,4], image biomarkers [5,6,7,8], photodynamic therapy [9] and others [10,11,12,13].
Organometallic compounds have been studied extensively due to their advantages over their organic ligands, primarily stemming from the metal center [14]. The higher the spin-orbit coupling (SOC) constant of the transition metal, the greater the potential for intersystem crossing, leading to enhanced exciton harvesting in the T1 triplet state and, consequently, increased emission intensity [15]. Notably, compounds containing metal ions from the second and third transition rows tend to exhibit greater emissivity due to the higher overlap of their 4d or 5d orbitals with the orbitals of the ligand [16]. Among the well-established efficient triplet emitters are complexes of platinum [17,18,19,20,21], ruthenium [22,23] and iridium [23,24]. However, the recent literature reports studies involving palladium(II) complexes with luminescent properties [14]. The current data are not particularly encouraging, as only a few Pd(II) complexes exhibit luminescence. This is attributed to both the weak SOC effect and the presence of thermally accessible metal-centered (MC) d-d* states, resulting in distortion of the excited state structure and efficient non-radiative decay [25,26]. Notably, orthometalated Pd(II) complexes have received significant attention, with several demonstrating luminescent properties, where the presence of the metal enhances the emission in comparison with that of the free ligand [27,28,29,30]. Consequently, the exploration of the photophysical properties of Pd(II) complexes remains a relatively unexplored, yet intriguing, field of research.
In addition to the transition metal, the nature of ligands plays an important role in the photophysical properties of organometallic compounds. Complexes containing chelating ligands with carbon and nitrogen donor atoms have been extensively reported in the relevant literature [14]. Specifically, complexes of 1,10-phenanthroline derivatives have been reported to exhibit attractive luminescence features [31]. Recently, de Franca et al. synthesized a complex of Pd(II) with 1,10-phenanthroline, achieving a remarkable quantum yield of over 50% [32]. Furthermore, a binuclear complex of Pd(II) with a bis-N,N’ chelating bridging perylene diimide ditopic ligand, while displaying a low quantum yield, exhibited a higher quantum yield than the free ligand [33].
Although allyl is a common anionic ligand in organometallic chemistry, mixed ligand complexes with phenanthroline-lik e N,N’ chelating ligands have rarely been reported in the literature [34,35,36,37,38]. In general, complexes with the η3-allyl-Pd(II) moiety have been investigated for their applications in catalytic processes [39,40,41] and their cytotoxic activity [42,43,44,45]. However, only a few reports deal with the examination of such type complexes regarding their photophysical feautures. Mayoral et al. reported that two compounds of [Pd(η3-allyl)] with β-diketone derivatives exhibited luminescent properties [46]. More recently, allyl palladium complexes bearing imidazo[1,5-a]pyridino-3-ylidene and dipyridoimidazolinylidene ligands have been synthesized and studied for their photophysical properties. The results demonstrated that these organometallic compounds exhibit luminescent characteristics, with the Φ value reaching up to 0.20 [47].
In this study, we synthesized and characterized six novel allyl palladium organometallic complexes (1)–(5), containing substituted 1,10-phenanthrolines and (6) containing 2-(2’-pyridyl)quinoxaline as N,N’ chelating ligands (Scheme 1). In addition, we investigated their photophysical properties and examined how the nature of the chelating ligand affects the fluorescence of the complexes.

2. Experimental

2.1. Materials and Methods

All solvents and chemicals were analytical grade and were used without further purification. Sodium tetrachloridopalladate(II) (99.9%) and 3-chloro-1-propene were purchased from Alfa Aesar. 2-acetyl-pyridine (99.0%) was purchased from Merck Chemical Company. The ligands 4,7-Diphenyl-1,10-phenanthroline (bphen), 2,9-Dimethyl-1,10-phenanthroline (ncp), 2,9-Dimethyl-4,7-diphenyl-1,10-phenanthroline (bcp), 5-Methyl-1,10-phenanthroline (mphen), and 3,4,7,8-Tetramethyl-1,10-phenanthroline (tmphen) were purchased from Sigma Aldrich. 2-(2′-pyridyl)quinoxaline (pqx) was synthesized according to the literature [48]. High-resolution electrospray ionization mass spectra (HR-ESI-MS) were obtained on a Thermo Scientific, LTQ Orbitrap XL™ system. C, H, and N determinations were performed on a Perkin-Elmer 2400 Series II analyzer. 1H NMR spectra were recorded on Bruker Avance NEO and Avance II spectrometers operating at proton frequencies of 400.13 and 500.13 MHz, while 13C at frequency 125 MHz, and processed using Topspin 4.07 (Bruker Analytik GmbH). Two-dimensional COSY, TOCSY, and NOESY spectra were recorded following standard Bruker procedures. NOESY spectra were recorded in mixing times in the range from 600 to 800 ms. The UV-Vis spectra of the complexes were recorded on Agilent Cary 60 UV-Vis Spectrophotometer with xenon source lamp, in solution of CH2Cl2 and in solid state. The fluorescence emission study was carried out using a Jasco FP-8300 fluorometer equipped with a xenon lamp source. The luminescence quantum yields of the solutions (C = 1 × 10−4 M) were determined using the equation Qs = Qr(Ar/As)(Es/Er)(ns/nr)2, using [Ru(bpy)3]Cl2 in air-equilibrated water solution as a reference standard (Qr = 0.04). ‘A’ stands for the absorbance of the solution, ‘E’ for the integrated fluorescence intensity of the emitted light, ‘n’ is the refractive index of the solvents and subscripts ‘r’ and ‘s’ correspond to the reference and the sample, respectively. Using the equation Q = S2/S0 − S1, the quantum yields of the solid state of the complexes were calculated. S2 denotes the integrated emission intensity of the sample and S0, S1 stand for the excitation intensities of the standard and the sample, respectively.

2.2. Synthesis of the Complexes

Complexes (1)–(6) were synthesized in a similar manner. In a typical procedure, 0.20 mmol of the ligand was added to a stirred solution of CH2Cl2 (10 mL) containing 0.05 mmol of the dimer [Pd(η3-C3H5)Cl]2 at room temperature, according to the literature [49] (Figures S1 and S2, in the Supplementary Materials). The reaction mixture was stirred for 2 h, and a pale-yellow precipitate appeared. It was collected with centrifugation, washed twice with 5 mL of CH2Cl2, and dissolved in 10 mL of acetone. An aqueous solution (20 mL) containing 1 mmol of NH4PF6 was added slowly, and the mixture was kept at 4 °C overnight. A microcrystalline yellow precipitate was collected, washed with 5 mL of H2O, and dried under vacuum.
[Pd(η3-C3H5)(bphen)]PF6 (1): Yield: 75%. C27H21F6N2PPd: calc.% C, 51.9; H, 3.39; N, 4.48; found C, 51.6; H, 3.41; N, 4.49. HR-ESI-MS, positive (m/z): found 479.0743 calc. 479.0743 for [C27H21N2Pd]+. 1H NMR: (500 MHz, 298 K, acetone-d6, δ in ppm), H2/9 9.57, (d, 2H), H3/8 8.18, (d, 2H), H5/6 8.27, (s, 2H), Ha/b/c 7.76, (m, 10H), Ha1″/c1″ 4.78, (d, 2H), Hb″ 6.34, (m, 1H), Ha2″/c2″ 3.90, (d, 2H).
[Pd(η3-C3H5)(ncp)]PF6 (2): Yield: 82%. C17F6H17N2PPd: calc.% C, 40.78; H, 3.42; N, 5.59; found C, 40.82; H, 3.40; N, 5.57. HR-ESI-MS, positive (m/z): found 355.0425 calc. 355.0421 for [C17H17N2Pd]+. 1H NMR: (500 MHz, 298 K, acetone-d6, δ in ppm), H4/7 8.76, (d, 2H), H3/8 8.04, (d, 2H), H5/6 8.15, (d, 2H), HCH3 3.10, (s, 6H), Ha1″/c1″ 4.90, (d, 2H), Hb″ 5.91, (m, 1H), Ha2″/c2″ 3.80, (d, 2H).
[Pd(η3-C3H5)(bcp)]PF6 (3): Yield: 87%. C29H25F6N2PPd: calc.% C, 53.35; H, 3.86; N, 4.29; found C, 53.30; H, 3.89; N, 4.31. HR-ESI-MS, positive (m/z): found 507.1065 calc. 507.1047 for [C29H25N2Pd]+. 1H NMR: (500 MHz, 298 K, acetone-d6, δ in ppm), H3/8 8.08, (s, 2H), H5/6 8.08, (s, 2H), Haa′/bb′/cc′ 7.69, (m, 10H), Ha1″/c1″ 4.99, (d, 2H), Hb″ 6.03, (m, 1H), Ha2″/c2″ 3.87, (d, 2H), HCH3 3.24, (s, 6H).
[Pd(η3-C3H5)(mphen)]PF6 (4): Yield: 91%. C16H15F6N2PPd: calc.% C, 39.49; H, 3.11; N, 5.76; found C, 39.53; H, 3.09; N, 5.74. HR-ESI-MS, positive (m/z): found 341.0274 calc. 341.0265 for [C16H15N2Pd]+. 1H NMR: (500 MHz, 298 K, acetone-d6, δ in ppm), H3/8 8.14, (m, 2H), H2 9.45, (d, 1H), H4 9.09, (d, 1H), H6 8.20, (t, 1H), H7 8.88, (d, 1H), H9 9.36, (d, 1H), Ha1″/c1″ 4.67, (d, 2H), Hb″ 6.24, (m, 1H), Ha2″/c2″ 3.70, (d, 2H), HCH3 2.94 (s, 3H). 13C NMR: (125 MHz, 298 K, acetone-d6, δ in ppm) C9/2 155.3/154.8, C11/13 146.0/146.1, C7/4 140.1/138.1, C5 136.7, C12/14 131.6/131.2, C3/8 127.5/127.4, C6 127.2, C(CH3) 18.8, Cb 121.2, Ca/c 63.3/63.0.
[Pd(η3-C3H5)(tmphen)]PF6 (5): Yield: 79%. C19H21 F6N2PPd: calc.% C, 43.16; H, 4.00; N, 5.30; found C, 43.10; H, 3.13; N, 5.76. HR-ESI-MS, positive (m/z): found 383.0742 calc. 383.0734 for [C19H21N2Pd]+. 1H NMR: (500 MHz, 298 K, acetone-d6, δ in ppm), H2/9 9.29, (s, 2H), H5/6 8.47, (s, 2H), Ha1″/c1″ 4.63, (d, 2H), Hb″ 6.19, (m, 1H), Ha2″/c2″ 3.73, (d, 2H), HCH3 2.91 (s, 6H), HCH3 2.68 (s, 6H). 13C NMR: (125 MHz, 298 K, acetone-d6, δ in ppm), C2/9 156.1, C11/13 148.7, C5/6 124.5, C3/8 136.1, C4/7 129.7, C12/14 145.3, C(3/8) CH3 17.5, C(4/7) CH3 15.3, Cb 120.7, Ca/c 62.7.
[Pd(η3-C3H5)(pqx)]PF6 (6): Yield: 82%. C16H14 F6N3PPd: calc.% C, 40.66; H, 4.17; N, 7.90; found C, 40.70; H, 4.16; N, 7.87. HR-ESI-MS, positive (m/z): found. 354.0224 calc. 354.0217 for [C16H14N3Pd]+. 1HNMR: (500 MHz, 298 K, acetone-d6, δ in ppm), H5 8.55, (d, 1H), H4′ 8.61(t, 1H), H6/7 8.21, (m, 2H), H3 10.25, (s, 1H), H6′ 9.30, (d,1H), H3′9.16, (d, 1H), H8 8.40, (d, 1H), H5′ 8.06, (t, 1H), Ha1″/c1″ 5.02, (d, 2H), Hb″ 6.30, (m, 1H), Ha2″/c2″ 4.08, (d, 2H). Suitable bright-yellow crystals of (6), for X-ray analysis, were grown from an aqueous acetone solution containing dissolved solids initially isolated from the reaction mixture in the presence of NH4PF6 in the fridge. However, crystals of (6a) were obtained from the reaction mixture layered with diethyl ether and characterized with 1H and 13C NMR. Pd(η3-C3H5)(pqx)] [Pd(η3-C3H5)Cl2] (6a): 1HNMR: (500 MHz, 298 K, methanol-d4, δ in ppm), [Pd(η3-C3H5)(pqx)]+: H5 8.43, (d, 1H), H4′ 8.44(t, 1H), H6/7 8.13, (m, 2H), H3 10.08, (s, 1H), H6′ 8.94, (d, 1H), H3′ 9.12, (d, 1H), H8 8.39, (d, 1H), H5′ 7.90, (t, 1H), Ha1″/c1″ 4.87, (d, 2H), Hb″ 6.18, (m, 1H), Ha2″/c2″ 3.93, (d, 2H). [Pd(η3-C3H5)Cl2]: Ha1″/c1″ 4.05, (d, 2H), Hb″ 5.58, (m, 1H), Ha2″/c2″ 3.03, (d, 2H). 13C NMR: (125 MHz, 298 K, methanol-d4, δ in ppm) [Pd(η3-C3H5)(pqx)]+: C2 151.7, C3 146.0, C2′ 155.3, C3′ 126.7, C4′ 130.1, C5′ 142.3, C6′ 156.0, C9 146.0, C10 142.3, C5/8 131.6/131.1, C6/7 134.6/134.1, Cb 120.5, Ca/c 62.6. [Pd(η3-C3H5)Cl2]: Cb 112.8, Ca/c 65.9.

2.3. Crystal Structure Determinations

Suitable crystals were glued to a thin glass fiber with cyanoacrylate (super glue) adhesive and placed on the goniometer head. Diffraction data were collected on a Bruker D8 Quest Eco diffractometer, equipped with a Photon II detector and a TRIUMPH (curved graphite) monochromator utilizing Mo Ka radiation (λ = 0.71073 Å) using the APEX 3 software package [50]. The collected frames were integrated with the Bruker SAINT software using a wide-frame algorithm. Data were corrected for absorption effects using the Multi-Scan method (SADABS) [51]. The structures were solved using the Bruker SHELXT Software Package and refined by full-matrix least squares techniques on F 2 (SHELXL 2018/3) [52] via the ShelXle interface [53]. The non-H atoms were treated anisotropically, whereas the organic H atoms were placed in calculated, ideal positions and refined as riding on their respective carbon atoms. PLATON [54] was used for geometric calculations, and X-Seed [55] for molecular graphics. Details on data collection and refinement are presented in Table 1. Full details on the structures can be found in the CIF files in the ESI. CCDC 2301621 and 2301622 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.

3. Results and Discussion

3.1. Synthesis and Characterization of Complexes (1)–(6)

Complexes (1)–(6) were formed from the reaction of the dimer [Pd(η3-C3H5)Cl]2 with the chelating ligand N,N’, in a non-coordinative solvent, such as CH2Cl2 and subsequently isolated as [PF6] salts, according to Scheme 2 as follows:
Τhe 1H NMR spectra of complexes (1)–(6) in acetone exhibit, in addition to signals corresponding to the aromatic protons, three signals attributed to the η3-allyl protons. These signals include a one-proton multiplet, ranging from 5.91 to 6.34 ppm, assigned to Hb″, a two-proton doublet ranging from 4.67 to 5.02 ppm with a 3J coupling constant of 6.5–7.5 Hz, assigned to Ha2″/Hc2″ syn to Hb″ and another two-proton doublet, ranging from 3.70 to 4.08 ppm with 3J = 13.5–14.5 Hz, assigned to Ha1″/Hc1″ anti to Hb″ (Figure 1).
The distinction between the doublets of Ha2′’/Hc2″ and Ha1″/Hc1″ was made based on the general principle that the 3J value for trans-protons in a double bond C-C is lower than that for the corresponding cis-protons [47]. Moreover, in the 600 ms NOESY spectrum of (2), we observe a cross-peak, opposite to diagonal, between the ncp methyl groups and the signal assigned to Ha1″/Hc1″. This indicates a proximity between these protons. Also, all the NOE connectivities between the η3-allyl protons were observed (Figure 2).
The chemical shifts of Hb″ appear to be affected by the intrinsic basicity of the chelated phenanthroline. In general, the methyl-substituted phenanthrolines are stronger bases than the phenanthroline itself, which, however, is stronger base than its phenyl-substituted derivatives. Thus, the deshielding effect on Hb’’ is more pronounced as the pKa of the chelated phenanthroline decreases. For instance, 2,9-dimethyl-1,10-phenanthroline [56] (pKa = 6.15 and δ = 5.91 ppm); 5-methyl-1,10-phenanthroline [56] (pKa = 5.2 and δ = 6.24 ppm) > 1,10-phenanthroline [57] (pKa = 4.92 and δ = 6.30 ppm) > 4,7-diphenyl-1,10-phenanthroline [58] (pKa = 4.55 and δ = 6.34 ppm). With respect to the chelated ligand, the protons at neighboring sites of the coordinated nitrogen atoms shifted downfield, as expected, in the range of 0.26 to 0.47 ppm (Figures S3–S8, in the Supplementary Materials). The 13C NMR spectra of the complexes (4), (5) and (6a) are presented in Figures S15–S17 (in the Supplementary Materials).
In the HR-ESI-MS of complexes (1)–(6), only one cluster peak assignable to the positive single-charged cation [Pd(η3-C3H5)(L)]+ was observed (S9–S14).

3.2. Crystal Structures of [Pd(η3-C3H5)(pqx)]PF6 (6) and [Pd(η3-C3H5)(pqx)][Pd(η3-C3H5)Cl2] (6a)

Suitable crystals for X-ray diffraction studies of (6) were grown from an aqueous acetone solution containing dissolved solids initially isolated from the reaction mixture in presence of NH4PF6 in the fridge, while crystals of (6a) were isolated by layering the reaction mixture with diethyl ether. The same procedure to obtain crystals for complexes (1)–(5) was followed without success. Selected bond lengths and angles for both compounds are presented in Table 2.
Compound (6) crystallizes in the orthorhombic space group Pnma with its asymmetric unit containing halves of the cation [Pd(η3-C3H5)(pqx)]+ and the anion PF6 lying on a mirror plane. All the atoms of the cation, except for the central allyl carbon atom, are on that mirror plane and its structure is presented in Figure 3. The palladium atom is formally five-coordinated, since the allyl ligand utilizes all its carbon atoms for coordination, adopting a η3-binding mode. Though, ignoring the coordination of C(2A), the coordination sphere of Pd(1) can be considered as distorted square planar. The planarity of the coordination sphere is imposed by the above-mentioned mirror plane, but the cis coordination sphere angles deviate significantly from 90°. This is due to the small bite angle of the pqx ligand and to the even smaller C(1A)–Pd(1)–C(3A) allyl angle. The Pd–N and Pd–C bond distances agree with literature values for similar complexes. (See, for example, Refs [59,60,61]).
There is a small difference between the Pd–N bond distances, reflecting the expected difference based on the basicity of pyridine and quinoxaline nitrogen atoms. Although the C(1A) and C(3A) atoms of the allyl group are on the mirror plane hosting the coordination sphere, C(2A) is disordered in two positions, above and below the plane, as imposed by the symmetry. The dihedral angle between the coordination sphere plane and the allyl plane is 67.2°. The geometrical characteristics of the allyl group deviate slightly from the ideal values, with the mean C–C distance and C–C–C angle being respectively 1.38 Å and 119.4° (ideal values: ca. 1.36 Å and 120°, respectively). The small difference between the Pd–C(1A) and Pd–C(3A) bond distances is related to the different trans influence of the N atoms of the heterocyclic ligand.
Symmetry-related (−x+1, y−1/2, −z+1) complex cations interact with π–π stacking interactions via the quinoxaline moiety, forming positively charged columns parallel to the b-axis of the unit cell. The characteristics of the interaction are: centroid distance, 3.90; least-square plane distance, 3.43; centroid offset, 1.86 Å. The counter anions are located between the columns interacting with non-conventional C–H ··· F H-bonds. A packing diagram is shown in Figure S18.
In absence of other counter anions (i.e., PF6) from the reaction system, compound (6a) can be isolated. It crystallizes in the monoclinic space group P21/c and consists of ionic pairs [Pd(η3-C3H5)(pqx)][Pd(η3-C3H5)Cl2] interacting with each other in the lattice via stacking interactions and non-conventional C–H ··· Cl H-bonds. A view of the ionic pair is shown in Figure 4. Despite the lowered accuracy of the bond lengths and angles in the structure of (6a), due to the disordered allyl groups, the structural characteristics, and the observations of the differences between the Pd–N and Pd–C bond lengths and angles are very similar to (6) and will not be described in detail.
To our surprise, ionic pairs containing palladiumallyl complexes in both cations and anions are scarce in the literature. Compound (6a) is the ninth member of this family[62] (CSD codes: AQICOS, AQICUY, BEHMIK, HAJKAE, IZIXIB, OZUHUN, OZUJAV, QOHPUZ). The structural characteristics of the anion are consistent within the observed values. (See for example, [63,64,65]).
In the lattice, the cations are arranged in a similar fashion as in (6), forming columns, but the stacking interactions of symmetry related (−x+2, −y+1, −z+1) quinoxaline moieties lead to the formation of stacked dimers. The characteristics of the interaction are centroid distance, 3.77; least-square plane distance, 3.39; centroid offset, 1.65 Å. The counter anions are located between the columns interacting with non-conventional C–H ··· Cl H-bonds. A packing diagram is shown in Figure S19.

3.3. Photophysical Properties of Complexes (1)–(6)

The photophysical data of complexes (1)–(6) are summarized in Table 3. Solid state absorption and emission spectra are presented in Figure 5, while the corresponding solution spectra are presented in Figure 6.
The solid-state UV-Vis spectra of compounds (1)–(6) (Figure 5) exhibit very similar bands below 350 nm, which are attributed to intraligand π → π* transitions. The low energy absorption bands ranging from 352 to 420 nm probably arise from metal-to-ligand charge transfer (MLCT) transitions involving the N,N’-chelated ligands and the η3- allyl group. Upon excitation of the complexes at 400 nm, a greenish-blue emission centered at approximately 480 and 530 nm is observed for all the compounds. The quantum yield (QY) varies from 3 to 14%, depending on the nature of the chelated ligand. The complexes containing methyl-substituted phenanthrolines, (2), (4), and (5), exhibit significantly higher QYs compared to those with phenyl substitutions, (1) and (3). However, in diluted solutions of CH2Cl2, lower QYs ranging from 1 to 2.8% were observed for all the complexes, (1)–(5). These results suggested that in solid-state the higher QYs arise from packing interactions, such as π-stacking, which are mainly favored by the presence of small substituents, like the methyl groups. In every case, the complexes exhibit higher QYs compared to that of the phenanthroline itself [31]. In similar platinum complexes with substituted phenanthrolines, the emission values solid state complexes are quite close to those of palladium, with the exception of the complex with bathophenanthroline, where a maximum of 620 nm is observed. However, the quantum yields are very low, with some cases even falling below 1% [66].
The UV/Vis spectra of (1)–(6) in CH2Cl2 show similarities among complexes (1)–(5), involving strong intraligand π → π* transitions below 350 nm and a weak intensity band (ε = 20−40 × 104 M−1cm−1) at 348–362 nm, which is assigned to MLCT. In the case of Pt complexes with substituted phenanthrolines, the absorption spectra in CH2Cl2, show a spectral red-shift in comparison with the corresponding Pd ones, moving the λmax of MLCT towards the visible region. Also, a similar phenomenon was observed in the emission spectra of Pt complex with bathophenanthroline (λem = 600 nm) [66]. However, the spectrum of (6) significantly differs from those of (1)–(5), showing two similar intense bands (ε254 = 20 × 104 M−1cm−1, ε364 = 16 × 104 M−1cm−1) at 254 and 364 nm, probably reflecting the lower rigidity of pqx compared to the phenanthroline ring.
The emission spectra of complexes (1)–(4) and (6) are similar, with λem ranging from 433–437 nm and low QYs (0.15–1.6%). However, complex (5) exhibits an almost green emission, having the highest QY (2.8%) among all the studied complexes.

4. Conclusions

The reaction between the dimer [Pd(η3-C3H5)Cl]2 and N,N’ aromatic diimines (L) yielded mixed-ligand cationic complexes of the general formula [Pd(η3-C3H5)(L)]+. The complexes were isolated as [PF6] salts and characterized using high-resolution ESI-MS, NMR spectroscopic techniques and X-ray single-crystal diffraction methods. It was observed that the chemical shift of the coordinated η3-allylHb’’ was affected by the intrinsic basicity of the chelated phenanthroline, reflecting the trans influence of the chelated ligand on the η3-coordinated allyl group. All the synthesized complexes emitted in the greenish-blue region when exited at λexc = 400 nm, with those containing methyl-substituted phenanthrolines having the higher quantum yield (≈14%) in the solid state.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry5040162/s1: Figures S1 and S2: 1H and 13C NMR spectra of the complex [Pd(η3-allyl)Cl]2 in CDCl3 at 298 K. Figures S3–S8: 1H NMR spectra of complexes (1)–(6). Figures S9–S14: HR-ESI-MS of complexes (1)–(6). Figures S15–S17: 13C NMR spectra of complexes (4), (5) and (6a). Figure S18: A packing diagram of compound (6), down to the b axis of the unit cell. Hydrogen atoms of the allyl group have been omitted for clarity. Figure S19: A packing diagram of compound (6a), down to the a axis of the unit cell.

Author Contributions

A.G. (Antonia Garypidou), E.S. and D.T.: methodology, validation, investigation and editing. M.M.: investigation. K.Y. and T.T.: conceptualization, writing—original draft preparation and supervision. A.G. (Achilleas Garoufis) and J.C.P.: conceptualization, supervision, writing—review, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

There was no funding for this project.

Data Availability Statement

Data are contained within the article and supplementary materials.

Acknowledgments

We acknowledge the Unit of Environmental, Organic and Biochemical high-resolution analysis-ORBITRAP-LC-MS, the X-ray Center of single crystal diffraction and the NMR Centre of the University of Ioannina for providing access to the facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zou, S.J.; Shen, Y.; Xie, F.M.; De Chen, J.; Li, Y.Q.; Tang, J.X. Recent advances in organic light-emitting diodes: Toward smart lighting and displays. Mater. Chem. Front. 2020, 4, 788–820. [Google Scholar] [CrossRef]
  2. Salehi, A.; Fu, X.; Shin, D.; So, F. Recent Advances in OLED Optical Design. Adv. Funct. Mater. 2019, 29, 1808803. [Google Scholar] [CrossRef]
  3. Hong, G.; Gan, X.; Leonhardt, C.; Zhang, Z.; Seibert, J.; Busch, J.M.; Bräse, S. A Brief History of OLEDs—Emitter Development and Industry Milestones. Adv. Mater. 2021, 33, 2005630. [Google Scholar] [CrossRef]
  4. Bauri, J.; Choudhary, R.B.; Mandal, G. Recent advances in efficient emissive materials-based OLED applications: A review. J. Mater. Sci. 2021, 56, 18837–18866. [Google Scholar] [CrossRef]
  5. Yao, B.; Giel, M.C.; Hong, Y. Detection of kidney disease biomarkers based on fluorescence technology. Mater. Chem. Front. 2021, 5, 2124–2142. [Google Scholar] [CrossRef]
  6. Zhang, X.; Yao, B.; Hu, Q.; Hong, Y.; Wallace, A.; Reynolds, K.; Ramsey, C.; Maeder, A.; Reed, R.; Tang, Y. Detection of biomarkers in body fluids using bioprobes based on aggregation-induced emission fluorogens. Mater. Chem. Front. 2020, 4, 2548–2570. [Google Scholar] [CrossRef]
  7. Gupta, G.; You, Y.; Hadiputra, R.; Jung, J.; Kang, D.K.; Lee, C.Y. Heterometallic bodipy-based molecular squares obtained by self-assembly: Synthesis and biological activities. ACS Omega 2019, 4, 13200–13208. [Google Scholar] [CrossRef]
  8. Qi, Y.L.; Wang, H.R.; Chen, L.L.; Duan, Y.T.; Yang, S.Y.; Zhu, H.L. Recent advances in small-molecule fluorescent probes for studying ferroptosis. Chem. Soc. Rev. 2022, 51, 7752–7778. [Google Scholar] [CrossRef]
  9. Pham, T.C.; Nguyen, V.-N.; Choi, Y.; Lee, S.; Yoon, J. Recent Strategies to Develop Innovative Photosensitizers for Enhanced Photodynamic Therapy. Chem. Rev. 2021, 121, 13454–13619. [Google Scholar] [CrossRef]
  10. Pal, T.K. Metal-organic framework (MOF)-based fluorescence ‘turn-on’ sensors. Mater. Chem. Front. 2022, 7, 405–441. [Google Scholar] [CrossRef]
  11. Tsutsui, T.; Kusaba, S.; Yamashina, M.; Akita, M.; Yoshizawa, M. Open versus Closed Polyaromatic Nanocavity: Enhanced Host Abilities toward Large Dyes and Pigments. Chem.—A Eur. J. 2019, 25, 4320–4324. [Google Scholar] [CrossRef]
  12. Wu, W.; Li, Z. Nanoprobes with aggregation-induced emission for theranostics. Mater. Chem. Front. 2021, 5, 603–626. [Google Scholar] [CrossRef]
  13. Lee, L.C.C.; Lo, K.K.W. Luminescent and Photofunctional Transition Metal Complexes: From Molecular Design to Diagnostic and Therapeutic Applications. J. Am. Chem. Soc. 2022, 144, 14420–14440. [Google Scholar] [CrossRef] [PubMed]
  14. Dalmau, D.; Urriolabeitia, E.P. Luminescence and Palladium: The Odd Couple. Molecules 2023, 28, 2663. [Google Scholar] [CrossRef] [PubMed]
  15. 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]
  16. Wegeberg, C.; Wenger, O.S. Luminescent First-Row Transition Metal Complexes. JACS Au 2021, 1, 1860–1876. [Google Scholar] [CrossRef]
  17. Cebrián, C.; Mauro, M. Recent advances in phosphorescent platinum complexes for organic light-emitting diodes. Beilstein J. Org. Chem. 2018, 14, 1459–1481. [Google Scholar] [CrossRef] [PubMed]
  18. Yang, X.; Yao, C.; Zhou, G. Highly Efficient Phosphorescent Materials Based on Platinum Complexes and Their Devices (OLEDs). Platin. Met. Rev. 2013, 57, 2–16. [Google Scholar] [CrossRef]
  19. Herberger, J.; Winter, R.F. Platinum emitters with dye-based σ-aryl ligands. Coord. Chem. Rev. 2019, 400, 213048. [Google Scholar] [CrossRef]
  20. Puttock, E.V.; Walden, M.T.; Williams, J.A.G. The luminescence properties of multinuclear platinum complexes. Coord. Chem. Rev. 2018, 367, 127–162. [Google Scholar] [CrossRef]
  21. Sifnaiou, E.; Garypidou, A.; Ypsilantis, K.; Plakatouras, J.C.; Garoufis, A. Synthesis, characterization and photophysical properties of mixed ligand cyclometalated platinum(II) complexes containing 2-phenylpyridine and pyridine carboxylic acids. Polyhedron 2023, 231, 116252. [Google Scholar] [CrossRef]
  22. Nazeeruddin, M.K.; Grätzel, M. Transition Metal Complexes for Photovoltaic and Light Emitting Applications. In Photofunctional Transition Metal Complexes; Springer: Berlin/Heidelberg, Germany, 2007; pp. 113–175. [Google Scholar]
  23. To, W.P.; Wan, Q.; Tong, G.S.M.; Che, C.M. Recent Advances in Metal Triplet Emitters with d6, d8, and d10 Electronic Configurations. Trends Chem. 2020, 2, 796–812. [Google Scholar] [CrossRef]
  24. Yang, X.; Zhou, G.; Wong, W.Y. Functionalization of phosphorescent emitters and their host materials by main-group elements for phosphorescent organic light-emitting devices. Chem. Soc. Rev. 2015, 44, 8484–8575. [Google Scholar] [CrossRef] [PubMed]
  25. Evans, R.C.; Douglas, P.; Winscom, C.J. Coordination complexes exhibiting room-temperature phosphorescence: Evaluation of their suitability as triplet emitters in organic light emitting diodes. Coord. Chem. Rev. 2006, 250, 2093–2126. [Google Scholar] [CrossRef]
  26. Kuwabara, J.; Ogawa, Y.; Taketoshi, A.; Kanbara, T. Enhancement of the photoluminescence of a thioamide-based pincer palladium complex in the crystalline state. J. Organomet. Chem. 2011, 696, 1289–1293. [Google Scholar] [CrossRef]
  27. Laga, E.; Dalmau, D.; Arregui, S.; Crespo, O.; Jimenez, A.I.; Pop, A.; Silvestru, C.; Urriolabeitia, E.P. Fluorescent Orthopalladated Complexes of 4-Aryliden-5(4H)-oxazolones from the Kaede Protein: Synthesis and Characterization. Molecules 2021, 26, 1238. [Google Scholar] [CrossRef]
  28. Expósito, J.E.; Aullón, G.; Bardají, M.; Miguel, J.A.; Espinet, P. Fluorescent perylenylpyridine complexes: An experimental and theoretical study. Dalt. Trans. 2020, 49, 13326–13338. [Google Scholar] [CrossRef]
  29. Li, G.; Zheng, J.; Zhao, X.; Fleetham, T.; Yang, Y.F.; Wang, Q.; Zhan, F.; Zhang, W.; Fang, K.; Zhang, Q.; et al. Tuning the Excited State of Tetradentate Pd(II) Complexes for Highly Efficient Deep-Blue Phosphorescent Materials. Inorg. Chem. 2020, 59, 13502–13516. [Google Scholar] [CrossRef]
  30. Wan, Q.; To, W.P.; Chang, X.; Che, C.M. Controlled Synthesis of PdII and PtII Supramolecular Copolymer with Sequential Multiblock and Amplified Phosphorescence. Chem 2020, 6, 945–967. [Google Scholar] [CrossRef]
  31. Accorsi, G.; Listorti, A.; Yoosaf, K.; Armaroli, N. 1,10-Phenanthrolines: Versatile building blocks for luminescent molecules, materials and metal complexes. Chem. Soc. Rev. 2009, 38, 1690. [Google Scholar] [CrossRef]
  32. de França, B.M.; Oliveira, S.S.C.; Souza, L.O.P.; Mello, T.P.; Santos, A.L.S.; Forero, J.S.B. Synthesis and photophysical properties of metal complexes of curcumin dyes: Solvatochromism, acidochromism, and photoactivity. Dye. Pigment. 2022, 198, 110011. [Google Scholar] [CrossRef]
  33. Büyükekşi, S.I.; Şengül, A.; Erdönmez, S.; Altindal, A.; Orman, E.B.; Özkaya, A.R. Spectroscopic, electrochemical and photovoltaic properties of Pt(II) and Pd(II) complexes of a chelating 1,10-phenanthroline appended perylene dIImide. Dalt. Trans. 2018, 47, 2549–2560. [Google Scholar] [CrossRef] [PubMed]
  34. Ramdeehul, S.; Barloy, L.; Osborn, J.A.; De Cian, A.; Fischer, J. Synthesis, Solution Dynamics, and Crystal Structures of (2,2′:6′,2″-Terpyridine)(η1- and η3- allyl)palladium(II) Complexes. Organometallics 1996, 15, 5442–5444. [Google Scholar] [CrossRef]
  35. Hansson, S.; Norrby, P.O.; Soegren, M.P.; Aakermark, B.; Cucciolito, M.E.; Giordano, F.; Vitagliano, A. Effects of Phenanthroline Type Ligands on the Dynamic Processes of (η3-Allyl) palladium Complexes. Molecular Structure of (2,9-Dimethyl-1,10-phenanthroline) [(l, 2, 3-η)-3-methyl-2-butenyl]-chloropalladium. Organometallics 1993, 12, 4940–4948. [Google Scholar] [CrossRef]
  36. Sjoegren, M.; Hansson, S.; Norrby, P.O.; Aakermark, B.; Cucciolito, M.E.; Vitagliano, A. Selective stabilization of the anti isomer of (. eta. 3-allyl)palladium and -platinum complexes. Organometallics 1992, 11, 3954–3964. [Google Scholar] [CrossRef]
  37. Vitagliano, A.; Aakermark, B.; Hansson, S. Convenient synthesis of cationic (. eta. 3-allyl)palladium complexes. Preparative and stereochemical aspects. Organometallics 1991, 10, 2592–2599. [Google Scholar] [CrossRef]
  38. Gogoll, A.; Oernebro, J.; Grennberg, H.; Baeckvall, J.-E. Mechanism of Apparent. pi.-Allyl Rotation in (. pi.-Allyl)palladium Complexes with Bidentate Nitrogen Ligands. J. Am. Chem. Soc. 1994, 116, 3631–3632. [Google Scholar] [CrossRef]
  39. Parisotto, S.; Deagostino, A. π-Allylpalladium Complexes in Synthesis: An Update. Synthesis 2019, 51, 1892–1912. [Google Scholar] [CrossRef]
  40. Liu, Y.; Oble, J.; Pradal, A.; Poli, G. Catalytic Domino Annulations through η3-Allylpalladium Chemistry: A Never-Ending Story. Eur. J. Inorg. Chem. 2020, 2020, 942–961. [Google Scholar] [CrossRef]
  41. Du, J.; Li, Y.-F.; Ding, C.-H. Recent advances of Pd-p-allyl zwitterions in cycloaddition reactions. Chinese Chem. Lett. 2023, 34, 108401. [Google Scholar] [CrossRef]
  42. Scattolin, T.; Pessotto, I.; Cavarzerani, E.; Canzonieri, V.; Orian, L.; Demitri, N.; Schmidt, C.; Casini, A.; Bortolamiol, E.; Visentin, F.; et al. Indenyl and Allyl Palladate Complexes Bearing N-Heterocyclic Carbene Ligands: An Easily Accessible Class of New Anticancer Drug Candidates. Eur. J. Inorg. Chem. 2022, 2022, e202200103. [Google Scholar] [CrossRef]
  43. Tupini, C.; Zurlo, M.; Gasparello, J.; Lodi, I.; Finotti, A.; Scattolin, T.; Visentin, F.; Gambari, R.; Lampronti, I. Combined Treatment of Cancer Cells Using Allyl Palladium Complexes Bearing Purine-Based NHC Ligands and Molecules Targeting MicroRNAs miR-221-3p and miR-222-3p: Synergistic Effects on Apoptosis. Pharmaceutics 2023, 15, 1332. [Google Scholar] [CrossRef]
  44. Scattolin, T.; Bortolamiol, E.; Visentin, F.; Palazzolo, S.; Caligiuri, I.; Perin, T.; Canzonieri, V.; Demitri, N.; Rizzolio, F.; Togni, A. Palladium(II)-η3-Allyl Complexes Bearing N-Trifluoromethyl N-Heterocyclic Carbenes: A New Generation of Anticancer Agents that Restrain the Growth of High-Grade Serous Ovarian Cancer Tumoroids. Chem. A Eur. J. 2020, 26, 11868–11876. [Google Scholar] [CrossRef] [PubMed]
  45. Scattolin, T.; Bortolamiol, E.; Rizzolio, F.; Demitri, N.; Visentin, F. Allyl palladium complexes bearing carbohydrate-based N-heterocyclic carbenes: Anticancer agents for selective and potent in vitro cytotoxicity. Appl. Organomet. Chem. 2020, 34, e5876. [Google Scholar] [CrossRef]
  46. Mayoral, M.J.; Ovejero, P.; Campo, J.A.; Heras, J.V.; Oliveira, E.; Pedras, B.; Lodeiro, C.; Cano, M. Exploring photophysical properties of new boron and palladium (II) complexes with β-diketone pyridine type ligands: From liquid crystals to metal fluorescence probes. J. Mater. Chem. 2011, 21, 1255–1263. [Google Scholar] [CrossRef]
  47. Scattolin, T.; Andreetta, G.; Mauceri, M.; Rizzolio, F.; Demitri, N.; Canzonieri, V.; Visentin, F. Imidazo[1,5-a]pyridine-3-ylidenes and dipyridoimidazolinylidenes as ancillary ligands in Palladium allyl complexes with potent in vitro anticancer activity. J. Organomet. Chem. 2021, 952, 122014. [Google Scholar] [CrossRef]
  48. Kasselouri, S.; Garoufis, A.; Katehanakis, A.; Kalkanis, G.; Perlepes, S.P.; Hadjiliadis, N. 1:1 Metal complexes of 2-(2′-pyridyl) quinoxaline, a ligand unexpectedly formed by the reaction between 2-acetylpyridine and 1,2-phenylenediamine. Inorganica Chim. Acta 1993, 207, 255–258. [Google Scholar] [CrossRef]
  49. Tatsuno, Y.; Yoshida, T.; Otsuka, S.E.I.; Syntheses, I. Inorganic Syntheses; Wiley: Hoboken, NJ, USA, 1990. [Google Scholar]
  50. APEX 3; SAINT, SHELXT; Bruker AXS Inc.: Fitchburg, WI, USA, 2016.
  51. Sheldrick, G.M. SADABS; University of Göttingen: Göttingen, Germany, 1996. [Google Scholar]
  52. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  53. Hübschle, C.B.; Sheldrick, G.M.; Dittrich, B. ShelXle: A Qt graphical user interface for SHELXL. J. Appl. Crystallogr. 2011, 44, 1281–1284. [Google Scholar] [CrossRef]
  54. Spek, A.L. Structure validation in chemical crystallography. Acta Crystallogr. Sect. D Biol. Crystallogr. 2009, 65, 148–155. [Google Scholar] [CrossRef]
  55. Barbour, L.J. X-Seed—A Software Tool for Supramolecular Crystallography. J. Supramol. Chem. 2001, 1, 189–191. [Google Scholar] [CrossRef]
  56. Yasuda, M.; Sone, K.; Yamasaki, K. Stability of Zinc and Cadmium Complexes with Some Methyl Derivatives of 1,10-Phenanthroline and 2,2′-Bipyridine. J. Phys. Chem. 1956, 60, 1667–1668. [Google Scholar] [CrossRef]
  57. Yamasaki, K.; Yasuda, M. Stability of Zinc and Cadmium Complexes with 2,2′-Bipyridine and 1,10-Phenanthroline. J. Am. Chem. Soc. 1956, 78, 1324. [Google Scholar] [CrossRef]
  58. Babenko, M.; Busev, A.I.; Blokh, M.S. Extraction-spectrometric study of complexes formed during reaction of vanadium (ii), 1, 10-phenanthroline and bengal rose-a. Zhurnal Neorg. KhimII 1973, 18, 1326–1330. [Google Scholar]
  59. Torralba, M.C.; Cano, M.; Campo, J.A.; Heras, J.V.; Pinilla, E.; Torres, M.R. Pyrazole-based allylpalladium complexes: Supramolecular architecture and liquid crystal behaviour. Inorg. Chem. Commun. 2006, 9, 1271–1275. [Google Scholar] [CrossRef]
  60. Torralba, M.C.; Cano, M.; Campo, J.A.; Heras, J.V.; Pinilla, E.; Torres, M.R. Liquid crystal behaviour of ionic allylpalladium complexes containing 2-pyrazolylpyridine as bidentate N,N′-ligand. J. Organomet. Chem. 2006, 691, 765–778. [Google Scholar] [CrossRef]
  61. Amadio, E.; Scrivanti, A.; Chessa, G.; Matteoli, U.; Beghetto, V.; Bertoldini, M.; Rancan, M.; Dolmella, A.; Venzo, A.; Bertani, R. Synthesis, characterization and low temperature self assembling of (η3-allyl)palladium complexes with 2-pyridyl-1,2,3-triazole bidentate ligands. Study of the catalytic activity in Suzuki–Miyaura reaction. J. Organomet. Chem. 2012, 716, 193–200. [Google Scholar] [CrossRef]
  62. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The Cambridge Structural Database. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 2016, 72, 171–179. [Google Scholar] [CrossRef]
  63. Böttcher, L.; Scholz, A.; Walther, D.; Weisbach, N.; Görls, H. Mononucleare Oxalamidinkomplexe von Kupfer(I), Palladium(II) und Ruthenium(II) durch gekoppelte Kation-/Anionkoordination. Zeitschrift Anorg. Allg. Chemie 2003, 629, 2103–2112. [Google Scholar] [CrossRef]
  64. Wild, U.; Kaifer, E.; Himmel, H. Redox Chemistry and Group 10 Metal Complexes of Aromatic Compounds with Bulky Bicyclic Guanidino Groups. Eur. J. Inorg. Chem. 2011, 2011, 4220–4233. [Google Scholar] [CrossRef]
  65. Löffler, J.; Gauld, R.M.; Feichtner, K.S.; Rodstein, I.; Zur, J.A.; Handelmann, J.; Schwarz, C.; Gessner, V.H. Ylide-Substituted Phosphines with a Cyclic Ylide-Backbone: Angle Dependence of the Donor Strength. Organometallics 2021, 40, 2888–2900. [Google Scholar] [CrossRef] [PubMed]
  66. Chan, S.C.; Chan, M.C.W.; Wang, Y.; Che, C.M.; Cheung, K.K.; Zhu, N. Organic light-emitting materials based on bis(arylacetylide)platinum(II) complexes bearing substituted bipyridine and phenanthroline ligands: Photo- and electroluminescence from 3MLCT excited states. Chem A Eur. J. 2001, 7, 4180–4190. [Google Scholar] [CrossRef]
Scheme 1. Structures and numbering of ligands used in this study. mphen = 5-methyl-1,10-phenanthroline, tmphen = 3,4,7,8-tetramethyl-1,10-phenanthroline, ncp = 2,9-dimethyl-1,10-phenanthroline, bphen = 4,7-diphenyl-1,10-phenanthroline, bcp = 2,9-dimethyl-4,7-diphenyl-1,10-phenanthroline and pqx = 2-(2′-pyridyl)quinoxaline.
Scheme 1. Structures and numbering of ligands used in this study. mphen = 5-methyl-1,10-phenanthroline, tmphen = 3,4,7,8-tetramethyl-1,10-phenanthroline, ncp = 2,9-dimethyl-1,10-phenanthroline, bphen = 4,7-diphenyl-1,10-phenanthroline, bcp = 2,9-dimethyl-4,7-diphenyl-1,10-phenanthroline and pqx = 2-(2′-pyridyl)quinoxaline.
Chemistry 05 00162 sch001
Scheme 2. Synthetic procedure for complexes (1)–(6).
Scheme 2. Synthetic procedure for complexes (1)–(6).
Chemistry 05 00162 sch002
Figure 1. 1H NMR spectrum of complex (2) in acetone-d6. Inset shows the expansion of the spectrum aliphatic area and the numbered structure of the η3-allyl anionic ligand. Peaks with asterisk denote the solvent and the HOD.
Figure 1. 1H NMR spectrum of complex (2) in acetone-d6. Inset shows the expansion of the spectrum aliphatic area and the numbered structure of the η3-allyl anionic ligand. Peaks with asterisk denote the solvent and the HOD.
Chemistry 05 00162 g001
Figure 2. The NOESY spectrum of (2) at mixing time 600 ms, in acetone-d6 at 298 K showing the NOE connectivities between the η3-allyl protons and between the methyl groups of ncp and the allyl syn protons, Hc1″/Ha1″.
Figure 2. The NOESY spectrum of (2) at mixing time 600 ms, in acetone-d6 at 298 K showing the NOE connectivities between the η3-allyl protons and between the methyl groups of ncp and the allyl syn protons, Hc1″/Ha1″.
Chemistry 05 00162 g002
Figure 3. A partially labeled thermal ellipsoid plot (50% probability level) of the cation in the crystal structure of (6). Only one part of the disordered allyl group is shown for clarity.
Figure 3. A partially labeled thermal ellipsoid plot (50% probability level) of the cation in the crystal structure of (6). Only one part of the disordered allyl group is shown for clarity.
Chemistry 05 00162 g003
Figure 4. A partially labeled thermal ellipsoid plot (50% probability level) of the ionic pair in the crystal structure of (6a). Only the parts with the highest occupancy of the disordered allyl groups are shown for clarity.
Figure 4. A partially labeled thermal ellipsoid plot (50% probability level) of the ionic pair in the crystal structure of (6a). Only the parts with the highest occupancy of the disordered allyl groups are shown for clarity.
Chemistry 05 00162 g004
Figure 5. Solid state absorption (a) and emission (b) spectra of complexes (1)–(6).
Figure 5. Solid state absorption (a) and emission (b) spectra of complexes (1)–(6).
Chemistry 05 00162 g005
Figure 6. UV-Vis (a) and emission (b) spectra of complexes (1)–(6) in CH2Cl2.
Figure 6. UV-Vis (a) and emission (b) spectra of complexes (1)–(6) in CH2Cl2.
Chemistry 05 00162 g006
Table 1. Crystal data and structural refinement for compounds [Pd(η3-C3H5)(pqx)]PF6 (6) and [Pd(η3-C3H5)(pqx)][Pd(η3-C3H5)Cl2] (6a).
Table 1. Crystal data and structural refinement for compounds [Pd(η3-C3H5)(pqx)]PF6 (6) and [Pd(η3-C3H5)(pqx)][Pd(η3-C3H5)Cl2] (6a).
Compound(6)(6a)
Empirical formulaC16H14F6N3PPdC19H19Cl2N3Pd2
Formula weight499.67573.12
Temperature (Κ)296(2)
Wavelength (Å)0.71073
Crystal systemOrthorhombicMonoclinic
Space groupPnmaP21/c
Unit cell dimensions
a, b, c (Å), α, β, γ (ο)
18.7765(10), 6.8527(3), 16.4449(10), 90, 90, 906.9296(3), 29.3104(12), 9.8311(5), 90, 98.680(2), 90
Volume (Å3)2115.96(19) 1973.92(15)
Z44
Density (calcd.) (g/cm3)1.569 1.928
Absorption coefficient (mm−1)1.008 2.100
F(000)9841120
Crystal size (mm3)0.600 × 0.120 × 0.080 0.500 × 0.040 × 0.040
θ range for data collection (ο)2.477 to 24.9962.780 to 24.991
Index ranges−22 ≤ h ≤ 22, −8 ≤ k ≤ 7, −19 ≤ l ≤ 19−8 ≤ h ≤ 7, −34 ≤ k ≤ 34, −11 ≤ l ≤ 11
Reflections collected3456744871
Independent reflections2031 [Rint = 0.0499]3452 [Rint = 0.0801]
Completeness to θ = 24.996°99.8%99.2%
Refinement methodFull-matrix least-squares on F2
Data/restraints/parameters2031/0/1583452/180/291
Goodness-of-fit1.1851.239
Final R indices [I > 2σ(I)]Robs = 0.0405, wRobs = 0.1089Robs = 0.0834, wRobs = 0.1664
R indices [all data]Rall = 0.0439, wRall = 0.1110Rall = 0.0947, wRall = 0.1699
Largest diff. peak and hole (e·Å−3)0.452 and −1.135 1.703 and −1.761
R = Σ||Fo| − |Fc||/Σ|Fo|, wR = {Σ[w(|Fo|2 − |Fc|2)2]/Σ[w(|Fo|4)]}1/2 and w = 1/[σ2(Fo2) + (aP)2 + bP] where p = (Fo2 + 2Fc2)/3 where for 6: a = 0.0527, b = 3.6386; 6a: a = 0, b = 49.5689.
Table 2. Selected bond distances (Å) and angles (o) for [Pd(η3-C3H5)(pqx)]PF6 (6) and [Pd(η3-C3H5)(pqx)][Pd(η3-C3H5)Cl2] (6a).
Table 2. Selected bond distances (Å) and angles (o) for [Pd(η3-C3H5)(pqx)]PF6 (6) and [Pd(η3-C3H5)(pqx)][Pd(η3-C3H5)Cl2] (6a).
Compound (6)
Pd(1)–N(1)2.144(4)Pd(1)–N(3)2.090(5)
Pd(1)–C(1A)2.150(6)Pd(1)–C(2A)2.137(7)
Pd(1)–C(3A)2.115(5)
N(1)–Pd(1)–C(1A)111.71(18)N(1)–Pd(1)–N(3)78.19(19)
N(1)–Pd(1)–C(2A)142.7(2)N(1)–Pd(1)–C(3A)179.63(19)
N(3)–Pd(1)–C(1A)170.11(19)N(3)–Pd(1)–C(3A)102.2(2)
N(3)–Pd(1)–C(2A)132.66(18)C(1A)–Pd(1)–C(3A)67.93(13)
Compound (6a)
Pd(1)–N(1)2.170(9)Pd(1)–N(3)2.082(11)
Pd(1)–C(1A)2.13(12)Pd(1)–C(2A)2.18(8)
Pd(1)–C(3A)2.17(12)Pd(2)–C(1B)2.07(6)
Pd(2)–Cl(1)2.369(4)Pd(2)–C(2B)2.11(3)
Pd(2)–Cl(2)2.381(4)Pd(2)–C(3B)2.07(5)
N(1)–Pd(1)–C(1A)167(3)N(1)–Pd(1)–N(3)78.0(4)
N(1)–Pd(1)–C(2A)141(2)N(1)–Pd(1)–C(3A)117(3)
N(3)–Pd(1)–C(1A)93(3)N(3)–Pd(1)–C(3A)163(3)
N(3)–Pd(1)–C(2A)134(2)C(1A)–Pd(1)–C(3A)74(4)
Cl(1)–Pd(2)–Cl(2)95.82(16)C(1B)–Pd(2)–Cl(1)96.5(15)
C(2B)–Pd(2)–Cl(1)127.1(9)C(3B)–Pd(2)–Cl(1)169.1(10)
C(1B)–Pd(2)–Cl(2)166.6(16)C(2B)–Pd(2)–Cl(2)133.8(9)
C(3B)–Pd(2)–Cl(2)95.1(10)
Table 3. Photophysical data of complexes (1)–(6) in solid state and CH2Cl2-diluted solutions. Excitation in solid state, λexc = 400 nm and in CH2Cl2, λexc = 365 nm.
Table 3. Photophysical data of complexes (1)–(6) in solid state and CH2Cl2-diluted solutions. Excitation in solid state, λexc = 400 nm and in CH2Cl2, λexc = 365 nm.
ComplexUV/Vis Absorbance
λmax (nm), (ε × 104 M−1cm−1))
Emission λem (nm)QY (%)
SolidSolutionSolidSolutionSolidSolution
(1) bp237, 309, 420232(18), 292(20), 321sh, 360(2)478, 5654333%1.6%
(2) ncp238, 311, 394sh234(40), 279(35), 295sh, 355(1)477, 56643612.6%1.2%
(3) bcp239, 311, 370sh233(41), 294(42), 359(2.5)478, 565433, 533sh6%1%
(4) mp232, 310, 358sh233(34), 277(24), 297sh, 362(1)478, 56643714%1.15%
(5) tmp236, 312, 352sh 236(37), 281(37), 304sh, 338(4.5)476, 565456sh, 486, 510sh14%2.8%
(6) pqx305, 394254(20), 364(16)480, 5654371.9%0.15%
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

Garypidou, A.; Ypsilantis, K.; Sifnaiou, E.; Manthou, M.; Thomos, D.; Plakatouras, J.C.; Tsolis, T.; Garoufis, A. Synthesis, Characterization and Photophysical Properties of Mixed Ligand (η3-Allyl)palladium(II) Complexes with N,N’Aromatic Diimines. Chemistry 2023, 5, 2476-2489. https://doi.org/10.3390/chemistry5040162

AMA Style

Garypidou A, Ypsilantis K, Sifnaiou E, Manthou M, Thomos D, Plakatouras JC, Tsolis T, Garoufis A. Synthesis, Characterization and Photophysical Properties of Mixed Ligand (η3-Allyl)palladium(II) Complexes with N,N’Aromatic Diimines. Chemistry. 2023; 5(4):2476-2489. https://doi.org/10.3390/chemistry5040162

Chicago/Turabian Style

Garypidou, Antonia, Konstantinos Ypsilantis, Evaggelia Sifnaiou, Maria Manthou, Dimitris Thomos, John C. Plakatouras, Theodoros Tsolis, and Achilleas Garoufis. 2023. "Synthesis, Characterization and Photophysical Properties of Mixed Ligand (η3-Allyl)palladium(II) Complexes with N,N’Aromatic Diimines" Chemistry 5, no. 4: 2476-2489. https://doi.org/10.3390/chemistry5040162

Article Metrics

Back to TopTop