DOI:
10.1039/D4QO01697A
(Research Article)
Org. Chem. Front., 2024,
11, 6651-6659
Ion-pairing assemblies of anion-responsive helical PtII complexes†
Received
10th September 2024
, Accepted 1st October 2024
First published on 2nd October 2024
Abstract
Naphthylisoquinoline-appended dipyrrolyldiketone PtII complexes as helical π-electronic systems were synthesized. The PtII complexes, showing chiroptical properties in solution, exhibited anion-binding behaviour, resulting in the formation of anion complexes as pseudo π-electronic anions. In combination with the π-electronic cation, the receptor–anion complexes formed charge-by-charge ion-pairing assemblies, with the narcissistic self-sorting columnar structures comprising either of the enantiomers, in the crystal state.
Introduction
Assemblies of π-electronic systems show fascinating electronic and electrooptical properties depending on the arrangements of constituting species. In particular, modulating the assembly modes of π-electronic systems with photoluminescence properties affords various electrooptical materials, such as light-emitting diodes, lasers and sensors.1 Among various luminescent π-electronic molecules reported thus far, organoplatinum(II) complexes with square-planar geometries exhibit photoluminescence properties,2,3 such as triplet energy transfer and phosphorescence anisotropy amplification.3h,j Unique assembling behaviours, including stacking structures and PtII⋯PtII interactions, result in characteristic electronic and photoluminescence properties.
The electronic and photoluminescence properties of PtII complexes in solution and assembly states can be controlled by introduced π-electronic ligand molecules.3 As phosphorescent π-electronic PtII complexes, dipyrrolyldiketone PtII complexes (e.g., 1, Fig. 1a)4 showing anion-responsive abilities form anion complexes as pseudo π-electronic anions by binding guest anions and resulting ion pairs in combination with countercations. The π-electronic anions form ion-pairing assemblies through iπ–iπ interactions with π-electronic cations to form charge-by-charge assemblies.5,6 The phosphorescence properties of the assemblies depend on countercations that modulate the arrangement of components. For example, the enhancement of the solid-state phosphorescence was observed by the formation of charge-by-charge assemblies with bulky countercations.4b Thus, an appropriate combination with countercations and modification of π-electronic ligands would result in modulated electronic properties.
|
| Fig. 1 (a) Anion complexation of dipyrrolyldiketone PtII complex 1 (X−: guest anion) and (b) the dipyrrolyldiketone PtII complex with a helical π-electronic unit (left) and the conceptual diagram of charge-by-charge ion-pairing assemblies (right). | |
Modification of the π-electronic ligands affects the chiroptical properties of PtII complexes. For example, chiral pendants attached to PtII complexes induce chiroptical properties.7 Furthermore, PtII complexes with helical π-electronic ligands, as seen in those with helicene-like structures, exhibit chiroptical properties.8 Such chiral PtII complexes with phosphorescence properties have been reported thus far for circularly polarized luminescence (CPL) properties. The chiroptical properties of PtII complexes can be modulated by binding guest species. In particular, ion-pairing assemblies of chiral PtII complexes would provide phosphorescent CPL materials. However, there are no reports on helical PtII complexes with anion-responsive properties (Fig. 1b). In light of this, it would be important to synthesize anion-responsive chiral PtII complexes that form ion-pairing assemblies in order to control the chiroptical characteristics of phosphorescence. This report describes the synthesis and ion-pairing assembly of helical PtII complexes to provide insights into future chiroptical ion-pairing materials.
Results and discussion
Synthesis and characterization of PtII complexes
As a preliminary examination for the reactivity in the PtII complexation of dipyrrolyldiketones9 and naphthylisoquinoline ligands,10 PtII complex 2a was synthesized in 32% yield by treating naphthylisoquinoline10a with [(PtMe2)2(SMe2)2]11 at r.t. in the presence of trifluoromethanesulfonic acid (TfOH) (1 equiv.), followed by the addition of dipyrrolyldiketone10a and K2CO3 (1.5 equiv.) (Fig. 2). As a helical π-electronic ligand for PtII complexation, 8-phenyl-substituted naphthylisoquinoline10b was selected for inducing chiral states in anion-complexing and ion-pairing forms. Under the same reaction conditions, PtII complexes 3a,b were prepared in 35% and 19% yields, respectively, as yellow solids from the corresponding dipyrrolyldiketones. The obtained PtII complexes were characterized by 1H and 13C NMR and ESI-TOF-MS. Theoretical investigations for the most stable conformations of 2a and 3a,b at the B3LYP/6-31G(d,p) level with LanL2DZ for Pt suggested that doubly pyrrole-inverted geometries suitable for anion binding were less stable by 6.35, 6.60 and 6.60 kcal mol−1, respectively (Fig. S19–21†),12 as observed in the previously reported dipyrrolyldiketone PtII complexes including 1.4a The naphthylisoquinoline unit in the optimized structures of 3a,b exhibited helical structures with a ligand-phenyl unit partially overlapped with the isoquinoline plane. The helical pitch of 3a (3.14 Å) was longer by 0.15 Å than that of 2a (2.99 Å), suggesting that the sterically hindered ligand-phenyl unit affords helical conformations without helicity inversion.
|
| Fig. 2 Dipyrrolyldiketone PtII complexes with a naphthylisoquinoline ligand 2a and 3a,b. | |
As observed in the optimized structures of 3a,b, the X-ray analysis of 3a for the single crystal obtained from CH2Cl2/n-hexane revealed a helical geometry and packing structure (Fig. 3 and Fig. S12†). The diketone-naphthylisoquinoline-coordinated PtII unit showed angles of 80.4°–94.7° around PtII with a τ4 value13 of 0.09, indicating an almost square-planar geometry. The ligand-phenyl unit was partially overlapped with the isoquinoline unit to induce the helical geometry with a helical pitch of 3.12 Å, being shorter by 0.02 Å than that of the optimized structure presumably due to the stacking effect. In the crystal structure, P- and M-helical enantiomers formed a stacked dimer with a stacking distance of 3.29 Å based on the mean plane of O, O, N, C and Pt and a Pt⋯Pt distance of 4.16 Å.
|
| Fig. 3 Single-crystal X-ray structure of 3a: (a) packing structure and (b) top and side views as the P-helix. The colours brown, pink, blue, red and grey in this figure and Fig. 6 represent carbon, hydrogen, nitrogen, oxygen and platinum, respectively. | |
Solution-state properties and anion-binding behaviours
PtII complexes 2a and 3a exhibited UV/vis absorption spectra with λmax at 380 and 386 nm and shoulders at 359/407 and 352/408 nm, respectively, in CH2Cl2, whereas phenyl-substituted 3b showed red-shifted absorption with λmax at 452 nm and shoulders at 382 and 430 nm (Fig. S4†). The main absorption band of 3a at 386 nm was assigned to the HOMO−1-to-LUMO+1 transition, whereas that of 3b at 452 nm was attributed to the HOMO-to-LUMO+1 transition (Fig. S23, 24, 26 and 27†). Interestingly, in contrast to the phosphorescence emissive properties of dipyrrolyldiketone PtII complexes, such as 1, possessing an arylpyridine ligand,42a and 3a,b showed luminescence with low quantum efficiencies at 509/676, 571/711 and 582/713 nm when excited at 380, 387 and 452 nm, respectively, in deoxygenated CH2Cl2 (Fig. S5†). The luminescence intensities at longer wavelengths for the PtII complexes decreased under atmospheric conditions, suggesting that the luminescence bands at the shorter and longer wavelengths were derived from fluorescence and phosphorescence, respectively.8g Theoretical calculations of 2a and 3a at CAM-B3LYP/6-31+G(d,p) with LanL2DZ for Pt revealed that the electron density differences of T1–S0 were mainly observed at the naphthylisoquinoline unit, whereas those at the dipyrrolyldiketone unit were seen for 1. The S0–T1 excitation energies of 2a and 3a are 1.64 and 1.54 eV, respectively, which are lower than that of 1 (2.41 eV). The low phosphorescence intensities of the PtII complexes with a naphthylisoquinoline unit resulted from lower T1 excited-state energy levels, inducing T1–S0 nonradiative transitions.
PtII complexes 3a,b, as racemic mixtures, were resolved via chiral HPLC separation to obtain the enantiopure P and M enantiomers (Fig. S6†). Circular dichroism (CD) spectra of the first fraction of 3a in CH2Cl2 (3 × 10−5 M) showed negative-to-positive signs in the 250–300 nm region (Fig. 4), whereas 3b exhibited the opposite CD signs for the first fraction (Fig. S7†). These CD spectra can be assigned as characteristic binaphthyl-type patterns.14 Theoretically estimated CD spectra of 3a,b at the B3LYP/6-31+G(d,p) level for the optimized structures suggest that the first fractions included the P- and M-helical enantiomers, respectively (Fig. S29†).
|
| Fig. 4 CD spectra of 3a in CH2Cl2 (3 × 10−5 M) (inset: optimized structures of 3a as P- and M-helices at B3LYP/6-31G(d,p)). The solid and dotted lines correspond to the spectra of the first and second fractions, respectively, in chiral HPLC separation. | |
Dipyrrolyldiketone PtII complexes exhibit anion-binding behaviour through multiple hydrogen bonding4 as seen in boron complexes. The solution-state anion-binding behaviour was evaluated using 1H NMR and UV/vis absorption spectral changes. Upon the addition of Cl− as a tetrabutylammonium (TBA+) salt to 3a in CD2Cl2 (1.0 × 10−3 M) at −50 °C, the signals at 9.51 and 6.51 ppm ascribable to those of pyrrole NH and bridging CH, respectively, were shifted downfield to 12.56/12.51 and 7.98 ppm, respectively, suggesting the formation of a [1 + 1]-type Cl− complex through hydrogen bonding (Fig. 5). Similarly, the [1 + 1]-type anion complexation of 2a and 3b was confirmed with 1H NMR spectral changes upon Cl− addition (Fig. S38 and 40†). In particular, the o-CH of the phenyl rings attached to the pyrrole units in 3b also supported weak hydrogen bonding for anion binding. The [1 + 1]-type Cl− binding constants (Ka) for 2a and 3a were estimated to be 1 × 103 and 2 × 103 M−1, respectively, based on the 1H NMR measurements in CD2Cl2 at −50 °C (Fig. S38 and 39†). Conversely, the Ka values of 3b were evaluated by UV/vis absorption spectral changes upon the addition of anions as TBA salts. The Ka values of 3b at r.t. were estimated to be 2400, 300 and 5600 M−1 for Cl−, Br− and CH3CO2−, respectively, which are similar to those of PtII complexes with arylpyridines.4a In contrast to 3b, α-unsubstituted 2a and 3a exhibited no drastic changes in the UV/vis absorption spectra upon the addition of anions (Fig. S34 and 35†). The conformation changes in the anion-free and anion-binding forms resulted in very small changes in the direction of the transition dipole moments. In addition, the chiral π-electronic ligand on PtII did not significantly affect the pyrrole-inverted structure and the resulting anion-binding properties. Furthermore, slightly decreased CD intensities at 300–400 nm were observed for 3a,b upon the addition of Cl− (Fig. S37†), suggesting that changes in the conformations and electronic states due to anion binding only marginally affected the chirality of the naphthylisoquinoline ligand.
|
| Fig. 5 (a) 1H NMR of 3a (1 mM) upon the addition of Cl− as a TBA salt in CD2Cl2 at −50 °C and (b) Cl−-binding behaviour of 3a shown by optimized structures at B3LYP/6-31G(d,p). | |
Ion-pairing assemblies of chiral PtII complexes
The exact structures of the anion complexes and ion-pairing assemblies were revealed using single-crystal X-ray analysis. The single crystal of 2a·Cl−-TATA+ (TATA+ = 4,8,12-tripropyl-4,8,12-triazatriangulenium cation)15 suitable for X-ray analysis was prepared by vapour diffusion of n-hexane into a CH2Cl2 solution of a 1:1 mixture of 2a and TATACl. In the solid state, 2a·Cl− exhibited a [1 + 1]-type binding mode with the pyrrole N(–H)⋯Cl− and bridging C(–H)⋯Cl− distances of 3.19/3.22 and 3.58 Å, respectively (Fig. 6a(i) and Fig. S13†). 2a·Cl−-TATA+ formed a charge-by-charge assembly with alternately stacked 2a·Cl− and TATA+ units with stacking distances of 3.42 and 3.48 Å (Fig. 6a(ii)). Single crystals of 3a·Cl−-TATA+ and 3b·Cl−-TATA+ were also obtained from vapour diffusion of n-hexane into a CH2Cl2 solution of a 1:1 mixture of 3a,b and TATACl (Fig. 6b,c(i) and Fig. S14, 15†). Similarly, 3a·Cl−-TATA+ and 3b·Cl−-TATA+ formed charge-by-charge assemblies of the Cl− complexes and TATA+ with stacking distances of 3.39 and 3.45/3.47 Å, respectively (Fig. 6b,c(ii)). TATA+ was mainly stacked with the plane of the anion complexes to form columnar structures aligned along the a-, b- and c-axis directions for 2a·Cl−-TATA+, 3a·Cl−-TATA+ and 3b·Cl−-TATA+, respectively. In all the ion pairs based on the PtII complexes including 2a, the charge-by-charge columns comprise either of the enantiomers and the corresponding enantiomers, exhibiting narcissistic self-sorting of the enantiomers via ion pairing with the π-electronic cation. This observation contrasts with that of the anion-free 3a, which shows social self-sorting stacking columns in the crystal structure. Notably, the arrangement of helical structures can be modulated via anion binding and ion pairing. The charge-by-charge columns of the same enantiomer were aligned along the b-, a- and b-axis directions for 2a·Cl−-TATA+, 3a·Cl−-TATA+ and 3b·Cl−-TATA+, respectively, to form layer structures, which were alternately arranged with the layers of opposite chirality along the c-, c- and a-axis directions (Fig. 6a–c(iii)).
|
| Fig. 6 Single-crystal X-ray structures of (a) 2a·Cl−-TATA+, (b) 3a·Cl−-TATA+ and (c) 3b·Cl−-TATA+ as (i) representative stacked ion pairs and (ii) side and (iii) top views of packing structures. Green spheres in (i) and (iii) represent chlorine. The colours magenta and cyan in (ii) represent anions and cations, respectively. The colours orange and green in (iii) represent P- and M-helical enantiomers, respectively. | |
Hirshfeld surface analysis16 for the charge-by-charge structures clearly showed red and blue triangles arranged in bow-tie shapes on the shape-index surface and a flat region on the curvedness, suggesting characteristic mapping patterns for π–π stacking structures of the planar charged species (Fig. S17 and 18†). The interaction energies for the stacking of TATA+ and the anion complexes were estimated by energy decomposition analysis (EDA)17 in the framework of the fragment molecular orbital (FMO) method at FMO2-MP2 using mixed basis sets including NOSeC-V-DZP with TZP for Pt with MCP (Tables S1–3 and Fig. S31, 32†).18–20 The EDA calculations determine electrostatic, dispersion, charge transfer and exchange-repulsion interaction energies along with the total interaction energy. In all the ion-pairing assemblies, the main interaction energies were derived from electrostatic and dispersion forces. These energies and total energies for the representative pairing Cl− complexes and TATA+ in 2a·Cl−-TATA+, 3a·Cl−-TATA+ and 3b·Cl−-TATA+ were calculated to be −89.1/–134.7/–202.4, −77.1/–143.8/–197.6 and −82.2/–155.7/–214.0 kcal mol−1, respectively, suggesting iπ–iπ interactions.6 Narcissistic self-sorting charge-by-charge columns were formed by stacking of the chiral PtII complexes with the achiral π-electronic cation through iπ–iπ interactions.
Conclusions
Dipyrrolyldiketone PtII complexes with a helical π-electronic ligand were prepared as anion-responsive molecules. The synthesized chiral PtII complexes, showing weak phosphorescence emissions, exhibited anion-binding properties, and their electronic properties and assembly behaviours were modulated by anion binding. The crystal-state ion-pairing assemblies based on the anion complexes of the chiral PtII complexes include narcissistic self-sorting charge-by-charge columns comprising either of the enantiomers and those of the corresponding enantiomers. The chiral charge-by-charge columns are aligned in one direction to form layers that are alternately arranged with layers of opposite chirality. Further modifications of introduced π-electronic ligands would control the electronic states, enhancing the phosphorescence quantum efficiency. Charge-by-charge ion-pairing assemblies comprising single enantiomers of chiral PtII complexes are fascinating chiroptical materials.
Experimental
General information
Starting materials were purchased from FUJIFILM Wako Pure Chemical Corp., Nacalai Tesque Inc., Tokyo Chemical Industry Co., Ltd and Sigma-Aldrich Co. and were used without further purification unless otherwise stated. NMR spectra used in the characterization of products were recorded using a JEOL ECA-600 600 MHz spectrometer. All NMR spectra were referenced to the solvent. UV-visible absorption spectra were recorded using a Hitachi U-4100 spectrometer. Electrospray ionization mass spectrometry (ESI-MS) was performed with a BRUKER microTOF using the ESI-TOF method. TLC analyses were carried out on aluminum sheets coated with silica gel 60 (Merck 5554). Column chromatography was performed using Wakogel C-300.
(1,3-Dipyrrol-2-yl-1,3-propanedionato-κO1,κO3)-[1-(naphthalen-1-yl-κC)isoquinoline-κN]platinum, 2a
According to the literature procedure,4a in a dried Schlenk flask, [(PtMe2)2(SMe2)2]11 (31.4 mg, 0.054 mmol) was dissolved in THF (1.0 mL) under a N2 atmosphere. To the solution was added 1-(naphthalen-1-yl)isoquinoline10a (25.7 mg, 0.100 mmol). The resulting mixture was stirred for 1 h at r.t., and TfOH (8.8 μL, 0.100 mmol) was added dropwise. The reaction mixture was stirred for 30 min, and then a solution of 1,3-dipyrrol-2-yl-1,3-propanedione9a (21.2 mg, 0.104 mmol) and K2CO3 (19.8 mg, 0.143 mmol) in MeOH/THF (1:1, 2.0 mL) was added. The mixture was stirred for 4 h. After the removal of the solvent under vacuum, the residue was purified by column chromatography over silica gel (Wakogel C-300; eluent: CH2Cl2) to give 2a (11.3 mg, 0.017 mmol, 32%) as a yellow solid. Rf = 0.79 (CH2Cl2). 1H NMR (600 MHz, DMSO-d6, 20 °C): δ (ppm) 11.76 (br, 1H, NH), 11.55 (br, 1H, NH), 9.30 (d, J = 6.0 Hz, 1H, Ar–H), 8.18 (d, J = 7.8 Hz, 1H, Ar–H), 8.12 (m, 2H, Ar–H), 7.94 (d, J = 7.8 Hz, 1H, Ar–H), 7.89 (m, 2H, Ar–H), 7.85 (d, J = 7.8 Hz, 1H, Ar–H), 7.74 (d, J = 8.4 Hz, 1H, Ar–H), 7.60 (t, J = 7.2 Hz, 1H, Ar–H), 7.41 (t, J = 6.6 Hz, 1H, Ar–H), 7.34 (t, J = 8.4 Hz, 1H, Ar–H), 7.25 (m, 2H, pyrrole-H), 7.15 (m, 2H, pyrrole-H), 6.65 (s, 1H, CH), 6.29 (m, 2H, pyrrole-H). 13C NMR (151 MHz, DMSO-d6, 20 °C): δ (ppm) 170.25, 169.43, 168.79, 145.60, 140.01, 138.95, 137.38, 131.19, 131.81, 131.44, 130.82, 130.46, 129.69, 128.90, 128.85, 128.43, 127.11, 126.79, 125.26, 124.73, 124.16, 123.77, 123.38, 123.34, 119.77, 113.09, 112.62, 110.24, 110.10, 93.01. UV/vis (CH2Cl2, λmax[nm] (ε, 104 M−1 cm−1)): 380 (3.3). ESI-TOF-MS (HR): 649.1210. Calcd for C30H20N3O2Pt ([M − H]−): 649.1209. This compound was further characterized as the ion-pairing form by single-crystal X-ray analysis.
(1,3-Dipyrrol-2-yl-1,3-propanedionato-κO1,κO3)-[1-(8-phenylnaphthalen-1-yl-κC)isoquinoline-κN]platinum, 3a
According to the literature procedure,4a in a dried Schlenk flask, [(PtMe2)2(SMe2)2]11 (29.5 mg, 0.051 mmol) was dissolved in THF (1.0 mL) under a N2 atmosphere. To the solution was added 1-(8-phenylnaphthalen-1-yl)isoquinoline10b (33.4 mg, 0.101 mmol). The resulting mixture was stirred for 1 h at r.t., and TfOH (8.8 μL, 0.100 mmol) was added dropwise. The reaction mixture was stirred for 30 min, and then a solution of 1,3-dipyrrol-2-yl-1,3-propanedione9a (22.1 mg, 0.109 mmol) and K2CO3 (28.6 mg, 0.207 mmol) in MeOH/THF (1:1, 2.0 mL) was added. The mixture was stirred for 3 h. After the removal of the solvent under vacuum, the residue was purified by column chromatography over silica gel (Wakogel C-300; eluent: CH2Cl2) to give 3a (12.8 mg, 0.018 mmol, 35%) as a yellow solid. Rf = 0.79 (CH2Cl2). 1H NMR (600 MHz, DMSO-d6, 20 °C): δ (ppm) 11.74 (br, 1H, NH), 11.53 (br, 1H, NH), 8.98 (d, J = 6.0 Hz, 1H, Ar–H), 8.08 (d, J = 7.8 Hz, 1H, Ar–H), 7.96 (d, J = 8.4 Hz, 1H, Ar–H), 7.89 (d, J = 8.4 Hz, 1H, Ar–H), 7.69 (d, J = 8.4 Hz, 1H, Ar–H), 7.60 (d, J = 8.4 Hz, 1H, Ar–H), 7.54 (m, 2H, Ar–H), 7.45 (d, J = 6.6 Hz, 1H, Ar–H), 7.34 (d, J = 7.2 Hz, 1H, Ar–H), 7.28 (t, J = 7.2 Hz, 1H, Ar–H), 7.24 (br, 2H, pyrrole-H), 7.18 (s, 1H, pyrrole-H), 7.15 (s, 1H, pyrrole-H), 6.69 (m, 5H, Ar–H), 6.65 (s, 1H, CH), 6.29 (m, 2H, pyrrole-H). 13C NMR (151 MHz, DMSO-d6, 20 °C): δ (ppm) 170.61, 170.06, 168.75, 144.61, 142.43, 139.08, 138.91, 136.46, 135.73, 132.57, 131.43, 130.80, 130.47, 129.14, 128.98, 128.93, 128.36, 127.86, 126.78, 126.52, 126.31, 125.61, 125.08, 123.71, 123.66, 123.31, 118.51, 112.98, 112.55, 110.19, 110.06, 92.94 (four signals were overlapped). UV/vis (CH2Cl2, λmax[nm] (ε, 104 M−1 cm−1)): 387 (4.1). ESI-TOF-MS (HR): 725.1525. Calcd for C36H24N3O2Pt ([M − H]−): 725.1525. This compound was further characterized by single-crystal X-ray analysis.
(1,3-Di(5-phenylpyrrol-2-yl)-1,3-propanedionato-κO1,κO3)-[1-(8-phenylnaphthalen-1-yl-κC)isoquinoline-κN]platinum, 3b
According to the literature procedure,4a in a dried Schlenk flask, [(PtMe2)2(SMe2)2]11 (30.5 mg, 0.053 mmol) was dissolved in THF (1.0 mL) under a N2 atmosphere. To the solution was added 1-(8-phenylnaphthalen-1-yl)isoquinoline10b (35.2 mg, 0.113 mmol). The resulting mixture was stirred for 1 h at r.t., and TfOH (8.8 μL, 0.100 mmol) was added dropwise. The reaction mixture was stirred for 30 min, and then a solution of 1,3-di(5-phenylpyrrol-2-yl)-1,3-propanedione9b (38.9 mg, 0.110 mmol) and K2CO3 (20.7 mg, 0.150 mmol) in MeOH/THF (1:1, 2.0 mL) was added. The mixture was stirred for 3 h. After the removal of the solvent under vacuum, the residue was purified with column chromatography over silica gel (Wakogel C-300; eluent: CH2Cl2) to give 3b (9.11 mg, 0.010 mmol, 19%) as a yellow solid. Rf = 0.72 (CH2Cl2/n-hexane = 2/1 (v/v)). 1H NMR (600 MHz, DMSO-d6, 20 °C): δ (ppm) 11.80 (br, 1H, NH), 11.65 (br, 1H, NH), 9.09 (d, J = 6.6 Hz, 1H, Ar–H), 8.08 (d, J = 8.4 Hz, 1H, Ar–H), 7.96 (d, J = 8.4 Hz, 1H, Ar–H), 7.93 (m, 3H, Ar–H), 7.90 (d, J = 8.4 Hz, 2H, Ar–H), 7.68 (d, J = 8.4 Hz, 1H, Ar–H), 7.63 (d, J = 8.4 Hz, 1H, Ar–H), 7.55 (m, 3H, Ar–H), 7.48 (m, 4H, Ar–H), 7.41 (m, 1H, pyrrole-H), 7.35 (m, 3H, Ar–H), 7.29 (t, J = 8.4 Hz, 1H, Ar–H), 7.25 (m, 1H, pyrrole-H), 6.88 (s, 1H, CH), 6.80 (m, 2H, pyrrole-H), 6.70 (m, 5H, Ar–H). 13C NMR (151 MHz, DMSO-d6, 20 °C): δ (ppm) 170.57, 170.00, 168.05, 144.64, 142.47, 139.25, 139.20, 136.95, 136.52, 136.25, 135.80, 133.30, 132.61, 132.37, 131.89, 131.82, 130.84, 129.17, 129.06, 129.01, 128.85, 128.81, 128.17, 127.94, 127.91, 127.26, 127.24, 127.21, 127.11, 126.81, 126.56, 126.33, 125.69, 125.66, 125.54, 125.35, 125.21, 125.11, 123.74, 118.71, 115.33, 114.75, 108.95, 108.80, 94.15 (three signals were overlapped). UV/vis (CH2Cl2, λmax[nm] (ε, 104 M−1 cm−1)): 452 (6.5). ESI-TOF-MS (HR): 877.2148. Calcd for C48H32N3O2Pt ([M − H]−): 877.2148. This compound was further characterized as the ion-pairing form by single-crystal X-ray analysis.
Method for single-crystal X-ray analysis
Crystallographic data are summarized in Table 1. A single crystal of 3a was obtained by vapour diffusion of n-hexane into a CH2Cl2 solution. The data crystal was an orange block of approximate dimensions 0.05 mm × 0.05 mm × 0.03 mm. A single crystal of 2a·Cl−-TATA+ was obtained by vapour diffusion of n-hexane into a CH2Cl2 solution of a mixture of 2a and TATA+ as a Cl− salt (TATACl)15c in a 1:1 ratio. The data crystal was a red prism of approximate dimensions 0.10 mm × 0.03 mm × 0.02 mm. A single crystal of 3a·Cl−-TATA+ was obtained by vapor diffusion of n-hexane into a CH2Cl2 solution of a mixture of 3a and TATACl in a 1:1 ratio. The data crystal was an orange prism of approximate dimensions 0.04 mm × 0.02 mm × 0.01 mm. A single crystal of 3b·Cl−-TATA+ was obtained by vapour diffusion of n-hexane into a CH2Cl2 solution of a mixture of 3b and TATACl in a 1:1 ratio. The data crystal was an orange prism of approximate dimensions 0.100 mm × 0.050 mm × 0.020 mm. The data of 3b·Cl−-TATA+ were collected at 100 K using a DECTRIS PILATUS3 CdTe 1M diffractometer with Si (311) monochromated synchrotron radiation (λ = 0.4136 Å) at BL02B1 (SPring-8),21 whereas those of 3a and 2a·Cl−-TATA+ were collected at 90 K on a DECTRIS EIGER X 1 M diffractometer with Si (111) monochromated synchrotron radiation (λ = 0.802 and 0.8105 Å, respectively) at BL40XU (SPring-8).22 The data of 3a·Cl−-TATA+ were collected at 90 K using a Bruker D8 Venture diffractometer with MoKα radiation (λ = 0.71073 Å) focused with a multilayer confocal mirror. All the structures were solved by the dual-space method. The structures were refined by a full-matrix least-squares method using SHELXL 201423 (Yadokari-XG).24 In each structure, the non-hydrogen atoms were refined anisotropically. For 2a·Cl−-TATA+, the disordered solvents, presumably n-hexane, were removed using the SQUEEZE protocol included in PLATON.25
Table 1 Crystallographic details
|
3a
|
2a·Cl−-TATA+ |
3a·Cl−-TATA+ |
3b·Cl−-TATA+ |
Synchrotron radiation.
MoKα radiation.
|
Formula |
C36H25N3O2Pt·CH2Cl2 |
C30H21N3O2PtCl·C28H30N3·C6H14 |
C36H25N3O2PtCl·C28H30N3·2.772CH2Cl2 |
C48H33N3O2PtCl·C28H30N3·2CH2Cl2 |
Fw |
811.60 |
1180.76 |
1406.13 |
1492.71 |
Crystal size, mm |
0.05 × 0.05 × 0.03 |
0.10 × 0.03 × 0.02 |
0.04 × 0.02 × 0.01 |
0.100 × 0.050 × 0.020 |
Crystal system |
Triclinic |
Monoclinic |
Monoclinic |
Orthorhombic |
Space group |
P (no. 2) |
P21/n (no. 14) |
P21/c (no. 14) |
Pna21 (no. 33) |
a, Å |
9.721(7) |
7.925(8) |
13.9514(15) |
36.864(7) |
b, Å |
12.828(7) |
19.815(13) |
14.7192(14) |
21.004(4) |
c, Å |
12.984(7) |
32.173(18) |
30.005(3) |
8.479(17) |
α, ° |
71.794(12) |
90 |
90 |
90 |
β, ° |
87.54(2) |
95.72(7) |
102.175(4) |
90 |
γ, ° |
81.166(11) |
90 |
90 |
90 |
V, Å3 |
1519.8(16) |
5027(7) |
6023.0(10) |
6565(2) |
ρ
calcd, g cm−3 |
1.774 |
1.560 |
1.551 |
1.510 |
Z
|
2 |
4 |
4 |
4 |
T, K |
90(2) |
90(2) |
90(2) |
100(2) |
μ, mm−1 |
6.539a |
4.050a |
2.671b |
0.592a |
No. of reflns |
18108 |
52646 |
95598 |
188969 |
No. of unique reflns |
6690 |
8882 |
10676 |
14998 |
Variables |
406 |
616 |
742 |
832 |
λ, Å |
0.802a |
0.8105a |
0.71073b |
0.4136a |
R
1 (I > 2σ(I)) |
0.0556 |
0.0342 |
0.0610 |
0.0255 |
wR2 (I > 2σ(I)) |
0.1379 |
0.0870 |
0.1450 |
0.0651 |
GOF |
1.060 |
1.042 |
1.251 |
1.021 |
Calculations for the geometrical optimizations and energy decomposition analysis
DFT calculations for the theoretical optimizations, electrostatic potentials (ESPs), molecular orbitals and UV/vis spectra were carried out using the Gaussian 16 program.12 Energy decomposition analysis (EDA) calculations for the packing structures were performed using the GAMESS program.18
Data availability
The data that support the findings of this study are available in the ESI† of this article.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was supported by JSPS KAKENHI Grant Numbers JP22H02067 and JP23K23335 for Scientific Research (B), JP22K05253 and JP24K08389 for Scientific Research (C), JP23K17951 for Challenging Research (Exploratory) and JP20H05863 for Transformative Research Areas (A) “Condensed Conjugation” and Ritsumeikan Global Innovation Research Organization (R-GIRO) project (2022–27). Theoretical calculations were partially performed at the Research Center for Computational Science, Okazaki, Japan (22-IMS-C077, 23-IMS-C069 and 24-IMS-C067) and Information Initiative Center, Hokkaido University. We thank Dr Yuiga Nakamura, Dr Kouhei Ichiyanagi, Dr Toshiyuki Sasaki and Dr Nobuhiro Yasuda, JASRI/SPring-8 (2023A1240 and 2023B1755) for synchrotron radiation single-crystal X-ray analysis, Prof. Osamu Tsutsumi, Ritsumeikan University, for single-crystal X-ray analysis and Prof. Hitoshi Tamiaki, Ritsumeikan University, for various measurements including single-crystal X-ray analysis.
References
- Selected books and report:
(a)
Supramolecular Soft Matter, ed. T. Nakanishi, Wiley, 2011 Search PubMed;
(b)
Supramolecular Chemistry: From Molecules to Nanomaterials, ed. P. A. Gale and J. W. Steed, Wiley, 2012 Search PubMed;
(c)
Supramolecular Materials for Opto-Electronics, ed. N. Koch, RSC, 2015 Search PubMed;
(d) P. Yu, Y. Zhen, H. Dong and W. Hu, Crystal Engineering of Organic Optoelectronic Materials, Chem, 2019, 5, 2814 CrossRef.
- Reviews:
(a) V. W.-W. Yam, V. K.-M. Au and S. Y.-L. Leung, Light-Emitting Self-Assembled Materials Based on d8 and d10 Transition Metal Complexes, Chem. Rev., 2015, 115, 7589–7728 CrossRef PubMed;
(b) T. Strassner, Phosphorescent Platinum(II) Complexes with C^C* Cyclometalated NHC Ligands, Acc. Chem. Res., 2016, 49, 2680–2689 CrossRef PubMed;
(c) L. Ravotto and P. Ceroni, Aggregation induced phosphorescence of metal complexes: From principles to applications, Coord. Chem. Rev., 2017, 346, 62–76 CrossRef;
(d) E. V. Puttock, M. T. Walden and J. A. G. Williams, The luminescence properties of multinuclear platinum complexes, Coord. Chem. Rev., 2018, 367, 127–162 CrossRef;
(e) M. Yoshida and M. Kato, Cation-controlled luminescence behavior of anionic cyclometalated platinum(II) complexes, Coord. Chem. Rev., 2020, 408, 213194 CrossRef;
(f) Y. Han, Z. Gao, C. Wang, R. Zhong and F. Wang, Recent progress on supramolecular assembly of organoplatinum(II) complexes into long-range ordered nanostructures, Coord. Chem. Rev., 2020, 414, 213300 CrossRef.
-
(a) V. W.-W. Yam, K. M.-C. Wong and N. Zhu, Solvent-Induced Aggregation through Metal⋯Metal/π⋯π Interactions: Large Solvatochromism of Luminescent Organoplatinum(II) Terpyridyl Complexes, J. Am. Chem. Soc., 2002, 124, 6506–6507 CrossRef PubMed;
(b) Y. Sun, K. Ye, H. Zhang, J. Zhang, L. Zhao, B. Li, G. Yang, B. Yang, Y. Wang, S.-W. Lai and C.-M. Che, Luminescent One-Dimensional Nanoscale Materials with PtII⋯PtII Interactions, Angew. Chem., Int. Ed., 2006, 45, 5610–5613 CrossRef PubMed;
(c) N. Komiya, M. Okada, K. Fukumoto, D. Jomori and T. Naota, Highly Phosphorescent Crystals of Vaulted trans-Bis(salicylaldiminato)platinum(II) Complexes, J. Am. Chem. Soc., 2011, 133, 6493–6496 CrossRef;
(d) C.-H. Chen, F.-I. Wu, Y.-Y. Tsai and C.-H. Cheng, Platinum Phosphors Containing an Aryl-modified β-Diketonate: Unusual Effect of Molecular Packing on Photo- and Electroluminescence, Adv. Funct. Mater., 2011, 21, 3150–3158 CrossRef;
(e) N. Komiya, N. Itami and T. Naota, Solid-State Phosphorescence of trans-Bis(salicylaldiminato)platinum(II) Complexes Bearing Long Alkyl Chains: Morphology Control towards Intense Emission, Chem. – Eur. J., 2013, 19, 9497–9505 CrossRef PubMed;
(f) M. Mauro, A. Aliprandi, C. Cebrián, D. Wang, C. Kübel and L. De Cola, Self-assembly of a neutral platinum(II) complex into highly emitting microcrystalline fibers through metallophilic interactions, Chem. Commun., 2014, 50, 7269–7272 RSC;
(g) K. Ohno, S. Yamaguchi, A. Nagasawa and T. Fujihara, Mechanochromism in the luminescence of novel cyclometalated platinum(II) complexes with α-aminocarboxylates, Dalton Trans., 2016, 45, 5492–5503 RSC;
(h) M.-J. Sun, Y. Liu, W. Zeng, Y. S. Zhao, Y.-W. Zhong and J. Yao, Photoluminescent Anisotropy Amplification in Polymorphic Organic Nanocrystals by Light-Harvesting Energy Transfer, J. Am. Chem. Soc., 2019, 141, 6157–6161 CrossRef PubMed;
(i) H. Yang, H. Li, L. Yue, X. Chen, D. Song, X. Yang, Y. Sun, G. Zhou and Z. Wu, Aggregation-induced phosphorescence emission (AIPE) behaviors in PtII(C^N)(N-donor ligand)Cl-type complexes through restrained D2d deformation of
the coordinating skeleton and their optoelectronic properties, J. Mater. Chem. C, 2021, 9, 2334–2349 RSC;
(j) F.-F. Xu, W. Zeng, M.-J. Sun, Z.-L. Gong, Z.-Q. Li, Y. S. Zhao, J. Yao and Y.-W. Zhong, Organoplatinum(II) Cruciform: A Versatile Building Block to Fabricate 2D Microcrystals with Full-Color and White Phosphorescence and Anisotropic Photon Transport, Angew. Chem., Int. Ed., 2022, 61, e202116603 CrossRef PubMed.
-
(a) A. Kuno, G. Hirata, H. Tanaka, Y. Kobayashi, N. Yasuda and H. Maeda, Dipyrrolyldiketone PtII Complexes: Ion-Pairing π-Electronic Systems with Various Anion-Binding Modes, Chem. – Eur. J., 2021, 27, 10068–10076 CrossRef CAS PubMed;
(b) Y. Haketa, K. Komatsu, H. Sei, H. Imoba, W. Ota, T. Sato, Y. Murakami, H. Tanaka, N. Yasuda, N. Tohnai and H. Maeda, Enhanced solid-state phosphorescence of organoplatinum π-systems by ion-pairing assembly, Chem. Sci., 2024, 15, 964–973 RSC;
(c) Y. Haketa, Y. Murakami and H. Maeda, Ion-pairing assemblies of π-extended anion-responsive organoplatinum complexes, Sci. Technol. Adv. Mater., 2024, 25, 2313958 CrossRef.
-
(a) Y. Haketa, K. Urakawa and H. Maeda, First decade of π-electronic ion-pairing assemblies, Mol. Syst. Des. Eng., 2020, 5, 757–771 RSC;
(b) Y. Haketa, K. Yamasumi and H. Maeda, π-Electronic ion pairs: building blocks for supramolecular nanoarchitectonics via iπ–iπ interactions, Chem. Soc. Rev., 2023, 52, 7170–7198 RSC.
-
i
π–iπ Interactions for assemblies: Y. Sasano, H. Tanaka, Y. Haketa, Y. Kobayashi, Y. Ishibashi, T. Morimoto, R. Sato, Y. Shigeta, N. Yasuda, T. Asahi and H. Maeda, Ion-pairing π-electronic systems: ordered arrangement and noncovalent interactions of negatively charged porphyrins, Chem. Sci., 2021, 12, 9645–9657 RSC.
-
(a) G. Fu, Y. He, W. Li, B. Wang, X. Lü, H. He and W.-Y. Wong, Efficient polymer light-emitting diodes (PLEDs) based on chiral [Pt(C^N)(N^O)] complexes with near-infrared (NIR) luminescence and circularly polarized (CP) light, J. Mater. Chem. C, 2019, 7, 13743–13747 RSC;
(b) Q.-Y. Yang, H.-H. Zhang, X.-L. Han, S.-D. Weng, Y. Chen, J.-L. Wu, L.-Z. Han, X.-P. Zhang and Z.-F. Shi, Enhanced Circularly Polarized Luminescence Activity in Chiral Platinum(II) Complexes With Bis- or Triphenylphosphine Ligands, Front. Chem., 2020, 8, 303 CAS;
(c) M. Ikeshita, S. Furukawa, T. Ishikawa, K. Matsudaira, Y. Imai and T. Tsuno, Enhancement of Chiroptical Responses of trans-Bis[(β-iminomethyl)naphthoxy]platinum(II) Complexes with Distorted Square Planar Coordination Geometry, ChemistryOpen, 2022, 11, e202100277 CrossRef PubMed;
(d) M. Ikeshita, T. Yamamoto, S. Watanabe, M. Kitahara, Y. Imai and T. Tsuno, Circularly Polarized Luminescence of Chiral Platinum(II) Complexes with Tetradentate Salen Ligands, Chem. Lett., 2022, 51, 832–835 CrossRef;
(e) J. Marikubo and T. Tsubomura, Circularly Polarized Luminescence of Cyclometalated Platinum(II) Complex Excimers: Large Difference between Isomers, Inorg. Chem., 2022, 61, 17154–17165 CrossRef PubMed.
-
(a) L. Norel, M. Rudolph, N. Vanthuyne, J. A. G. Williams, C. Lescop, C. Roussel, J. Autschbach, J. Crassous and R. Réau, Metallahelicenes: Easily Accessible Helicene Derivatives with Large and Tunable Chiroptical Properties, Angew. Chem., Int. Ed., 2010, 49, 99–102 CrossRef;
(b) E. Anger, M. Rudolph, L. Norel, S. Zrig, C. Shen, N. Vanthuyne, L. Toupet, J. A. G. Williams, C. Roussel, J. Autschbach, J. Crassous and R. Réau, Multifunctional and Reactive Enantiopure Organometallic Helicenes: Tuning Chiroptical Properties by Structural Variations of Mono- and Bis(platinahelicene)s, Chem. – Eur. J., 2011, 17, 14178–14198 CrossRef;
(c) C. Shen, E. Anger, M. Srebro, N. Vanthuyne, K. K. Deol, T. D. Jefferson, G. Muller, J. A. G. Williams, L. Toupet, C. Roussel, J. Autschbach, R. Reau and J. Crassous, Straightforward access to mono- and bis-cycloplatinated helicenes displaying circularly polarized phosphorescence by using crystallization resolution methods, Chem. Sci., 2014, 5, 1915–1927 RSC;
(d) J. R. Brandt, X. Wang, Y. Yang, A. J. Campbell and M. J. Fuchter, Circularly Polarized Phosphorescent Electroluminescence with a High Dissymmetry Factor from PHOLEDs Based on a Platinahelicene, J. Am. Chem. Soc., 2016, 138, 9743–9746 CrossRef CAS PubMed;
(e) T. Biet, T. Cauchy, Q. Sun, J. Ding, A. Hauser, P. Oulevey, T. Bürgi, D. Jacquemin, N. Vanthuyne, J. Crassous and N. Avarvari, Triplet state CPL active helicene–dithiolene platinum bipyridine complexes, Chem. Commun., 2017, 53, 9210–9213 RSC;
(f) Z.-P. Yan, X.-F. Luo, W.-Q. Liu, Z.-G. Wu, X. Liang, K. Liao, Y. Wang, Y.-X. Zheng, L. Zhou, J.-L. Zuo, Y. Pan and H. Zhang, Configurationally Stable Platinahelicene Enantiomers for Efficient Circularly Polarized Phosphorescent Organic Light-Emitting Diodes, Chem. – Eur. J., 2019, 25, 5672–5676 CrossRef PubMed;
(g) P. Vázquez-Domínguez, O. Journaud, N. Vanthuyne, D. Jacquemin, L. Favereau, J. Crassous and A. Ros, Helical donor–acceptor platinum complexes displaying dual luminescence and near-infrared circularly polarized luminescence, Dalton Trans., 2021, 50, 13220–13226 RSC;
(h) D. Tauchi, T. Koida, Y. Nojima, M. Hasegawa, Y. Mazaki, A. Inagaki, K. Sugiura, Y. Nagaya, K. Tsubaki, T. Shiga, Y. Nagata and H. Nishikawa, Aggregation-induced circularly polarized phosphorescence of Pt(II) complexes with an axially chiral BINOL ligand, Chem. Commun., 2023, 59, 4004–4007 RSC.
-
(a) H. Meada and Y. Kusunose, Dipyrrolyldiketone Difluoroboron Complexes: Novel Anion Sensors With C-H⋯X− Interactions, Chem. – Eur. J., 2005, 11, 5661–5666 CrossRef PubMed;
(b) H. Meada, Y. Haketa and T. Nakanishi, Aryl-Substituted C3-Bridged Oligopyrroles as Anion Receptors for Formation of Supramolecular Organogels, J. Am. Chem. Soc., 2007, 129, 13661–13674 CrossRef PubMed.
-
(a) A. Ros, B. Estepa, R. Lopez-Rodriguez, E. Alvarez, R. Fernandez and J. M. Lassaletta, Use of Hemilabile N,N Ligands in Nitrogen-Directed Iridium-Catalyzed Borylations of Arenes, Angew. Chem., Int. Ed., 2011, 50, 11724–11728 CrossRef PubMed;
(b) Z. Domínguez, R. López-Rodríguez, E. Álvarez, S. Abbate, G. Longhi, U. Pischel and A. Ros, Azabora[5]helicene Charge-Transfer Dyes Show Efficient and Spectrally Variable Circularly Polarized Luminescence, Chem. – Eur. J., 2018, 24, 12660–12668 CrossRef PubMed.
- G. S. Hill, M. J. Irwin, C. J. Levy, L. M. Rendina and R. J. Puddephatt, Platinum(II) Complexes of Dimethyl Sulfide, Inorg. Synth., 1998, 32, 149–151 CrossRef CAS.
-
M. J. Frisch, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 16, Revision C.01, Gaussian, Inc., Wallingford CT, 2016 Search PubMed.
- L. Yang, D. R. Powell and R. P. Houser, Structural variation in copper(I) complexes with pyridylmethylamide ligands: structural analysis with a new four-coordinate geometry index, τ4, Dalton Trans., 2007, 955–964 RSC.
- L. Di Bari, G. Pescitelli and P. Salvadori, Conformational Study of 2,2′-Homosubstituted 1,1′-Binaphthyls by Means of UV and CD Spectroscopy, J. Am. Chem. Soc., 1999, 121, 7998–8004 CrossRef.
-
(a) B. W. Laursen and F. C. Krebs, Synthesis of a Triazatriangulenium Salt, Angew. Chem. Int. Ed., Angew. Chem. Int. Ed., 2000, 39, 3432–3434 CrossRef;
(b) B. W. Laursen and F. C. Krebs, Synthesis, Structure, and Properties of Azatriangulenium Salts, Chem. Eur. J., 2001, 7, 1773–1783 CrossRef PubMed;
(c) Y. Haketa, S. Sasaki, N. Ohta, H. Masunaga, H. Ogawa, N. Mizuno, F. Araoka, H. Takezoe and H. Maeda, Oriented Salts: Dimension-Controlled Charge-by-Charge Assemblies from Planar Receptor–Anion Complexes, Angew. Chem., Int. Ed., 2010, 49, 10079–10083 CrossRef PubMed.
- P. R. Spackman, M. J. Turner, J. J. McKinnon, S. K. Wolff, D. J. Grimwood, D. Jayatilaka and M. A. Spackman,
CrystalExplorer: a program for Hirshfeld surface analysis, visualization and quantitative analysis of molecular crystals, J. Appl. Crystallogr., 2021, 54, 1006–1011 CrossRef PubMed.
- M. J. S. Phipps, T. Fox, C. S. Tautermann and C.-K. Skylaris, Energy decomposition analysis approaches and their evaluation on prototypical protein–drug interaction patterns, Chem. Soc. Rev., 2015, 44, 3177–3211 RSC.
- Articles for GAMESS:
(a) M. W. Schmidt, K. K. Baldridge, J. A. Boatz, S. T. Elbert, M. S. Gordon, J. H. Jensen, S. Koseki, N. Matsunaga, K. A. Nguyen, S. Su, T. L. Windus, M. Dupuis and J. A. Montgomery Jr., General atomic and molecular electronic structure system, J. Comput. Chem., 1993, 14, 1347–1363 CrossRef CAS;
(b)
M. S. Gordon and M. W. Schmidt, in Theory and Applications of Computational Chemistry: The First Forty Years, ed. C. E. Dykstra, G. Frenking, K. S. Kim and G. E. Scuseria, Elsevier, 2005 Search PubMed.
- Report for FMO: K. Kitaura, E. Ikeo, T. Asada, T. Nakano and M. Uebayasi, Fragment molecular orbital method: an approximate computational method for large molecules, Chem. Phys. Lett., 1999, 313, 701–706 CrossRef.
- Report for pair interaction energy decomposition analysis (PIEDA): D. G. Fedorov and K. Kitaura, Pair interaction energy decomposition analysis, J. Comput. Chem., 2007, 28, 222–237 CrossRef PubMed.
- K. Sugimoto, H. Ohsumi, S. Aoyagi, E. Nishibori, C. Moriyoshi, Y. Kuroiwa, H. Sawa and M. Takata, Extremely High Resolution Single Crystal Diffractometory for Orbital Resolution using High Energy Synchrotron Radiation at SPring-8, AIP Conf. Proc., 2010, 1234, 887–890 CrossRef.
-
(a) N. Yasuda, H. Murayama, Y. Fukuyama, J. E. Kim, S. Kimura, K. Toriumi, Y. Tanaka, Y. Moritomo, Y. Kuroiwa, K. Kato, H. Tanaka and M. Takata, X-ray diffractometry for the structure determination of a submicrometre single powder grain, J. Synchrotron Radiat., 2009, 16, 352–357 CrossRef;
(b) N. Yasuda, Y. Fukuyama, K. Toriumi, S. Kimura and M. Takata, Submicrometer Single Crystal Diffractometry for Highly Accurate Structure Determination, AIP Conf. Proc., 2010, 1234, 147–150 CrossRef.
- G. M. Sheldrick, A short history of SHELX, Acta Crystallogr., Sect. A: Found. Crystallogr., 2008, 64, 112–122 CrossRef PubMed.
-
(a)
K. Wakita, Yadokari-XG, Software for Crystal Structure Analyses, 2001 Search PubMed;
(b) C. Kabuto, S. Akine, T. Nemoto and E. Kwon, Release of Software (Yadokari-XG 2009) for Crystal Structure Analyses, J. Crystallogr. Soc. Jpn., 2009, 51, 218–224 CrossRef.
- A. L. Spek,
PLATON SQUEEZE: a tool for the calculation of the disordered solvent contribution to the calculated structure factors, Acta Crystallogr., Sect. C: Struct. Chem., 2015, 71, 9–18 CrossRef PubMed.
Footnote |
† Electronic supplementary information (ESI) available: Synthetic procedures, analytical data, computational details and cif files for the single-crystal X-ray analysis. CCDC 2381822–2381825. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4qo01697a |
|
This journal is © the Partner Organisations 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.