Next Article in Journal
Nano-MnO2 Decoration of TiO2 Microparticles to Promote Gaseous Ethanol Visible Photoremoval
Previous Article in Journal
Preparation of Nickel Nanoparticles by Direct Current Arc Discharge Method and Their Catalytic Application in Hybrid Na-Air Battery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phosphorescent Modulation of Metallophilic Clusters and Recognition of Solvents through a Flexible Host-Guest Assembly: A Theoretical Investigation

1
College of Chemical Engineering and Technology, Key Laboratory for New Molecule Design and Function of Gansu Universities, Tianshui Normal University, Tianshui 741001, China
2
School of Electronic Information and Electrical Engineering, Tianshui Normal University, Tianshui 741001, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2018, 8(9), 685; https://doi.org/10.3390/nano8090685
Submission received: 15 August 2018 / Revised: 27 August 2018 / Accepted: 28 August 2018 / Published: 2 September 2018

Abstract

:
MP2 (Second order approximation of Møller–Plesset perturbation theory) and DFT/TD-DFT (Density functional theory/Time-dependent_density_functional_theory) investigations have been performed on metallophilic nanomaterials of host clusters [Au(NHC)2]+⋅⋅⋅[M(CN)2]⋅⋅⋅[Au(NHC)2]+ (NHC = N-heterocyclic carbene, M = Au, Ag) with high phosphorescence. The phosphorescence quantum yield order of clusters in the experiments was evidenced by their order of μS1ES1−T1 values ( μ S 1 : S0 → S1 transition dipole, E S 1 T 1 : splitting energy between the lowest-lying singlet S1 and the triplet excited state T1 states). The systematic variation of the guest solvents (S1: CH3OH, S2: CH3CH2OH, S3: H2O) are employed not only to illuminate their effect on the metallophilic interaction and phosphorescence but also as the probes to investigate the recognized capacity of the hosts. The simulations revealed that the metallophilic interactions are mainly electrostatic and the guests can subtly modulate the geometries, especially metallophilic Au⋅⋅⋅M distances of the hosts through mutual hydrogen bond interactions. The phosphorescence spectra of hosts are predicted to be blue-shifted under polar solvent and the excitation from HOMO (highest occupied molecular orbital) to LUMO (lowest unoccupied molecular orbital) was found to be responsible for the 3MLCT (triplet metal-to-ligand charge transfer) characters in the hosts and host-guest complexes. The results of investigation can be introduced as the clues for the design of promising blue-emitting phosphorescent and functional materials.

1. Introduction

The relationship between the luminescent property and metallophilic Au(I)⋅⋅⋅M bond distance (M = AuI, AgI, CuI, TlI, HgII, BiIII, etc.) has attracted a great deal of attention in the last few years [1,2,3,4,5]. Many experimental and theoretical studies indicated that Au(I) complexes luminesce strongly, especially when the metallophilic interaction is present [6,7,8,9]. The organometallic complexes were applied extensively as emitters in organic light-emitting diodes (OLEDs) and phosphorescent OLEDs with green and red spectral range, which have already been demonstrated to be high efficiency and stability [10,11,12]. However, the blue-emitting OLEDs, which are essential for the commercial launch of devices for lighting, still lack stability and efficiency. Indeed, designing new materials to show higher energy, such as deep-blue emission, encounters more obstacles than the progress made for obtaining green and red colors.
Considerable investigations have been carried out in developing blue OLEDs with high external quantum efficiency as well as a deeper blue color [13,14,15,16,17]. Recently, the highly phosphorescent organometallic nanomaterials of polymeric double salts [Au(NHC)2][M(CN)2] (NHC = N-heterocyclic carbene, M = Au or Ag) were prepared, which can provide a deep blue shifted phosphorescence spectrum with emission quantum yields of up to 90% [2]. It was found that the extended metallophilic d10⋅⋅⋅d10 interactions played significant role in the phosphorescence of quasi-2D polymeric nanostructures.
Metallophilic attractions behave similarly to hydrogen bonds [18] which can be easily disrupted/modulated by other crystal packing motifs [18,19,20,21]. Furthermore, the photoluminescent characters might be modulated as metallophilic interaction changed by the functional guest molecules/ions introduced into the M···M interaction systems. Some groups [4,21,22,23,24,25,26,27,28,29,30,31,32] have made significant progress in this area. For example, Catalano et al. introduced BF4 and CH3OH through anion-cation/anion–π interaction to the hetero-metallic coordinated [AuCu-(PPh2py)3](BF4)2 Au(I)···Cu(I) complex [4]. López-de-Luzuriaga and coworkers [30] employed halogen bonding to bimetallic gold–silver clusters. Dichloromethane and tetrahydrofuran also can tune the phosphorescent properties of a Ag/Au metallophilic tetranuclear complex [31]. The coexisting cations [32] can considerably enhance the emission yields of [Au(CN)2−] oligomers in aqueous solutions. However, these investigations only observed red-shift phosphorescence and therefore exploration of the blue and highly efficient phosphorescent materials is essential [2].
In the present paper, the metallophilic characters and the phosphorescent properties of the host clusters were investigated by density functional theory (DFT). Introducing the guest solvents CH3OH (S1), CH3CH2OH (S2), and H2O (S3), we mainly focus on: (a) the nature of homo/hetero-metallophilic interactions; (b) how the guest molecule affects the metallophilic distance, and (c) how the metallophilic distance affects the photophysical properties of the complexes. Our calculated results reported herein are predicted to provide blue-shift phosphorescence, which is helpful for the further synthesis of the organometallic compounds for blue-emitting OLEDs.

2. Computational Methodology

Although metallophilic interactions have been investigated with a series of quantum chemical methods, the size of the systems studied renders the use of more computationally expensive methods prohibitive and unfeasible [33,34,35,36,37,38,39,40,41,42,43,44]. The B3LYP [45,46,47], M06-2X [48], ωB97XD [49], B3LYP-D3 [50,51], PBE0 [52,53,], and MP2 [54,55] methods were employed respectively to optimize the polymeric ground-state geometries of clusters X (X = I, II, and III, see Figure 1) on the basis of the experimental X-ray structures. Two basis sets (BS1: SDD pseudopotential and basis set for Au and Ag, 6-31G(d) for other atoms; BS2: SDD pseudopotential and basis set for Au and Ag, 6-31+G(d,p) for other atoms) were used to obtain reliable geometries of complexes [56,57,58,59,60]. The ground state (S0), the lowest excited states (singlet: S1; triplet: T1) were fully optimized. Triplet states were calculated at the spin unrestricted UPBE0 level with a spin multiplicity of 3.
The MP2 methods [58,59,60,61,62] with BS1, BS3 (aug-cc-pVDZ-PP [63] pseudopotential and basis set for Au and Ag, 6-311++G(d,p) [64,65] for other atoms), and BS4 (aug-cc-pVTZ-PP pseudopotential and basis set for Au and Ag, 6-311++G(d,p) for other atoms) were also used to investigate the interaction energies with basis set superposition error (BSSE), ECP [66]. The different contributions of the interaction energies were obtained by using the GKS-EDA (generalized Kohn−Sham based energy decomposition analysis) [67] methods on the GAMESS [68] platform using B3LYP-D3 with BS5 (MCP-TZP [69,70,71,72,73,74] for Au and Ag atoms, MCP-DZP [74,75] for other atoms) and BS6 (MCP-TZP [69,70,71,72,73,74] for all atoms) basis sets.
The geometries of all the structures were fully optimized using the GAUSSIAN09 program suite [76]. The orbital composition analysis is performed by the Multiwfn 3.3 suite of program [77]. The natural bond orbital (NBO) analysis is achieved by NBO 5.0 procedure [78].

3. Results and Discussion

We initially optimized the representative host cluster I which has seven oligomeric units (Figure S1, see SI) and then calculated II and III with a similar procedure. The geometrical parameters of the ground state (S0) for clusters I, II, and III at various calculated levels are in good agreement with their experimental X-ray structures, respectively. After further extensive testing, PBE0/BS1 was employed in the subsequent qualitative analysis as it was found to be time-economical and reliable to evaluate the geometrical, electronic, and spectral properties of weakly bound metal complexes (Table S1, see SI) [42,43,44,79,80].

3.1. Ground States Properties

3.1.1. Structures of Clusters X

It should be noted that [Au(NHC)2]+ is approached vertically to [Au(CN)2] with metallophilic interaction Au···Au in I. The Au⋅⋅⋅Au distance of 3.11 Å between [Au(NHC)2]+- [Au(CN)2] reveals the aurophilic interaction in cluster I [56].
Clusters II and III have similar geometries to I. That is, the anion [Au(CN)2] lies vertically between two cations [Au(NHC)2]+ and the distances of Au⋅⋅⋅Au and Au⋅⋅⋅Ag are 3.15 and 3.06 Å for II and III, respectively. Therefore, the metallophilic interaction is decreased in II while it is strengthened in III, as compared with I. Both ∠CMC of anion [M(CN)2] and ∠AuMAu (M = Au, Ag) are 180.0° in cluster X, which indicates that three metal atoms together with two CN ions are in the same plane. The Au⋅⋅⋅M distances in X are shorter at least by 0.2 Å than that reported for the tetrameric CF3Au⋅CO [5,81], which suggests stronger metallophilic interactions in X than that in [CF3Au⋅CO]4.
The NBO results show that the metallophilic interactions between two adjacent metal atoms are all synergistically bidirectional (outward and inward), which is displayed using cluster I as an example (Figure 2, Table S2 in SI). In outward aurophilic interactions, the electron is delocalized from the LP(5) orbital of central Au2 atom to the LP*(7) orbitals of two lateral Au1 and Au3 atoms E i j ( 2 ) : ~30 kcal/mol respectively), whereas the electron is donated back from the LP(4) orbitals of Au1 and Au3 atoms to the LP*(6) orbital of central Au1 ( E i j ( 2 ) : ~28 kcal/mol respectively) in the inward aurophilic interactions. The E i j ( 2 ) of outward (62.83 kcal/mol) is 7 kcal/mol higher than that of inward (55.83 kcal/mol), and also, NBO results show that whether in inward or outward modes, the d orbits of Au atoms act as the electron donors and the electron acceptors originate from the p orbits of Au atoms.

3.1.2. Structures of X⋅⋅⋅(S)x Complexes

In this section, the solvents CH3OH (S1), CH3CH2OH (S2), and H2O (S3) are introduced to explore how they affect the metallophilic interactions and the phosphorescent properties of X. Figure 3 shows the formation model of complexes X⋅⋅⋅(S)x (X = IIII; x = 1–4). The skeletal diagram and key structural parameters are also listed in Figure S2 and Table S3, respectively.
Compared to free I, the structure of segment I in complex I⋅⋅⋅S1 is changed. There are mutual H⋅⋅⋅N and O⋅⋅⋅H hydrogen bonds in complex I⋅⋅⋅S1 and which leads NHC ligands of [Au(NHC)2]+ to be torsional and the distances of Au1⋅⋅⋅M2 and M2⋅⋅⋅Au3 are not equal anymore.
Figure 3 shows there are three complexation modes between X and two solvent molecules. The energy analysis revealed that mode 1 of X⋅⋅⋅(S)2 is ~ 3 kcal/mol lower than those in modes 2 and 3. Therefore, the following discussions focus on the mode 1. In I⋅⋅⋅(S1)2 (Figure 4), two Au⋅⋅⋅Au distances are elongated to 3.46 Å and the angles ∠H1Au2H2 and ∠Au1Au2Au3 both are 180°. Besides the aurophilic interaction, two types of hydrogen bonds are also involved: two N⋅⋅⋅H–O hydrogen bonds (1.87 Å) and four O⋅⋅⋅H–C bonds (2.42 Å). The I⋅⋅⋅(S1)2 is complexed together by hydrogen bond interactions.
For I⋅⋅⋅(S1)3, both neighbor Au⋅⋅⋅Au bond distances are 3.40 Å and there are three N⋅⋅⋅H–O (~ 1.9 Å) and six O⋅⋅⋅H–C hydrogen bonds (2.3–2.5 Å). Different from the I⋅⋅⋅(S1)2, the angle ∠Au1Au2Au3 of I⋅⋅⋅(S1)3 is decreased to 158.5°, which reflects that the three gold atoms are not in the same plane anymore.
Complex I···(S1)4 possesses D2h symmetry. The skeletons of [Au(CN)2] and S1 lie in the same plane and they are encapsulated by two [Au(NHC)2]+ segments. The fourth S1 for I⋅⋅⋅(S1)3, the Au⋅⋅⋅Au (3.25 Å) is decreased by 0.15 Å and the N⋅⋅⋅H and O⋅⋅⋅H bonds are 1.95 Å and 2.47 Å, respectively.
The structures for I⋅⋅⋅(S2)x and I⋅⋅⋅(S3)x are similar to those of corresponding I⋅⋅⋅(S1)x with the same x. For example, the Au⋅⋅⋅Au distances are 3.50, 3.45, 3.40, and 3.25Å for I⋅⋅⋅S2, I⋅⋅⋅(S2)2, I⋅⋅⋅(S2)3, and I⋅⋅⋅(S2)4, which are relatively close to those for I⋅⋅⋅(S3)x. It can be seen from Table S3, the other two metallophilic clusters II and III can be also adjusted by solvent and their structures are also similar to the those of I⋅⋅⋅(S)x respectively. Therefore, the Au⋅⋅⋅M (M = Au, Ag) distances are dramatically elongated with one S molecule inserted between two [Au(NHC)2]+. However, metallophilic Au⋅⋅⋅M distance is decreased gradually as the number of S x increased. The phosphorescence character then can be predicted to be modulated with the Au⋅⋅⋅M distance changed [1,2,3,4,5].

3.1.3. Interaction Energies

Interaction energies of Au⋅⋅⋅M in clusters X

The BSSE-corrected interaction energy (obtained from the electronic energy) computations using the GAUSSIAN09 program will be denoted with the superscript CP. The E add CP (added) and E tot CP (total) interaction energies defined in Equations (1) and (2) [5] are investigated.
E add CP [ X ] = E [ X ] E [ ( Au ( NHC ) 2 ) + ( M ( CN ) 2 ) ] E [ ( Au ( NHC ) 2 ) + ] + BSSE
E tot CP [ X ] = E [ X ] 2 E [ ( Au ( NHC ) 2 ) + E ( M ( CN ) 2 ) ] + BSSE
The calculated level test showed that the PBE0-D3/BS3 method is the most reliable and financial in time to estimate ECP. Table 1 shows that the interaction energies are very close for the isolated I, II, and III, in which E add CP and E tot CP are ~−30 and ~−102 kcal/mol, respectively. The clusters X E tot CP values for two metallophilic bonds are 3–4 times of those of E add CP of X, suggesting a degree of cooperativity [5,82,83,84].
The EDA interaction energies of clusters IIII are reported in Table 2, where the contributors Ees, Eex, Epol, Edisp, and Ecorr are attractive and Erep is repulsive. It should be noted that Eadd and Etot values are very close to their corresponding electrostatic energy (Ees) term, indicating that the clusters are mainly stabilized by the electrostatic interactions [5,84].

Interaction energies of complexes X⋅⋅⋅(S)x

In this section, the total interaction energies ECP and the ECPn (n = 1, 2, 3, and 4) are considered. The ECP1, ECP2, ECP3, and ECP4 respectively correspond to the interaction energies between one, two, three, and four S molecules and the remainder parts in X⋅⋅⋅(S)x (Table 3). For I⋅⋅⋅(S1)x, the ECP1 value is decreased from −16.6 to −11.8 kcal/mol as the number of S1 increasing from one to four. The ECPn value for I⋅⋅⋅(S2)x is very similar with the case in I⋅⋅⋅(S1)x, indicating that the solvents S1 and S2 behave similar in controlling the interaction energies for I⋅⋅⋅(S)x. Furthermore, the solvents S1 and S2 also act very similarly to adjust the interaction energy for II⋅⋅⋅(S)x or III⋅⋅⋅(S)x. Therefore, the interaction energy might not be adjusted by the carbon chain growth. All ECPn values for X⋅⋅⋅(S)x are negative, which shows S can stabilize the X⋅⋅⋅(S)x complexes.
The GKS-EDA results (Figure S3a–c) of complexes X⋅⋅⋅(S2)x [85] further reveal that the solvents S1 and S2 contribute very similar behavior to change the interaction energy for II⋅⋅⋅(S)x or III⋅⋅⋅(S)x. In other words, the EDA values have small differences between X⋅⋅⋅(S1)x and X⋅⋅⋅(S2)x at the same x with the same X. The EDA results showed that Eex is the most important energy component for all complexes except X⋅⋅⋅S2, in which electrostatic force is most important (Table S4).
Interestingly, the energy contributors of Eex, Erep, Epol, Edisp, and Ecorr are summed close to zero and an excellent correlation R = 1.00 between the Etot and the Ees values was found for I⋅⋅⋅(S2)x, II⋅⋅⋅(S2)x, and III⋅⋅⋅(S2)x with the linear equation (Figure S3a′–c′). These results further reinforce our finding that the investigated interaction is governed by the electrostatic term (Figure S4).

3.2. Excited State’s Properties

The M-related bond lengths, bond angles, and the major geometrical changes between the ground state S0 and lowest lying triplet excited state T1 are summarized in Table 4. The parameters reveal that the Au1⋅⋅⋅M2 and Au3⋅⋅⋅M2 bonds are shorter maximally by ~0.40 Å for T1 than those for the corresponding S0 of the cluster X. The metallophilic Au⋅⋅⋅M distance in S0 can be increased by introducing S into cluster X. However, the Au⋅⋅⋅M distances are shortened by 0.35 Å for triplet III⋅⋅⋅(S1)4 and even by 0.80 Å for triplet III⋅⋅⋅S1 as compared with the corresponding S0 of III⋅⋅⋅(S)x, respectively.
Table 5 shows the calculated emission energies, the electron transition assignments, and the experimental values of complexes IIII. The calculated emission wavelengths of 457, 483, and 417 nm are in good agreement with the experimental emission values of 448, 465, and 446 nm, which respectively correspond to the clusters I, II, and III. The electron transition from HOMO to LUMO is responsible for the emission at 457 nm for I. The HOMO of cluster I mainly consist of natural atomic orbital (NAO) of Au (77.23%, d: 45.29%), while LUMO has less NAO of Au (35.71%) and more significant NAO of the ligands (64.29%), which is MLCT character from the metal-centred to ligands excited states (Table S5). II and III also display similar MLCT phosphorescence nature to I (Table S6).
The phosphorescent emission wavelengths can be tuned as the Au–M distances change with the solvent effect (Table S3). The Au⋅⋅⋅M distances in S0 of X⋅⋅⋅(S)x are lengthened compared to those in free X. However, the Au⋅⋅⋅M distances in X⋅⋅⋅(S)x are dramatically shorter in the T1 state than those in S0. Actually, Au⋅⋅⋅M distances in X⋅⋅⋅(S)x are changed little compared with those in free X no matter with the type and the number of S in the triplet state, which is significantly different from the case in S0. Therefore, the Au⋅⋅⋅M distance change in S0 as x increases can reflect the photoluminescence change since the phosphorescent emission involves electron transition from T1 to S0.
It can be seen obviously that the photoluminescence of X can be tuned (Δλ) by Au⋅⋅⋅M distance change (Δd) through introducing S to X, and photoluminescence of the X⋅⋅⋅(S)x complexes shows blue shift compared with free X (Table S3 and Figure S5). Especially, X⋅⋅⋅(S)4 (X = I, III) provide the largest blue-shifted photoluminescence, while II⋅⋅⋅(S)4 has the smallest blue-shifted emission wavelengths. The photoluminescent emission wavelength for II⋅⋅⋅S is significantly different from those for II⋅⋅⋅(S)x (x = 2, 3, 4) because I is distorted in I⋅⋅⋅S. However, similar to the photoluminescence of X, the excitations from HOMO to LUMO of X⋅⋅⋅(S)x are still responsible for their emission at their maximum wavelength which mainly originates from MLCT character between metal-centred and ligands (Tables S5 and S6). The MLCT character becomes most obvious at x = 4 because of the charge transfer capacity from metal-centred to ligands of ~38%, ~44%, and ~31% for I⋅⋅⋅(S)4, II⋅⋅⋅(S)4, and III⋅⋅⋅(S)4, respectively. Moreover, different solvents, S1 and S2, have little effect on MLCT character of X at specific x. For example, the charge transfer capacity for I⋅⋅⋅(S1)4 and I⋅⋅⋅(S2)4 are respectively 37.63% and 37.60% (Table S5) so that the maximum photoluminescent wavelengths of them are also the same with 412 nm.
The radiative (kr) and the non-radiative (knr) rate constants are linked to the phosphorescence quantum yield (ΦPL) from an emissive excited state to the ground state by Equation (3).
ΦPL = kr/(kr + knr)
Theoretically, kr is related to the mixing between S1 and T1, which is proportional to the spin-orbit coupling (SOC) rate constants. SOC rate constants are linked to the phosphorescence quantum efficiency between the two states, according to Equation (4).
k r = γ Ψ S 1 | H S 0 | Ψ S 0 2 · μ S 1 2 ( Δ E S 1 T 1 ) 2 ,   with   γ = 16 π 2 10 6 n 3 E em 3 3 h ε 0
In general, smaller Δ E S 1 T 1 and larger μ S 1 or μ S 1 Δ E S 1 T 1 for the system could induce a higher ΦPL [86]. The μS1ES1−T1 values for complexes X are in the order: 0.99 (I) 0.91 (II) 0.83 (III), which is rationalized by the ΦPL in experimental order: 90% (I) 67% (II) 11% (III). An excellent correlation (R = 0.943) was observed between the theoretical μ S 1 Δ E S 1 T 1 values and the experimental ΦPL values with the linear equation ΦPL = −393.3 − 493.8 × μ S 1 Δ E S 1 T 1 for free X (Figure S6). We further employ this equation to predict the ΦPL values of X⋅⋅⋅(S)x, which reveal that fourteen complexes have high phosphorescence efficiency with ΦPL larger than 50% (Table S7, see SI).

4. Conclusions

In summary, a detailed study of metallophilic Au⋅⋅⋅M bonding in host clusters from self-assembled [Au(NHC)2]+ and [M(CN)2] (M = Au, Ag) and the characters of their host–guest complexes are presented by the second-order Møller−Plesset (MP2) method, density functional theory, and qualitative analysis via GKS-EDA and NBO methods with a series of basis sets.
The largest and smallest μS1ES1−T1 values for cluster I and III, respectively, were the highest and lowest ΦPL among three experimental clusters IIII. Based on host–guest complexation, the phosphorescence of hosts can be modulated and the guest solvents can be recognized. The metallophilic interactions are mainly derived from electrostatic interaction. The solvents methanol, ethanol, and water can adjust the geometries of Au(I)⋅⋅⋅Au(I)/Ag(I) clusters with H-bond interaction between the cluster and the solvents, especially changing the distance between two neighbor metal centres, leading to blue-shift phosphorescence of the clusters. These results highlight that the metal⋅⋅⋅metal interaction and the photoluminescence characters of the clusters can be functionally controlled by solvent molecules. The linear relationship between Δλ and Δd and the binding energies of host–guest complexes suggest that the phosphorescence wavelength shift can be predicted through the M⋅⋅⋅M distances in T1 states, and the interactions in clusters can offer potential applications in solvent and catalysis transport and recognition using synthetic functional materials in the future. This work will be expanded upon by employing related binuclear complexes under systematic variation of the metal ions and solvents.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/8/9/685/s1, Figure S1: Structure diagrams of clusters, Figure S2: Top view (a) and face view (b) of the representative cluster X⋅⋅⋅(S)x (X = IIII; x = 1–4), Figure S3: Trends of interaction energies (a, b, c) and the relationships between the Ees and Etot (a′, b′, c′), Figure S4, The relationships between the Ees and Etot (a) I⋅⋅⋅(S2)x, (b) II⋅⋅⋅(S1)x, (c) II⋅⋅⋅(S2)x and (d) III⋅⋅⋅(S2)x for the ECP1 of clusters, Figure S5: The relationship between Δd and Δλ for complexes X⋅⋅⋅(S)x in T1 states, Figure S6: The relationship between μS1ES1−T1 values in theory and the ΦPL values in experiment, Table S1: Comparing the calculated and experimental parameters for the different oligomeric I (bond length: Ǻ, wavelength: nm), Table S2: The NBO results of the clusters, Table S3: The key parameters and the phosphorescent character of complexes X⋅⋅⋅(S)x, Table S4: Selected GKS-EDA results of X⋅⋅⋅(Sn)x, Table S5: Molecular Orbitals and their main consist of natural atomic orbital and the characters, Table S6: Calculated phosphorescent emission (in nm) of the studied complexes with the TDDFT method, along with experimental values, Table S7: The singlet-triplet splitting energy (ΔES1−T1, in eV), transition electric dipole moment (μS1, in atomics units), the μS1ES1−T1, and the predicted ΦPL ( 50%) for complexes X⋅⋅⋅(S)x, Table S8: The ECP of cluster I at different calculated levels (kcal/mol), Table S9: The EDA results of cluster I at different calculated levels (kcal/mol).

Author Contributions

Z.-F.L. and X.-P.Y. conceived and designed the experiments; H.-X.L. performed the calculations; Z.-F.L. and Hui-Xue Li analyzed the data wrote the paper. Z.-F.L., H.-X.L. and G.-F.Z. revised the paper.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 21463023, 21465021) and the Natural Science Foundation of Gansu Province, China (Grant No. 17JR5RE010, 17YF1FE100).

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant No. 21463023, 21465021), the Natural Science Foundation of Gansu Province, China (Grant No. 17JR5RE010, 17YF1FE100). We thank the high-performance grid computing platform of Sun Yat-Sen University and Guangdong Province Key Laboratory of Computational Science for generous computer time. We would also like to thank Prof. PeiFeng Su for some instructive suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

References and Notes

  1. Pan, Q.-J.; Zhang, H.-X. An ab initio study on luminescent properties and aurophilic attraction of binuclear gold(I) complexes with phosphinothioether ligands. Inorg. Chem. 2004, 43, 593–601. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, Y.; Cheng, G.; Li, K.; Shelar, D.P.; Lu, W.; Che, C.-M. Phosphorescent polymeric nanomaterials with metallophilic d10···d10 interactions self-assembled from [Au(NHC)2]+ and [M(CN)2]. Chem. Sci. 2014, 5, 1348–1353. [Google Scholar] [CrossRef]
  3. López-de-Luzuriaga, J.M.; Monge, M.; Olmos, M.E.; Pascual, D.; Rodrı́guez-Castillo, M.A. Very short metallophilic interactions induced by three-center–two-electron perhalophenyl ligands in phosphorescent Au–Cu complexes. Organometallics 2012, 31, 3720–3729. [Google Scholar] [CrossRef]
  4. Chen, K.; Strasser, C.E.; Schmitt, J.C.; Shearer, J.; Catalano, V.J. Modulation of luminescence by subtle anion–cation and anion−π interactions in a trigonal AuI···CuI complex. Inorg. Chem. 2012, 51, 1207–1209. [Google Scholar] [CrossRef] [PubMed]
  5. Li, Z.-F.; Yang, X.-P.; Hui-Xue, L.; Guo, Z. Electronic structure of gold carbonyl compounds RAu-L (R = CF3, BO, Br, Cl, CH3, HCC, Mes3P, SIDipp; L = CO, N2, BO) and origins of aurophilic interactions in the clusters [RAuL]n (n = 2–4): A theoretical study. Organometallics 2014, 33, 5101–5110. [Google Scholar] [CrossRef]
  6. Tang, S.S.; Chang, C.-P.; Lin, I.J.B.; Liou, L.-S.; Wang, J.-C. Molecular aggregation of annular dinuclear gold(I) compounds containing bridging diphosphine and dithiolate ligands. Inorg. Chem. 1997, 36, 2294–2300. [Google Scholar] [CrossRef] [PubMed]
  7. Leung, K.H.; Phillips, D.L.; Tse, M.-C.; Che, C.-M.; Miskowski, V.M. Resonance raman investigation of the Au(I)−Au(I) interaction of the 1[dσ*pσ] excited state of Au2(dcpm)2(ClO4)2 (dcpm = bis(dicyclohexylphosphine)methane). J. Am. Chem. Soc. 1999, 121, 4799–4803. [Google Scholar] [CrossRef]
  8. Fernández, E.J.; Gimeno, M.C.; Laguna, A.; López-de-Luzuriaga, J.M.; Monge, M.; Pyykkö, P.; Sundholm, D. Luminescent characterization of solution oligomerization process mediated gold−gold interactions. Dft calculations on [Au2Ag2R4L2]n moieties. J. Am. Chem. Soc. 2000, 122, 7287–7293. [Google Scholar] [CrossRef]
  9. Rawashdeh-Omary, M.A.; Omary, M.A.; Patterson, H.H.; Fackler, J.P. Excited-state interactions for [Au(CN)2]n and [Ag(CN)2]n oligomers in solution. Formation of luminescent gold−gold bonded excimers and exciplexes. J. Am. Chem. Soc. 2001, 123, 11237–11247. [Google Scholar] [CrossRef] [PubMed]
  10. Adachi, C.; Baldo, M.A.; Thompson, M.E.; Forrest, S.R. Nearly 100% internal phosphorescence efficiency in an organic light-emitting device. J. Appl. Phys. 2001, 90, 5048–5051. [Google Scholar] [CrossRef]
  11. Baldo, M.A.; Lamansky, S.; Burrows, P.E.; Thompson, M.E.; Forrest, S.R. Very high-efficiency green organic light-emitting devices based on electrophosphorescence. Appl. Phys. Lett. 1999, 75, 4–6. [Google Scholar] [CrossRef]
  12. Tsuboyama, A.; Iwawaki, H.; Furugori, M.; Mukaide, T.; Kamatani, J.; Igawa, S.; Moriyama, T.; Miura, S.; Takiguchi, T.; Okada, S.; et al. Homoleptic cyclometalated iridium complexes with highly efficient red phosphorescence and application to organic light-emitting diode. J. Am. Chem. Soc. 2003, 125, 12971–12979. [Google Scholar] [CrossRef] [PubMed]
  13. Shen, J.Y.; Lee, C.Y.; Huang, T.-H.; Lin, J.T.; Tao, Y.-T.; Chien, C.-H.; Tsai, C. High Tg blue emitting materials for electroluminescent devices. J. Mater. Chem. 2005, 15, 2455–2463. [Google Scholar] [CrossRef]
  14. McDowell, J.J.; Gao, D.; Seferos, D.S.; Ozin, G. Synthesis of poly(spirosilabifluorene) copolymers and their improved stability in blue emitting polymer leds over non-spiro analogs. Polym. Chem. 2015, 6, 3781–3789. [Google Scholar] [CrossRef]
  15. Chung, Y.-H.; Sheng, L.; Xing, X.; Zheng, L.; Bian, M.; Chen, Z.; Xiao, L.; Gong, Q. A pure blue emitter (CIEy [approximate] 0.08) of chrysene derivative with high thermal stability for OLED. J. Mater. Chem. C 2015, 3, 1794–1798. [Google Scholar] [CrossRef]
  16. Kim, S.O.; Lee, K.H.; Kim, G.Y.; Seo, J.H.; Kim, Y.K.; Yoon, S.S. A highly efficient deep blue fluorescent OLED based on diphenylaminofluorenylstyrene-containing emitting materials. Synth. Met. 2010, 160, 1259–1265. [Google Scholar] [CrossRef]
  17. Yam, V.W.-W.; Au, V.K.-M.; Leung, S.Y.-L. Light-emitting self-assembled materials based on d8 and d10 transition metal complexes. Chem. Rev. 2015, 115, 7589–7728. [Google Scholar] [CrossRef] [PubMed]
  18. Sluch, I.M.; Miranda, A.J.; Elbjeirami, O.; Omary, M.A.; Slaughter, L.M. Interplay of metallophilic interactions, π–π stacking, and ligand substituent effects in the structures and luminescence properties of neutral PtII and PdII aryl isocyanide complexes. Inorg. Chem. 2012, 51, 10728–10746. [Google Scholar] [CrossRef] [PubMed]
  19. Hunter, C.A.; Sanders, J.K.M. The nature of π-π Interactions. J. Am. Chem. Soc. 1990, 112, 5525–5534. [Google Scholar] [CrossRef]
  20. Lindeman, S.V.; Rathore, R.; Kochi, J.K. Silver(I) complexation of (poly)aromatic ligands. Structural criteria for depth penetration into cis-stilbenoid cavities. Inorg. Chem. 2000, 39, 5707–5716. [Google Scholar] [CrossRef] [PubMed]
  21. Aguilo, E.; Gavara, R.; Baucells, C.; Guitart, M.; Lima, J.C.; Llorca, J.; Rodriguez, L. Tuning supramolecular aurophilic structures: The effect of counterion, positive charge and solvent. Dalton T. 2016, 45, 7328–7339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Penney, A.A.; Sizov, V.V.; Grachova, E.V.; Krupenya, D.V.; Gurzhiy, V.V.; Starova, G.L.; Tunik, S.P. Aurophilicity in action: Fine-tuning the gold(I)–gold(I) distance in the excited state to modulate the emission in a series of dinuclear homoleptic gold(I)–NHC complexes. Inorg. Chem. 2016, 55, 4720–4732. [Google Scholar] [CrossRef] [PubMed]
  23. Deák, A.; Jobbágy, C.; Marsi, G.; Molnár, M.; Szakács, Z.; Baranyai, P. Anion-, solvent-, temperature-, and mechano-responsive photoluminescence in gold(I) diphosphine-based dimers. Chem.-Eur. J. 2015, 21, 11495–11508. [Google Scholar] [CrossRef] [PubMed]
  24. Berenguer, J.R.; Lalinde, E.; Martín, A.; Moreno, M.T.; Ruiz, S.; Sánchez, S.; Shahsavari, H.R. Photophysical responses in Pt2Pb clusters driven by solvent interactions and structural changes in the PbII environment. Inorg. Chem. 2014, 53, 8770–8785. [Google Scholar] [CrossRef] [PubMed]
  25. Romanova, J.; Ranga Prabhath, M.R.; Sadik, Y.; Jarowski, P.D. Molecular design of organometallic materials: Effect of the metallophilic interactions, ligand, metal, and oxidation state. In Quantum Systems in Physics, Chemistry, and Biology: Advances in Concepts and Applications; Tadjer, A., Pavlov, R., Maruani, J., Brändas, E.J., Delgado-Barrio, G., Eds.; Springer International Publishing: Cham, Zuerich, Switzerland, 2017; pp. 139–158. [Google Scholar]
  26. Kruppa, S.V.; Bappler, F.; Klopper, W.; Walg, S.P.; Thiel, W.R.; Diller, R.; Riehn, C. Ultrafast excited-state relaxation of a binuclear Ag(I) phosphine complex in gas phase and solution. Phys. Chem. Chem. Phys. 2017, 19, 22785–22800. [Google Scholar] [CrossRef] [PubMed]
  27. Imoto, H.; Nishiyama, S.; Yumura, T.; Watase, S.; Matsukawa, K.; Naka, K. Control of aurophilic interaction: Conformations and electronic structures of one-dimensional supramolecular architectures. Dalton T. 2017, 46, 8077–8082. [Google Scholar] [CrossRef] [PubMed]
  28. Koshevoy, I.O.; Chang, Y.-C.; Karttunen, A.J.; Haukka, M.; Pakkanen, T.; Chou, P.-T. Modulation of metallophilic bonds: Solvent-induced isomerization and luminescence vapochromism of a polymorphic Au–Cu cluster. J. Am. Chem. Soc. 2012, 134, 6564–6567. [Google Scholar] [CrossRef] [PubMed]
  29. Mirzadeh, N.; Drumm, D.W.; Wagler, J.; Russo, S.P.; Bhargava, S. Different solvates of the dinuclear cyclometallated gold(i) complex [Au22-C6H4AsMe2)2]: A computational study insight into solvent-effected optical properties. Dalton T. 2013, 42, 12883–12890. [Google Scholar] [CrossRef] [PubMed]
  30. Laguna, A.; Lasanta, T.; López-de-Luzuriaga, J.M.; Monge, M.; Naumov, P.; Olmos, M.E. Combining aurophilic interactions and halogen bonding to control the luminescence from bimetallic gold−silver clusters. J. Am. Chem. Soc. 2010, 132, 456–457. [Google Scholar] [CrossRef] [PubMed]
  31. Donamaría, R.; Gimeno, M.C.; Lippolis, V.; López-de-Luzuriaga, J.M.; Monge, M.; Olmos, M.E. Tuning the luminescent properties of a Ag/Au tetranuclear complex featuring metallophilic interactions via solvent-dependent structural isomerization. Inorg. Chem. 2016, 55, 11299–11310. [Google Scholar] [CrossRef] [PubMed]
  32. Wakabayashi, R.; Maeba, J.; Nozaki, K.; Iwamura, M. Considerable enhancement of emission yields of [Au(CN)2] oligomers in aqueous solutions by coexisting cations. Inorg. Chem. 2016, 55, 7739–7746. [Google Scholar] [CrossRef] [PubMed]
  33. Pyykkö, P. Strong closed-shell interactions in inorganic chemistry. Chem. Rev. 1997, 97, 597–636. [Google Scholar] [CrossRef] [PubMed]
  34. Runeberg, N.; Schütz, M.; Werner, H.-J. The aurophilic attraction as interpreted by local correlation methods. J. Chem. Phys. 1999, 110, 7210–7215. [Google Scholar] [CrossRef]
  35. O’Grady, E.; Kaltsoyannis, N. Does metallophilicity increase or decrease down group 11? Computational investigations of [Cl-M-PH3]2 (M = Cu, Ag, Au, [111]). Phys. Chem. Chem. Phys. 2004, 6, 680–687. [Google Scholar] [CrossRef]
  36. Fernández, E.J.; Laguna, A.; López-de-Luzuriaga, J.M.; Monge, M.; Olmos, M.E.; Puelles, R.C. Au(I)···Ag(I) metallophilic interactions between anionic units:  Theoretical studies on a AuAg4 square pyramidal arrangement. J. Phys. Chem. B 2005, 109, 20652–20656. [Google Scholar] [CrossRef] [PubMed]
  37. Otero-de-la-Roza, A.; Mallory, J.D.; Johnson, E.R. Metallophilic interactions from dispersion-corrected density-functional theory. J. Chem. Phys. 2014, 140, 18A504. [Google Scholar] [CrossRef] [PubMed]
  38. Latouche, C.; Skouteris, D.; Palazzetti, F.; Barone, V. TD-DFT benchmark on inorganic Pt(II) and Ir(III) complexes. J. Chem. Theory Comput. 2015, 11, 3281–3289. [Google Scholar] [CrossRef] [PubMed]
  39. Tsipis, A.C. Dft challenge of intermetallic interactions: From metallophilicity and metallaromaticity to sextuple bonding. Coordin. Chem. Rev. 2017, 345, 229–262. [Google Scholar] [CrossRef]
  40. Paenurk, E.; Gershoni-Poranne, R.; Chen, P. Trends in metallophilic bonding in Pd–Zn and Pd–Cu complexes. Organometallics 2017, 36, 4854–4863. [Google Scholar] [CrossRef]
  41. Kelly, J.T.; McClellan, A.K.; Joe, L.V.; Wright, A.M.; Lloyd, L.T.; Tschumper, G.S.; Hammer, N.I. Competition between hydrophilic and argyrophilic interactions in surface enhanced raman spectroscopy. ChemPhysChem 2016, 17, 2782–2786. [Google Scholar] [CrossRef] [PubMed]
  42. Amar, A.; Meghezzi, H.; Boixel, J.; Le Bozec, H.; Guerchais, V.; Jacquemin, D.; Boucekkine, A. Aggregation effect on the luminescence properties of phenylbipyridine Pt(II) acetylide complexes. A theoretical prediction with experimental evidence. J. Phys. Chem. A 2014, 118, 6278–6286. [Google Scholar] [CrossRef] [PubMed]
  43. Prabhath, M.R.R.; Romanova, J.; Curry, R.J.; Silva, S.R.P.; Jarowski, P.D. The role of substituent effects in tuning metallophilic interactions and emission energy of bis-4-(2-pyridyl)-1,2,3-triazolatoplatinum(ii) complexes. Angew. Chem. Int. Ed. 2015, 54, 7949–7953. [Google Scholar] [CrossRef] [PubMed]
  44. Sivchik, V.V.; Grachova, E.V.; Melnikov, A.S.; Smirnov, S.N.; Ivanov, A.Y.; Hirva, P.; Tunik, S.P.; Koshevoy, I.O. Solid-state and solution metallophilic aggregation of a cationic [Pt(NCN)L]+ cyclometalated complex. Inorg. Chem. 2016, 55, 3351–3363. [Google Scholar] [CrossRef] [PubMed]
  45. Lee, C.; Yang, W.; Parr, R.G. Development of the colle-salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785. [Google Scholar] [CrossRef]
  46. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  47. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  48. Zhao, Y.; Truhlar, D. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  49. Chai, J.-D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [Google Scholar] [CrossRef] [PubMed]
  50. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104–154119. [Google Scholar] [CrossRef] [PubMed]
  51. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef] [PubMed]
  52. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  53. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1997, 78, 1396. [Google Scholar] [CrossRef]
  54. Møller, C.; Plesset, M.S. Note on an approximation treatment for many-electron systems. Phys. Rev. 1934, 46, 618–622. [Google Scholar] [CrossRef]
  55. Head-Gordon, M.; Pople, J.A.; Frisch, M.J. MP2 energy evaluation by direct methods. Chem. Phys. Lett. 1988, 153, 503–506. [Google Scholar] [CrossRef]
  56. Pyykkö, P.; Runeberg, N.; Mendizabal, F. Theory of the d10–d10 closed-shell attraction: 1. Dimers near equilibrium. Chem.-Eur. J. 1997, 3, 1451–1457. [Google Scholar] [CrossRef]
  57. Li, Z.F.; Fan, Y.Z.; DeYonker, N.J.; Zhang, X.T.; Su, C.Y.; Xu, H.Y.; Xu, X.Y.; Zhao, C.Y. Platinum(II)-catalyzed cyclization sequence of aryl alkynes via C(sp3)-H activation: A DFT study. J. Org. Chem. 2012, 77, 6076–6086. [Google Scholar] [CrossRef] [PubMed]
  58. Liu, R.-F.; Franzese, C.A.; Malek, R.; Żuchowski, P.S.; Ángyán, J.G.; Szczȩśniak, M.M.; Chałasiński, G. Aurophilic interactions from wave function, symmetry-adapted perturbation theory, and rangehybrid approaches. J. Chem. Theory Comput. 2011, 7, 2399–2407. [Google Scholar] [CrossRef] [PubMed]
  59. Granatier, J.; Lazar, P.; Otyepka, M.; Hobza, P. The nature of the binding of au, ag, and pd to benzene, coronene, and graphene: From benchmark CCSD(T) calculations to plane-wave DFT calculations. J. Chem. Theory Comput. 2011, 7, 3743–3755. [Google Scholar] [CrossRef] [PubMed]
  60. Riley, K.E.; Pitoňák, M.; Jurečka, P.; Hobza, P. Stabilization and structure calculations for noncovalent interactions in extended molecular systems based on wave function and density functional theories. Chem. Rev. 2010, 110, 5023–5063. [Google Scholar] [CrossRef] [PubMed]
  61. Vener, M.V.; Egorova, A.N.; Churakov, A.V.; Tsirelson, V.G. Intermolecular hydrogen bond energies in crystals evaluated using electron density properties: DFT computations with periodic boundary conditions. J. Comput. Chem. 2012, 33, 2303–2309. [Google Scholar] [CrossRef] [PubMed]
  62. Chen, Y.; Pan, X.; Yan, H.; Tan, N. Multiple hydrogen-bonding interactions between macrocyclic triurea and F, Cl, Br, I and NO3: A theoretical investigation. Phys. Chem. Chem. Phys. 2011, 13, 7384–7395. [Google Scholar] [CrossRef] [PubMed]
  63. Peterson, K.A.; Puzzarini, C. Systematically convergent basis sets for transition metals. II. Pseudopotential-based correlation consistent basis sets for the group 11 (Cu, Ag, Au) and 12 (Zn, Cd, Hg) elements. Theor. Chem. Acc. 2005, 114, 283–296. [Google Scholar] [CrossRef]
  64. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  65. McLean, A.D.; Chandler, G.S. Contracted gaussian basis sets for molecular calculations. I. Second row atoms, Z=11-18. J. Chem. Phys. 1980, 72, 5639–5648. [Google Scholar] [CrossRef]
  66. Boys, S.F.; Bernardi, F. The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  67. Su, P.; Jiang, Z.; Chen, Z.; Wu, W. Energy decomposition scheme based on the generalized kohn-sham scheme. J. Phys. Chem. A 2014, 118, 2531–2542. [Google Scholar] [CrossRef] [PubMed]
  68. Schmidt, M.W.; Baldridge, K.K.; Boatz, J.A.; Elbert, S.T.; Gordon, M.S.; Jensen, J.H.; Koseki, S.; Matsunaga, N.; Nguyen, K.A.; Su, S.; et al. General atomic and molecular electronic structure system. J. Comput. Chem. 1993, 14, 1347–1363. [Google Scholar] [CrossRef]
  69. Miyoshi, E.; Mori, H.; Hirayama, R.; Osanai, Y.; Noro, T.; Honda, H.; Klobukowski, M. Compact and efficient basis sets of s- and p-block elements for model core potential method. J. Chem. Phys. 2005, 122, 074104–074108. [Google Scholar] [CrossRef] [PubMed]
  70. Sakai, Y.; Miyoshi, E.; Tatewaki, H. Model core potentials for the lanthanides. J. Mol. Struc.-Theochem. 1998, 451, 143–150. [Google Scholar] [CrossRef]
  71. Miyoshi, E.; Sakai, Y.; Tanaka, K.; Masamura, M. Relativistic dsp-model core potentials for main group elements in the fourth, fifth and sixth row and their applications. J. Mol. Struc.-Theochem. 1998, 451, 73–79. [Google Scholar] [CrossRef]
  72. Sakai, Y.; Miyoshi, E.; Klobukowski, M.; Huzinaga, S. Model potentials for molecular calculations. I. The sd-MP set for transition metal atoms Sc through Hg. J. Comput. Chem. 1987, 8, 226–255. [Google Scholar] [CrossRef]
  73. Nakashima, H.; Mori, H.; Mon, M.S.; Miyoshi, E. Theoretical study on structures of gold, silver and copper clusters using relativistic model core potentials. In Proceedings of the 1st WSEAS International Conference on Computational Chemistry, Cairo, Egypt, 29 –31 December 2007; World Scientific and Engineering Academy and Society (WSEAS): Cairo, Egypt, 2007; pp. 11–13. [Google Scholar]
  74. Sakai, Y.; Miyoshi, E.; Klobukowski, M.; Huzinaga, S. Model potentials for main group elements Li through Rn. J. Chem. Phys. 1997, 106, 8084–8092. [Google Scholar] [CrossRef]
  75. Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  76. Frisch, M.J.; Trucks, G.W.; Schlegel, G.W.; Scuseria, G.W. Gaussian 09, Revision d.01; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  77. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  78. Glendening, E.D.; Badenhoop, J.K.; Reed, A.E.; Carpenter, J.E.; Bohmann, J.A.; Morales, C.M.; Weinhold, F. NBO 5.0; Theoretical Chemistry Institute, University of Wisconsin: Madison, WI, USA, 2001. [Google Scholar]
  79. Pérez Paz, A.; Espinosa Leal, L.A.; Azani, M.-R.; Guijarro, A.; Sanz Miguel, P.J.; Givaja, G.; Castillo, O.; Mas-Ballesté, R.; Zamora, F.; Rubio, A. Supramolecular assembly of diplatinum species through weak PtII⋅⋅⋅PtII intermolecular interactions: A combined experimental and computational study. Chem.-Eur. J. 2012, 18, 13787–13799. [Google Scholar] [CrossRef] [PubMed]
  80. Romanova, J.; Ranga Prabhath, M.R.; Jarowski, P.D. Relationship between metallophilic interactions and luminescent properties in Pt(II) complexes: TD-DFT guide for the molecular design of light-responsive materials. J. Phys. Chem. C 2016, 120, 2002–2012. [Google Scholar] [CrossRef]
  81. Martínez-Salvador, S.; Forniés, J.; Martín, A.; Menjón, B. [Au(CF3)(CO)]: A gold carbonyl compound stabilized by a trifluoromethyl group. Angew. Chem. Int. Ed. 2011, 50, 6571–6574. [Google Scholar] [CrossRef] [PubMed]
  82. Frontera, A.; Quinonero, D.; Costa, A.; Ballester, P.; Deya, P.M. MP2 study of cooperative effects between cation-π, anion-π and π-π interactions. New J. Chem. 2007, 31, 556–560. [Google Scholar] [CrossRef]
  83. Vijay, D.; Sastry, G.N. The cooperativity of cation–π and π–π interactions. Chem. Phys. Lett. 2010, 485, 235–242. [Google Scholar] [CrossRef]
  84. Li, Z.-F.; Li, H.-X.; Yang, X.-P. The mutual interactions based on amphipathic tetraoxacalix[2]arene[2]triazine: Recognition cases of anion and cation investigated by computational study. Phys. Chem. Chem. Phys. 2014, 16, 25876–25882. [Google Scholar] [CrossRef] [PubMed]
  85. Because whatever S1, S2 or S3, the corresponding energy components of Ecpn is comparable and therefore, the GKS-EDA analysis is selected and the results listed.
  86. Godefroid, G.; Yuqi, L.; Yanling, S.; Juanjuan, S.; Xiaochun, Q.; Xiaohong, S.; Zhijian, W. Enhancing the blue phosphorescence of iridium complexes with a dicyclometalated phosphite ligand via aza-substitution: A density functional theory investigation. J. Mater. Chem. C 2014, 2, 8364–8372. [Google Scholar] [CrossRef]
Figure 1. Schematic structure of metallophilic host clusters X (X = I, II, and III).
Figure 1. Schematic structure of metallophilic host clusters X (X = I, II, and III).
Nanomaterials 08 00685 g001
Figure 2. The orbital interaction of aurophilic interaction (outward and inward) in cluster I.
Figure 2. The orbital interaction of aurophilic interaction (outward and inward) in cluster I.
Nanomaterials 08 00685 g002
Figure 3. The model of complex formation between host clusters X (X = I, II, and III) and solvents (CH3OH, CH3CH2OH, and H2O).
Figure 3. The model of complex formation between host clusters X (X = I, II, and III) and solvents (CH3OH, CH3CH2OH, and H2O).
Nanomaterials 08 00685 g003
Figure 4. Top view (a) and side view (b) of the complex I⋅⋅⋅(CH3OH)2 (bond length: Å).
Figure 4. Top view (a) and side view (b) of the complex I⋅⋅⋅(CH3OH)2 (bond length: Å).
Nanomaterials 08 00685 g004
Table 1. Interaction energies (ECPadd and ECPtot) for clusters IIII at PBE0-D3/BS3// PBE0/BS1 level (kcal/mol).
Table 1. Interaction energies (ECPadd and ECPtot) for clusters IIII at PBE0-D3/BS3// PBE0/BS1 level (kcal/mol).
I II III
ECPaddECPtot ECPaddECPtot ECPaddECPtot
−29.0−102.5 −29.7−101.2 −29.7−103.6
Table 2. The GKS-EDA results of IIII at B3LYP-D3/BS5 level (kcal/mol).
Table 2. The GKS-EDA results of IIII at B3LYP-D3/BS5 level (kcal/mol).
ModeEesEexErepEpolEdispEcorrEtot
IEadd−23.3 −46.0 74.8 −15.6 −14.0 −5.6 −29.6
Etot−95.8 −99.5 159.8 −29.9 −26.2 −12.5 −104.0
IIEadd−22.5 −43.0 69.7 −15.1 −14.8 −5.1 −30.7
Etot−92.2 −94.0 150.7 −29.4 −27.0 −11.2 −103.0
III *Eadd−23.7 −36.7 60.3 −11.7 −16.9 −28.7
Etot−96.1 −78.9 127.8 −22.0 −31.6 −100.9
* Calculated with LMO-EDA (localized molecular orbital energy decomposition analysis) at MP2/BS6 level.
Table 3. Interaction energies (kcal/mol) of X⋅⋅⋅(S)x.
Table 3. Interaction energies (kcal/mol) of X⋅⋅⋅(S)x.
(S)xI II III
ECP1ECP2ECP3ECP4ECP ECP1ECP2ECP3ECP4ECP ECP1ECP2ECP3ECP4ECP
S1−16.6 −111.6 −15.8 −109.8 −17.1 −111.5
(S1)2−15.3−30.7 −124.8 −15.4−30.9 −123.9 −15.8−31.5 −124.7
(S1)3−12.5−25.7−42.0 −136.2 −13.7−30.0−44.0 −137.7 −12.9−26.4−43.1 −136.6
(S1)4−11.8−23.6−37.1−50.7−146.3 −13.0−26.0−40.6−55.2−150.5 −12.2−24.3−38.2−52.0−147.5
S2−16.7 −112.0 −16.4 −110.2 −17.3 −111.8
(S2)2−15.6−31.2 −125.4 −16.0−32.1 −125.0 −16.0−32.0 −125.2
(S2)3−12.8−26.2−42.8 −179.8 −14.1−28.3−45.2 −139.1 −13.3−30.2−43.8 −137.3
(S2)4−12.2−25.4−38.0−51.6−147.5 −13.4−26.7−41.6−56.5−152.0 −12.5−25.0−39.0−53.0−148.4
S3−15.2 −110.5 −14.5 −108.4 −15.8 −110.4
(S3)2−14.1−28.3 −122.5 −14.0−28.0 −121.4 −14.6−29.2 −122.3
(S3)3−10.6−23.5−38.8 −132.6 −10.9−27.7−41.4 −133.4 −10.9−24.1−39.8 −133.0
(S3)4−10.5−20.9−33.3−45.7−141.8 −11.9−23.9−37.7−51.5−147.5 −10.8−21.5−34.3−47.0−142.3
Table 4. Selected optimized parameters for IIII (bond length: Å, bond angle: °) a.
Table 4. Selected optimized parameters for IIII (bond length: Å, bond angle: °) a.
StateCplx bAu-C1Au-C2M-C3M-C4Au-C5Au-C6Au1-M2M2-Au3Au1M2Au3
I2.04 2.04 2.00 2.00 2.04 2.04 3.12 3.11 179.98
S0II2.04 2.04 2.00 2.00 2.04 2.04 3.15 3.15 180.00
III2.04 2.04 2.05 2.05 2.04 2.04 3.06 3.06 180.00
I2.03 2.03 1.99 1.99 2.03 2.03 2.77 2.77 179.99
T1II2.03 2.03 1.99 1.99 2.03 2.03 2.76 2.75 179.98
III2.03 2.03 2.02 2.02 2.03 2.03 2.76 2.76 179.99
I−0.01 −0.01 −0.01 −0.01 −0.01 −0.01 −0.35 −0.34 0.01
ΔII−0.01 −0.01 −0.01 −0.01 −0.01 −0.01 −0.39 −0.40 −0.02
III−0.01 −0.01 −0.03 −0.03 −0.01 −0.01 −0.30 −0.01 −0.30
a the atom numbering scheme is given in Figure 1. b Cplx = Complex. Δ = the different between bond length/bond angle in T1 and S0 state.
Table 5. Calculated phosphorescent emission (in nm) of the studied clusters IIII with TD-DFT method, along with experimental values.
Table 5. Calculated phosphorescent emission (in nm) of the studied clusters IIII with TD-DFT method, along with experimental values.
Complexλ/E(eV)ConfigurationCharacter Exp aδ
I457/2.72HOMO-LUMO (98%)LMCT4489
II483/2.57HOMO-LUMO (97%)LMCT46518
III417/2.98HOMO-LUMO (97%)LMCT446-29
a reference [11], δ = cal-exp.

Share and Cite

MDPI and ACS Style

Li, Z.-F.; Yang, X.-P.; Li, H.-X.; Zuo, G.-F. Phosphorescent Modulation of Metallophilic Clusters and Recognition of Solvents through a Flexible Host-Guest Assembly: A Theoretical Investigation. Nanomaterials 2018, 8, 685. https://doi.org/10.3390/nano8090685

AMA Style

Li Z-F, Yang X-P, Li H-X, Zuo G-F. Phosphorescent Modulation of Metallophilic Clusters and Recognition of Solvents through a Flexible Host-Guest Assembly: A Theoretical Investigation. Nanomaterials. 2018; 8(9):685. https://doi.org/10.3390/nano8090685

Chicago/Turabian Style

Li, Zhi-Feng, Xiao-Ping Yang, Hui-Xue Li, and Guo-Fang Zuo. 2018. "Phosphorescent Modulation of Metallophilic Clusters and Recognition of Solvents through a Flexible Host-Guest Assembly: A Theoretical Investigation" Nanomaterials 8, no. 9: 685. https://doi.org/10.3390/nano8090685

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop