Ragheb Khalil
Bouriche
ab,
Douniazed
Hannachi
*cd,
Amel
Messai
a,
Christophe
Morell
e,
Amor
Azizi
b and
Henry
Chermette
*e
aLaboratory of Engineering and Advanced Materials Science (ISMA), Abbes Laghrour University, khenchela, 40000, Algeria
bLaboratory of Chemistry, Molecular Engineering and Nanostructures, Chemistry Department, Faculty of Science, Ferhat Abbas-Sétif University, Sétif 19000, Algeria
cDépartement de Chimie, Faculté des Sciences de la Matière, Université de Batna-1, Algeria
dLaboratoire d'Électrochimie, d'Ingénierie Moléculaire et de Catalyse Redox (LEIMCR), Département d’Enseignement de Base en Technologie, Faculté de Technologie, Université Ferhat Abbas, Sétif-1, Algeria. E-mail: douniazed_hannachi@univ-setif.dz; h_douniazed@yahoo.fr
eInstitut des Sciences Analytiques, Université de Lyon, Université Claude Bernard Lyon 1, UMR CNRS 5280, 69622 Villeurbanne Cedex, France. E-mail: henry.chermette@univ-lyon1.fr
First published on 6th May 2025
This study presents a comprehensive quantum computational investigation into the electronic structure, topological characteristics, and optical properties of two distinct series of copper hydride nanoclusters [Cu3L3(H)i(BH4)j]+ and [Cu3L2(H)i(BH4)j]+, where i = 0 to 2 and j = 0 and 1. The electronic stability and reactivity are systematically evaluated through chemical hardness and chemical potential analyses, revealing that [Cu3L3(H)(BH4)]+ exhibits the highest electronic stability, whereas [Cu3L2(H)(BH4)]+ presents the greatest susceptibility to electronic perturbations. To gain deeper insights into the nature of the M–L interactions, EDA-NOCV is employed, demonstrating pronounced covalent character in the bonding interaction between the copper center and the ligand fragments. The linear and nonlinear optical parameters have been rigorously computed under both static and dynamic regimes. Notably, [Cu3L2(H)]+, [Cu3L2(H)(BH4)]+ and [Cu3L3(H)]+ exhibit remarkably high first hyperpolarizability (β0tot), reaching up to 1500 a.u., whereas Cu3L2(H)2 presents the lowest β0tot value (166 a.u.). These significant variations in NLO activity are attributed to the intrinsic topological framework of the nanoclusters, coupled with kinetic and potential energy at the AIM critical points BCPs and RCPs, which critically modulate charge delocalization and electron density redistribution. Additionally, the pronounced polarizability anisotropy observed in [Cu3L2(H)]+ and [Cu3L2(H)(BH4)]+ is identified as a key factor in increasing their NLO efficiency. Furthermore, second-order hyperpolarizability analysis indicates that Cu3L3 exhibits the highest value (173.48 × 104 a.u.), whereas Cu3L2(H)2 presents the lowest (5.688 × 104 a.u.). Frequency-dependent analyses at laser wavelengths of 1340 nm and 1064 nm reveal a pronounced enhancement of dynamic hyperpolarizability over its static counterpart. Moreover, TD-DFT calculations validate the exceptional near-infrared (NIR) transparency of these nanoclusters. Specifically, nanoclusters [Cu3L3(H)(BH4)]+, [Cu3L3(H)2]+, [Cu3L2(H)2]+ and [Cu3L2(H)]+ exhibit outstanding transparency across both the NIR and visible spectral regions, underscoring their potential for advanced applications in nonlinear optics and optoelectronic devices.
Copper hydrides have garnered significant interest due to their diverse applications in catalysis and materials science.19 Their history dates back to 1844, when Wurtz first isolated a pyrophoric red solid, CuH.20 Since then, a broad range of polynuclear copper(I) hydride complexes (LxCuyHn) have been identified as crucial intermediates in various chemical transformations, including selective reductions, cycloaddition reactions,21 and enantioselective functionalization processes such as hydroamination, hydroalkylation, and hydrobromination.22–26 In addition to their catalytic significance, copper hydrides have also been investigated for their potential in hydrogen storage and CO2 activation, among other applications.27–30
Theoretical and experimental investigations of hydride nanoclusters have consistently attracted considerable scientific interest.31–33 In 2016, Li et al. conducted a comprehensive study on the synthesis, structural elucidation, and gas-phase reactivity of bis(diphenylphosphino)amino copper hydride nanoclusters.34 These new complexes, represented as Cu3(X)(μ3-H)((PPh2)2NH)3, where X = μ2-Cl and μ3-BH4, were prepared and extensively characterized by high-resolution electrospray ionization mass spectrometry (ESI-MS), X-ray crystallography, nuclear magnetic resonance (NMR), and infrared (IR) spectroscopy. This study underscores the significant potential of copper hydride nanoclusters, which serve as a model system for the activation of BH4, facilitating the formation of coordinated hydrides. These hydrides subsequently undergo H2 elimination, promoting the generation of metallic clusters. Furthermore, the research provides crucial structural and mechanistic insights into polynuclear copper hydrides, laying the groundwork for an expanded understanding of their electronic properties and optical applications. Moreover, these findings open new avenues for the discovery of additional functionalities and potential applications of these nanoclusters in advanced materials science.
In this study, a comprehensive investigation is conducted utilizing quantum theory of atoms in molecules (QTAIM), energy decomposition analysis-natural orbitals for chemical valence (EDA-NOCV) and time-dependent density functional theory (TD-DFT) to study two distinct series of copper hydride nanoclusters designed and synthesized by Li et al.,34 namely [Cu3L3(H)i(BH4)j]+ and [Cu3L2(H)i(BH4)j]+, where i = 0 to 2 and j = 0 or 1 (ranging from A to H), as illustrated in Fig. 1. In this study, two specific cases were excluded from consideration. The first case, corresponding to i = 2 and j = 1, was eliminated due to severe steric congestion and pronounced geometric distortion, which rendered the incorporation of a second hydride ligand both energetically unfavourable and structurally impractical. The second case, i = 1 and j = 1, was excluded because the optimized geometry of the [Cu3L2(BH4)]+ nanocomplex exhibited complete structural distortion, thereby preventing its use in meaningful comparative analysis.
The primary objective of this study is to predict the optical properties of these nanoclusters and to systematically assess the impact of hydride units (H), BH4 ligands, and the bis(dimethylphosphino)amine (dmpa) ligand (L) on their nonlinear optical (NLO) responses and electronic structures. Notably, to achieve this DFT calculation, the bis(diphenylphosphino)amine (dppa) ligand, as identified in experimental results,34 was systematically substituted with bis(dimethylphosphino)amine (dmpa = L). This strategic modification enables a thorough investigation of the influence of L, H, and BH4 on the electronic structure and optical properties of the nanoclusters, providing deeper insight into their fundamental characteristics. To enable a precise comparison between the two nanocluster series, [Cu3L3]+ and [Cu3L2]+ (D and G, respectively) were designated as reference compounds.
On the other hand, Amsterdam density functional (ADF2023)44 software was used for the QTAIM study45–48 and EDA-NOCV analyses.49,50 This calculation used the PBE51 functional with the triple-ζ plus polarization (TZP) basis set, with numerical quality “VeryGood”. The integration grid was the Becke Grid with quality “VeryGood”.52 Relativistic corrections were taken into account with the use of the relativistic scalar zero-order-regular approximation (ZORA);53,54 no dispersion correction was used.
The average polarizability (α) and its anisotropy (Δα) are defined through the relations55
![]() | (1) |
![]() | (2) |
![]() | (3) |
The total magnitude of static second hyperpolarizability γ is estimated as56–58
![]() | (4) |
On the other hand, the quadratic nonlinear refractive index (n2) is calculated from the second hyperpolarizability coefficients, utilizing the following equation:59
n2 (cm2 W−1) = 8.28 × 10−23γDFWM (a.u.) | (5) |
![]() | (6) |
![]() | (7) |
![]() | (8) |
D index = [(Dx)2 + (Dy)2 + (Dz)2]1/2 | (9) |
Additionally, tindex is introduced to evaluate the separation degree between the electron and hole in the CT direction. A tindex value below zero indicates minimal separation between the electron and hole, whereas a distinctly positive value signifies a noticeable separation between them.
tindex = DCT − HCT | (10) |
HCT = |H·uCT| | (11) |
On the other hand, the variations of dipole moment of the excited state with respect to the ground state in X, Y and Z directions can be calculated as
![]() | (12) |
In our study, the crucial excited-state wavefunctions have been determined by natural transition orbitals (NTOs), which represent the electronic transition density matrix in terms of an “excited particle” and an “empty hole.” Additionally, charge transfer (CT) indices were computed using the MULTIWFN program.64
On the other hand, the effect of solvent environments on the nonlinear optical responses of the A nanocluster was investigated using the conductor-like polarizable continuum model (CPCM).65,66 Geometry optimizations of the A nanocluster were performed in solvent phases (water, methanol, and benzene) employing CPCM at the same level of theory as used for the gas-phase calculations.34
According to the literature, experimental and calculated values of nonlinear optical parameters exhibit a consistent trend. In other words, an increase in the experimental hyperpolarizability of some compound is reliably accompanied by a corresponding rise in theoretical predictions. This emphasizes that the computational method does not influence the relative ordering of properties within a series of molecules but significantly impacts the absolute magnitude of the computed values.36–40,67–69
Exp. | Gas | Water | Methanol | Benzene | |
---|---|---|---|---|---|
Exp: experimental data from ref. 34. X: centroid of the triangular arrangement of copper atoms. | |||||
X–hydride | 0.866 | 0.925 | 0.855 | 0.862 | 0.870 |
X–B | 2.072 | 2.218 | 2.219 | 2.215 | 2.196 |
X–Cu | 1.533 | 1.509 | 1.528 | 1.525 | 1.521 |
Cu–B | 2.577 | 2.681 | 2.694 | 2.689 | 2.671 |
Cu–hydride | 1.760 | 1.707 | 1.751 | 1.751 | 1.752 |
B–X–hydride | 178 | 180 | 180 | 179 | 180 |
Fig. 2 presents the chemical hardness and chemical potential of nanoclusters A to H, as well as their HOMO–LUMO orbitals. A comparison of chemical hardness (η) among nanoclusters A to H reveals significant variations in their electronic stability. For the [Cu3L3(H)i(BH4)j]+ series (k = 3), nanocluster A exhibits the highest η (7.074 eV), indicating the greatest stability and resistance to charge transfer, followed by B (6.219 eV), F (4.754 eV), and D (4.081 eV). This trend suggests that the presence of BH4 and hydride fragments enhances the stability of the nanoclusters compared to those with only two hydrides or one hydride. Conversely, for the Cu3L2(H)i(BH4)j series (k = 2), the trend reverses, with nanocluster E showing the highest η among H, G, and C, indicating that the two hydride fragments in E contribute more significantly to stability than the combination of one hydride or hydride with BH4 in C. This difference in trends between the two series can be attributed to the varying number of L ligands, which play a crucial role in modulating the electronic stability of the nanoclusters. In summary, the hardness trend from highest to lowest is A > B > E > H > F > G > C > D, with A being the most stable and D the most reactive, highlighting the inverse relationship between hardness and reactivity across the nanoclusters.
![]() | ||
Fig. 2 Shapes of the HOMO and LUMO, chemical potential (μ, eV) and chemical hardness (η, eV) of nanoclusters A to H. |
Regarding the chemical potential (μ), nanoclusters H, E, F, and A exhibit the most negative μ values (−6.724, −6.562, −6.087, and −6.016 eV, respectively), indicating a lower tendency to release electrons, and thus enhanced stability. These are followed by G (−5.862 eV), C (−5.826 eV) and B (−5.683 eV). In contrast, nanocluster D has the least negative μ (−5.196 eV), suggesting that it is the most nucleophilic and has a greater tendency to donate electrons. The overall order of chemical potential from most negative to least negative for both series is F ≈ A> B > D and H > E > G ≈ C, suggesting that nanoclusters without hydride fragments possess a greater electron-donating ability.
Within the QTAIM framework, critical points—such as bond critical points (BCPs) and ring critical points (RCPs)—are identified as positions where the gradient of the electron density vanishes. These critical points are characterized by several parameters, including the electron density (ρ), its Laplacian (∇2(ρ)), total energy density (H), kinetic energy density (G), and potential energy density (V). These descriptors provide insights into the nature of bonding and electronic interactions within the system. The computed parameters for nanoclusters A through H are systematically presented in tables and depicted in figures, providing a comprehensive visualization of the electron density topology. In these graphical representations, green circles indicate ring critical points (RCPs), while red circles correspond to bond critical points (BCPs).
The QTAIM analysis of nanocluster A provides detailed insights into its electronic structure and bonding interactions. From the results in Table 2 and Fig. 3 we can note that the RCPs associated with Cu–Cu–P–N–P (RCP1–RCP3), the low electron density (ρ ≈ 0.0143–0.0144 a.u.) and positive Laplacian (∇2ρ ≈ 0.0332–0.0333 a.u.) indicate weak, closed-shell interactions. Conversely, RCPs linked to Cu–Cu–H–B–H (RCP1′–RCP3′) exhibit higher ρ (≈ 0.0209–0.0210 a.u.) and (∇2(ρ)) (0.0410–0.0411 a.u.), suggesting stronger interactions with slight covalent contributions, as evidenced by negative total energy densities (H ≈ −0.0012 a.u.).
ρ | ∇2(ρ) | G | V | H | |
---|---|---|---|---|---|
RCP: Cu–Cu–P–N–P; RCP′: Cu–Cu–H–B–H; RCP′′: 3Cu-H-BH4. BCP: Cu⋯H; BCP′: Cu⋯H–B. Values are given in atomic units (a.u.). | |||||
RCP1 | 0.0144 | 0.0333 | 0.0080 | −0.0076 | 0.0003 |
RCP2 | 0.0143 | 0.0333 | 0.0080 | −0.0076 | 0.0003 |
RCP3 | 0.0143 | 0.0332 | 0.0080 | −0.0076 | 0.0003 |
RCP1′ | 0.0209 | 0.0410 | 0.0114 | −0.0126 | 0.0011 |
RCP2′ | 0.0210 | 0.0411 | 0.0114 | −0.0126 | −0.0012 |
RCP3′ | 0.0210 | 0.0410 | 0.0114 | −0.0126 | −0.0012 |
RCP′′ | 0.0201 | 0.0370 | 0.0104 | −0.0116 | −0.0012 |
BCP1 | 0.0681 | 0.1521 | 0.0579 | −0.0779 | −0.0199 |
BCP2 | 0.0678 | 0.1517 | 0.0576 | −0.0773 | −0.0197 |
BCP3 | 0.0679 | 0.1517 | 0.0577 | −0.0775 | −0.0198 |
BCP1′ | 0.1419 | −0.1877 | 0.0796 | −0.2061 | −0.1265 |
BCP2′ | 0.1420 | −0.1885 | 0.0796 | −0.2063 | −0.1267 |
BCP3′ | 0.1418 | −0.1874 | 0.0795 | −0.206 | −0.1264 |
The nature of hydrogen bonding interactions is classified according to the criteria established by Rozas et al.78 Based on the parameters at the BCPs, two distinct types of hydrogen bonds are identified. The Cu⋯H interactions (BCP1–BCP3) exhibit moderate electron density (ρ ≈ 0.0678–0.0681 a.u.) and positive Laplacian values (∇2(ρ) ≈ 0.152 a.u.), accompanied by negative total energy densities (H ≈ −0.0197 to −0.0198 a.u.). These characteristics align with moderate hydrogen bonds, which display mixed electrostatic and covalent character. In contrast, the Cu⋯H–B interactions (BCP1′–BCP3′) are characterized by significantly higher electron density (ρ ≈ 0.142 a.u.) and negative Laplacian values (∇2(ρ) ≈ −0.187 a.u.), along with highly negative total energy densities (H ≈ −0.126 a.u.). These parameters are indicative of strong hydrogen bonds with a covalent-like character.
Overall, the hydrogen bonding framework in nanocluster A showcases a dual nature, with moderate Cu⋯H bonds contributing to structural flexibility and strong Cu⋯H–B bonds providing stability through robust covalent interactions.
For nanocluster B, it can be observed that RCP1, RCP2 and RCP3 exhibit low electron density (ρ ≈ 0.014–0.015 a.u.) and positive Laplacian (∇2ρ ≈ 0.034–0.038 a.u.), indicating weak, closed-shell interactions (Table 3 and Fig. 4). The RCP′ reveals ρ ≈ 0.043 a.u. and ∇2(ρ) ≈ 0.086 a.u., indicating closed-shell interactions. The Cu⋯H interactions (BCP1–BCP3) display electron density values ranging from ρ = 0.0532 to 0.0773 a.u., with positive Laplacian values (∇2(ρ) = 0.1200 to 0.1480 a.u.) and negative total energy densities (H = −0.028 to −0.012 a.u.). These parameters indicate the presence of moderate hydrogen bonding with a hybrid electrostatic–covalent character. A similar character is observed in BCP′1 and BCP′2 (where ρ = 0.0706 and 0.0686 a.u., ∇2(ρ) = 0.1373 and 0.1367 a.u. and H = −0.0232 and −0.0216 a.u., respectively).
ρ | ∇2(ρ) | G | V | H | |
---|---|---|---|---|---|
RCP: Cu–Cu–P–N–P; RCP′: Cu–H–Cu–H. BCP: Cu⋯H (H between 3 Cu); BCP′: Cu⋯H (H between 2 Cu). Values are given in atomic units (a.u.). | |||||
RCP1 | 0.0147 | 0.0348 | 0.0083 | −0.0080 | 0.0003 |
RCP2 | 0.0153 | 0.0342 | 0.0084 | −0.0082 | 0.0001 |
RCP3 | 0.0148 | 0.0385 | 0.0090 | −0.0083 | 0.0006 |
RCP1′ | 0.0427 | 0.0860 | 0.0293 | −0.0371 | −0.0078 |
BCP1 | 0.0532 | 0.1200 | 0.0416 | −0.0533 | −0.0116 |
BCP2 | 0.0773 | 0.1480 | 0.0650 | −0.0929 | −0.0280 |
BCP3 | 0.0565 | 0.1246 | 0.0447 | −0.0582 | −0.0135 |
BCP1′ | 0.0706 | 0.1373 | 0.0575 | −0.0807 | −0.0232 |
BCP2′ | 0.0686 | 0.1367 | 0.0558 | −0.0774 | −0.0216 |
From Table 4, the RCPs associated with Cu–Cu–P–N–P (RCP1–RCP3) exhibit low ρ (0.0152–0.0318 a.u.) and positive ∇2(ρ) (0.0386–0.0612 a.u.), indicating weak closed-shell interactions. The Cu⋯H interactions (BCP1–BCP3) show ρ = 0.062 to 0.070 a.u., ∇2ρ = 0.142 to 0.152 a.u., and H = −0.02146 to −0.0165 a.u., suggesting moderate hydrogen bonding with a mixed electrostatic–covalent character. A similar trend is observed for BCP′1 and BCP′2, characterized by ρ = 0.065 and 0.066 a.u., ∇2ρ = 0.186 and 0.184 a.u., and H = −0.014 and −0.015 a.u., respectively (Fig. 5).
ρ | ∇2(ρ) | G | V | H | |
---|---|---|---|---|---|
RCP: Cu–Cu–P–N–P; RCP′: Cu–BH4–Cu. BCP: 3Cu⋯H; BCP′: Cu⋯H–B. Values are given in atomic units (a.u.). | |||||
RCP1 | 0.0173 | 0.0444 | 0.0107 | −0.0103 | 0.0004 |
RCP2 | 0.0318 | 0.0612 | 0.0194 | −0.0235 | −0.0041 |
RCP3 | 0.0152 | 0.0386 | 0.0091 | −0.0086 | 0.0005 |
BCP1 | 0.0626 | 0.1426 | 0.0521 | −0.0686 | −0.0165 |
BCP2 | 0.0691 | 0.1483 | 0.0581 | −0.0791 | −0.0210 |
BCP3 | 0.0700 | 0.1526 | 0.0596 | −0.0811 | −0.0215 |
BCP1′ | 0.0652 | 0.1860 | 0.0613 | −0.0762 | −0.0149 |
BCP2′ | 0.0660 | 0.1841 | 0.0616 | −0.0770 | −0.0155 |
For nanocluster D, Table 5 highlights distinct bonding characteristics in RCPs, where Cu–Cu–P–N–P (RCP1–RCP3) interactions exhibit ρ = 0.0152–0.0153 a.u. and Δ2ρ = 0.0371–0.0372 a.u., indicating weak closed-shell interactions. In contrast, RCP1′ presents a higher electron density (ρ = 0.0403 a.u.) and a lower Laplacian (Δ2ρ = 0.0304 a.u.), suggesting a stronger yet electrostatically influenced bond (see Fig. 6).
ρ | ∇2(ρ) | G | V | H | |
---|---|---|---|---|---|
RCP: Cu–Cu–P–N–P; RCP′: trinuclear copper. Values are given in atomic units (a.u.). | |||||
RCP1 | 0.0152 | 0.0371 | 0.0088 | −0.0085 | 0.0004 |
RCP2 | 0.0153 | 0.0372 | 0.0089 | −0.0085 | 0.0004 |
RCP3 | 0.0153 | 0.0371 | 0.0089 | −0.0085 | 0.0004 |
RCP1′ | 0.0403 | 0.0304 | 0.0187 | −0.0298 | −0.0111 |
For E, the data in Table 6 indicate that RCPs (Cu–Cu–P–N–P) exhibit low ρ (0.0131–0.0167 a.u.) and positive Δ2ρ, suggesting weak closed-shell interactions. In contrast, BCPs (Cu⋯H) show higher ρ (0.0663–0.0801 a.u.) and negative H, indicating moderate hydrogen bonding with a mixed electrostatic–covalent nature, where BCP′2 (Cu1⋯H2) is the strongest interaction (see Fig. 7).
ρ | ∇2(ρ) | G | V | H | |
---|---|---|---|---|---|
RCP: Cu–Cu–P–N–P. BCP1: Cu3⋯H1; BCP2: Cu4⋯H1; BCP′1: Cu3⋯H2; BCP′2: Cu1⋯H2. Values are given in atomic units (a.u.). | |||||
RCP1 | 0.0131 | 0.0325 | 0.0075 | −0.0069 | 0.0006 |
RCP2 | 0.0167 | 0.0432 | 0.0103 | −0.0098 | 0.0005 |
BCP1 | 0.0801 | 0.1550 | 0.0686 | −0.0984 | −0.0298 |
BCP2 | 0.0663 | 0.1361 | 0.0539 | −0.0737 | −0.0199 |
BCP′1 | 0.0764 | 0.1558 | 0.0655 | −0.0920 | −0.0265 |
BCP′2 | 0.0880 | 0.1530 | 0.0754 | −0.1126 | −0.0372 |
For nanocluster F, the data in Table 7 show that RCPs (Cu–Cu–P–N–P) have low ρ (0.0144–0.0169 a.u.) and positive ∇2ρ, indicating weak closed-shell interactions. In contrast, BCPs (3Cu⋯H) exhibit higher ρ (0.0663–0.0766 a.u.) and negative H, suggesting moderate hydrogen bonding with a mixed electrostatic–covalent nature, where BCP1 (3Cu⋯H) represents the strongest interaction in this nanocluster (see Fig. 8).
ρ | ∇2(ρ) | G | V | H | |
---|---|---|---|---|---|
RCP: Cu–Cu–P–N–P; BCP: 3Cu⋯H. Values are given in atomic units (a.u.). | |||||
RCP1 | 0.0155 | 0.0357 | 0.0087 | −0.0085 | 0.0002 |
RCP2 | 0.0144 | 0.0302 | 0.0075 | −0.0074 | 0.00007 |
RCP3 | 0.0169 | 0.0402 | 0.0100 | −0.0097 | 0.0001 |
BCP1 | 0.0766 | 0.1582 | 0.0660 | −0.9257 | −0.2650 |
BCP2 | 0.0663 | 0.1527 | 0.0567 | −0.0751 | −0.0185 |
BCP3 | 0.0668 | 0.1504 | 0.0567 | −0.0758 | −0.0191 |
In the case of nanocluster G, the data in Table 8 indicate that RCP1 and RCP′ exhibit low ρ (ρ < 0.1 a.u.) and positive Δ2ρ, suggesting weak closed-shell interactions. In contrast, RCP2 presents a negative H (−0.0108 a.u.) and positive Δ2ρ (0.041 a.u.), indicating a stronger interaction with partial covalent character.
ρ | ∇2(ρ) | G | V | H | |
---|---|---|---|---|---|
RCP: Cu–Cu–P–N–P; RCP′: trinuclear copper. Values are given in atomic units (a.u.). | |||||
RCP1 | 0.0151 | 0.0420 | 0.0096 | −0.0088 | 0.0008 |
RCP2 | 0.0415 | 0.0419 | 0.0213 | −0.0321 | −0.0108 |
RCP1′ | 0.0150 | 0.0418 | 0.0096 | −0.0087 | 0.0008 |
A comparison between nanoclusters G and D reveals that G exhibits stronger Cu–Cu–P–N–P interactions than D, as evidenced by RCP2 (ρ = 0.0415 a.u., H = −0.0108 a.u.), which indicates greater covalent character than the corresponding interaction in D (ρ = 0.015 a.u.). However, RCP1′ remains similar in both clusters, suggesting that the effect of the third ligand is negligible (Fig. 9).
For nanocluster H, the data in Table 8 highlight distinct bonding characteristics in RCPs and BCPs. The RCPs exhibit low ρ (0.016–0.023 a.u.) and positive ∇2ρ (0.0317–0.0423 a.u.), indicating weak closed-shell interactions with predominantly electrostatic character. In contrast, the BCPs (Cu⋯H) show higher ρ (0.087–0.096 a.u.) and negative H (−0.0350 to −0.0436 a.u.), suggesting moderate hydrogen bonding with a mixed electrostatic–covalent nature. Notably, BCP2 (ρ = 0.0963 a.u., H = −0.0436 a.u.) exhibits a stronger interaction than BCP1, implying greater charge delocalization in this bonding region (Fig. 10).
ρ | ∇2(ρ) | G | V | H | |
---|---|---|---|---|---|
RCP: Cu–Cu–P–N–P; BCP: Cu⋯H. Values are given in atomic units (a.u.). | |||||
RCP1 | 0.0167 | 0.0370 | 0.0093 | −0.0093 | 0.0000 |
RCP2 | 0.0238 | 0.0317 | 0.0109 | −0.0140 | −0.0030 |
RCP3 | 0.0176 | 0.0423 | 0.0106 | −0.0104 | 0.0001 |
BCP1 | 0.0873 | 0.1716 | 0.0780 | −0.1130 | −0.0350 |
BCP2 | 0.0963 | 0.1740 | 0.0871 | −0.1308 | −0.0436 |
The EDA-NOCV method decomposes the intrinsic interaction energy (ΔEint) between two molecular fragments into three fundamental components:
ΔEint = ΔEelstat + ΔEPauli + ΔEorb |
To investigate the bonding characteristics of the studied nanoclusters, the EDA-NOCV approach was applied, partitioning each system into two molecular fragments: fragment 1, consisting of the three copper atoms, and fragment 2, comprising the coordinating ligands (L), along with BH4 and hydride units. The computed EDA-NOCV results are presented in Table 10. For the two series, the computed EDA-NOCV results reveal that the intrinsic interaction energies follow the trend:
[Cu3L3(H)i(BH4)j]+: F (−7025 kcal mol−1) < B (−7084 kcal mol−1) ≈ D (−7089 kcal mol−1) < A (−7529 kcal mol−1) |
[Cu3L2(H)i(BH4)j]+: H (−4662 kcal mol−1) < E (−4715 kcal mol−1) ≈ G (−4727 kcal mol−1) < C (−5160 kcal mol−1). |
ΔEPauli | ΔEelstat | ΔEorb | ΔEint | |
---|---|---|---|---|
A | 23![]() |
−4769 (15%) | −26![]() |
−7529 |
B | 22![]() |
−4595 (16%) | −24![]() |
−7084 |
C | 16![]() |
−3223 (15%) | −18![]() |
−5160 |
D | 22![]() |
−4571 (15%) | −24![]() |
−7089 |
E | 15![]() |
−3067 (16%) | −16![]() |
−4715 |
F | 22![]() |
−4573 (16%) | −24![]() |
−7025 |
G | 14![]() |
−3051 (16%) | −16![]() |
−4727 |
H | 14![]() |
−3055 (16%) | −16![]() |
−4662 |
This order indicates that [Cu3L2(H)i(BH4)j]+ fragments exhibit stronger interactions than [Cu3L3(H)i(BH4)j]+ analogs. The strongest bonding was observed for the [Cu3Lk(H)(BH4)j]+ nanocluster, whereas the weakest bonding was noted for [Cu3Lk(H)]+, where k = 2 and 3.
Despite considerable Pauli repulsion, the high intrinsic interaction energies primarily arise from orbital contributions, with covalent interactions (ΔEorb) accounting for approximately 84% of the total attractive interactions. In contrast, electrostatic contributions (ΔEelstat) are relatively small but exhibit systematic variations, being more pronounced in A and B compared to D and F. A similar trend is observed in the [Cu3L2(H)i(BH4)j]+ series, where electrostatic contributions are the lowest in reference nanoclusters (D and G) and increase slightly with the addition of one or two hydride fragments. Notably, replacing a single hydride unit with a BH4 group results in an approximate 5% increase in electrostatic interactions. Overall, these findings underscore the dominant role of orbital interactions in stabilizing the nanoclusters, while electrostatic effects remain an important secondary factor influenced by ligand composition and substitution.
The results, as presented in Fig. 11, reveal that the main absorption regions of nanoclusters D, F, and G are confined within the ultraviolet-visible (UV-VIS) spectrum, while nanoclusters A, B, E, and H primarily absorb in the ultraviolet (UV) domain. Notably, the absorption spectrum of nanocluster C extends into the near-infrared (NIR) region, reaching up to 900 nm. From Table 11, we can note that the first electronic transitions for nanoclusters A, B, E, and H occur at 244 nm, 303 nm, 349 nm, and 348 nm, respectively. In contrast, nanoclusters C, D, F, and G exhibit red-shifted first electronic transitions, appearing at 596 nm, 644 nm, 454 nm, and 505 nm, respectively. Detailed spectral analysis reveals that all the studied nanoclusters exhibit transparency in the near-infrared (NIR) region. Notably, nanoclusters A, B, E, and H display exceptional transparency across both the NIR and visible spectral regions, a property ascribed to the absence of notable absorption features within the wavelength range of 400–700 nm.
Sn | Δλ0→n | E 0→n | f | S r | D CT | Δμ | H | t index | Transition | |
---|---|---|---|---|---|---|---|---|---|---|
A | S1 | 244 | 5.072 | 0.001 | 0.562 | 0.615 | 2.957 | 3.373 | −0.658 | H → L+3 (42%); H → L (20%) |
S8 | 214 | 5.776 | 0.159 | 0.612 | 0.440 | 2.105 | 3.394 | −1.398 | H → L+5 (32%); H → L+7 (22%) | |
S10 | 214 | 5.781 | 0.165 | 0.610 | 0.537 | 2.562 | 3.351 | −1.292 | H → L+6 (34%); H → L+8 (24%) | |
B | S1 | 303 | 4.086 | 0.032 | 0.518 | 1.628 | 7.859 | 2.764 | 0.033 | H → L (56%); H → L+1 (20%) |
S4 | 354 | 4.867 | 0.065 | 0.562 | 0.208 | 1.001 | 3.376 | −1.770 | H → L+2 (16%); H → L+ (32%) | |
S5 | 251 | 4.939 | 0.065 | 0.543 | 0.990 | 4.752 | 3.436 | −1.042 | H → L+1 (27%); H−2→ L (17%) | |
S19 | 219 | 5.656 | 0.173 | 0.582 | 0.311 | 1.493 | 3.808 | −1.555 | H → L+10 (12%); H → L+3 (8%) | |
S27 | 207 | 5.970 | 0.132 | 0.583 | 0.374 | 1.787 | 3.963 | −1.453 | H → L+3 (10%); H → L+4 (8%) | |
C | S1 | 596 | 2.079 | 0.095 | 0.730 | 0.967 | 4.626 | 2.749 | −0.666 | H → L (80%); H → L+3 (8%) |
S2 | 513 | 2.418 | 0.011 | 0.593 | 1.477 | 7.069 | 2.896 | −0.007 | H → L+1 (88%); H → L (10%) | |
S3 | 396 | 3.134 | 0.034 | 0.679 | 0.996 | 4.760 | 3.091 | −0.810 | H → L+3 (62%); H → L+2 (14%) | |
S4 | 375 | 3.307 | 0.046 | 0.657 | 1.178 | 5.632 | 3.124 | −0.644 | H → L+4 (74%); H → L+5 (8%) | |
S5 | 331 | 3.743 | 0.052 | 0.417 | 3.274 | 15.572 | 3.354 | 1.198 | H → L+2 (38%); H → L+6 (36%) | |
S10 | 276 | 4.500 | 0.147 | 0.527 | 1.715 | 8.146 | 3.947 | −0.550 | H → L+10 (44%); H → L+9 (28%) | |
D | S1 | 644 | 1.926 | 0.009 | 0.657 | 0.096 | 0.463 | 2.858 | −1.335 | H → L (98%) |
S2 | 388 | 3.198 | 0.061 | 0.560 | 0.739 | 3.548 | 3.120 | −1.071 | H → L+4 (88%) | |
S6 | 340 | 3.650 | 0.097 | 0.366 | 0.963 | 4.572 | 4.094 | −1.684 | H → L+2 (70%); H → L+6 (8%) | |
S9 | 320 | 3.878 | 0.138 | 0.532 | 0.500 | 2.383 | 3.733 | −1.934 | H → L+10 (28%); H → L+9 (20%) | |
S12 | 296 | 4.1851 | 0.1101 | 0.448 | 1.415 | 6.768 | 4.154 | −1.358 | H → L+6 (44%); H → L+9 (26%) | |
E | S1 | 349 | 3.55 | 0.012 | 0.519 | 0.686 | 3.288 | 2.579 | −0.887 | H → L (92%) |
S2 | 277 | 4.476 | 0.003 | 0.518 | 0.488 | 2.367 | 2.721 | −1.261 | H → L+1 (76%); H → L+2 (8%) | |
S6 | 236 | 5.247 | 0.148 | 0.624 | 1.081 | 5.169 | 2.923 | −0.683 | H → L+5 (28%); H → L+3 (20%) | |
S8 | 226 | 5.489 | 0.116 | 0.641 | 0.384 | 1.834 | 3.024 | −1.533 | H → L+4 (16%); H → L+5 (14%) | |
S14 | 211 | 5.888 | 0.069 | 0.453 | 0.898 | 4.285 | 2.738 | −0.630 | H → L+7 (54%) | |
F | S1 | 454 | 2.732 | 0.002 | 0.459 | 2.550 | 12.239 | 2.726 | 0.870 | H → L (91%) |
S2 | 347 | 3.313 | 0.002 | 0.469 | 1.636 | 7.867 | 2.872 | −0.147 | H → L+1 (46%); H → L+2 (42%) | |
S6 | 288 | 4.310 | 0.014 | 0.475 | 2.551 | 12.242 | 3.262 | 0.457 | H → L+4 (56%); H → L+5 (16%) | |
S7 | 281 | 4.419 | 0.02 | 0.563 | 1.643 | 7.883 | 3.293 | −0.484 | H−4 → L (22%); H−2→ L+1 (20%) | |
S10 | 260 | 4.776 | 0.035 | 0.456 | 3.271 | 15.640 | 3.709 | 0.800 | H → L+4 (12%); H → L+1 (10%) | |
S12 | 252 | 4.920 | 0.05 | 0.392 | 4.224 | 20.205 | 3.397 | 2.055 | H → L+7 (24%); H → L+3 (18%) | |
G | S1 | 505 | 2.454 | 0.017 | 0.629 | 0.794 | 3.811 | 2.701 | −0.702 | H → L (98%) |
S2 | 413 | 2.999 | 0.116 | 0.675 | 0.639 | 3.059 | 2.799 | −1.091 | H → L+1 (96%) | |
S4 | 308 | 4.030 | 0.086 | 0.573 | 0.156 | 0.745 | 3.122 | −1.602 | H → L+5 (82%) | |
S9 | 280 | 4.434 | 0.102 | 0.697 | 0.209 | 1.002 | 3.124 | −1.948 | H → L+2 (34%); H → L+9 (34%) | |
S11 | 262 | 4.723 | 0.362 | 0.648 | 0.672 | 3.224 | 3.480 | −1.352 | H → L+9 (38%); H → L+2 (20%) | |
H | S1 | 348 | 3.563 | 0.068 | 0.507 | 2.511 | 12.061 | 2.658 | 0.797 | H → L (92%) |
S2 | 339 | 3.658 | 0.008 | 0.471 | 2.195 | 10.536 | 2.602 | 0.575 | H → L (90%) | |
S3 | 311 | 3.988 | 0.024 | 0.401 | 2.890 | 13.881 | 2.550 | 1.339 | H → L+2 (90%) | |
S6 | 265 | 4.670 | 0.127 | 0.565 | 2.357 | 11.317 | 2.801 | 0.551 | H → L+3 (32%); H → L+5 (20%) | |
S9 | 251 | 4.941 | 0.105 | 0.533 | 2.003 | 9.620 | 2.746 | 0.307 | H → L+3 (54%); H → L+2 (10%) | |
S11 | 242 | 5.120 | 0.105 | 0.662 | 1.076 | 5.168 | 3.106 | −0.954 | H → L+3 (58%); H → L+4 (12%) |
To achieve a comprehensive understanding of the electronic transitions in the studied nanocluster, hole–electron descriptors (DCT, Δμ, Sr, HCT, and tindex) were computed, and the results are summarized in Table 11, and the charge density difference (CDD) between the excited state and the ground state is presented in Fig. S1 in the ESI.† The analysis indicates that the first singlet excitation (S1) of the investigated system is primarily governed by a HOMO → LUMO transition, contributing approximately 70%. For nanoclusters C, D, and G, the S1 state is mainly attributed to intramolecular charge transfer (ICT), largely confined to the copper atoms (Fig. 2). This is evidenced by the calculated overlap integrals (Sr) of 0.73, 0.657, and 0.629, along with DCT of 0.967, 0.096, and 0.794 Å. The corresponding negative tindex values suggest localized electronic excitation within these nanoclusters.
In contrast, nanoclusters B, F and H exhibit distinct charge transfer (CT) pathways, reflecting variations in their electronic excitation behavior. Specifically, in B, the S1 state is governed by hydride-to-ligand charge transfer (HLCT) and metal-to-ligand CT (MLCT). In F, the excitation corresponds to ligand-to-hydride CT (LHCT), while in H, it is primarily associated with ligand-to-metal CT (LMCT). These transitions exhibit a positive tindex, with overlap integrals (Sr) of 0.518, 0.459, and 0.507, and the corresponding DCT of 1.628, 2.550, and 2.511 Å, respectively. Furthermore, nanocluster F exhibits the largest dipole moment variation (Δμ = 12.239 Debye), surpassing H (12.061 Debye) and B (7.859 Debye). Meanwhile, nanocluster A demonstrates a mixed CT character, involving both LLCT and MLCT, with associated descriptors Sr = 0.562 and DCT = 0.615 Å.
In the case of reference nanoclusters D and G, the hole–electron distributions corresponding to the S6 and S2 electronic transitions, occurring at 340 nm and 413 nm, respectively, are predominantly localized on the copper centers and surrounding molecular fragments. The computed centroid separation between the hole and electron distributions remains relatively small, with DCT values of 0.963 and 0.639 Å, respectively. Furthermore, these transitions exhibit moderate dipole moment variations (Δμ = 4.572 and 3.059 Debye), indicative of localized electronic excitation (Fig. S1, ESI†). For nanoclusters A and C, the S8 and S10 excitations are mainly concentrated on the BH4 group, hydride unit, and copper atoms. The corresponding centroid separations are computed as 0.440 Å for A and 1.715 Å for C. Notably, nanocluster C exhibits a significantly larger dipole moment variation (Δμ = 8.146 Debye) compared to A (Δμ = 2.105 Debye), suggesting a stronger polarization effect in the excited state.
Regarding nanocluster F, the S2 and S6 excitations, located at 347 nm and 288 nm, respectively, exhibit considerable charge separation, as reflected by centroid distances of DCT = 1.636 and 2.551 Å, respectively. Additionally, these transitions are characterized by significant dipole moment variations (Δμ = 7.867, 12.242 Debye), suggesting enhanced charge redistribution upon excitation. The computed overlap integrals (Sr = 0.469 and 0.475), along with the hole delocalization indices (H = 2.872 and 3.262), further support the presence of extended charge transfer. A similar trend is observed for nanocluster H, where the S3, S6, and S9 transitions at 311, 265, and 251 nm, respectively, exhibit notable centroid separations (2 Å < DCT < 2.9 Å). The corresponding oscillator strengths (f = 0.024, 0.127, and 0.105) indicate varying transition intensities. Additionally, these excitations display substantial dipole moment variations (Δμ = 13.881, 11.31 and 9.620 Debye), signifying pronounced charge redistribution. The hole–electron densities in these transitions are predominantly localized on the copper centers and ligand framework, reinforcing their involvement in the electronic excitation process.
α (a.u.) | Δα (a.u.) | μ (Debye) | |
---|---|---|---|
A | 379 | 91 | 1.954 |
B | 374 | 90 | 1.894 |
C | 332 | 87 | 6.130 |
D | 403 | 148 | 0.965 |
E | 264 | 72 | 4.694 |
F | 368 | 92 | 11.511 |
G | 285 | 104 | 3.563 |
H | 266 | 62 | 9.789 |
From our results we can note that nanoclusters A and B exhibit strikingly similar values for dipole moment, polarizability, and polarizability anisotropy (1.924 D, 376 a.u., and 90 a.u., respectively). This observation suggests that the substitution of the BH4 group with a hydride ligand induces negligible changes in these electronic parameters, reflecting minimal structural or electronic perturbations associated with this modification. Likewise, nanocluster F demonstrates comparable polarizability and anisotropy values (368 and 92 a.u., respectively). However, the removal of either the hydride or the BH4 group from A and B results in a substantial increase in the dipole moment of F (11.511 D), highlighting a dramatic redistribution of electronic density upon this structural change. For nanocluster D, the elimination of both the hydride and BH4 ligands yields the highest polarizability (α = 403 a.u.) and polarizability anisotropy (Δα = 148 a.u.), indicative of a highly flexible electron cloud and substantial structural anisotropy. Despite these characteristics, the low dipole moment (μ = 0.965 D) suggests a nearly symmetric charge distribution, highlighting a minimal degree of charge separation within the cluster.
On the other hand, compounds E and H have the lowest polarizability (α = 264 and 266, respectively), with correspondingly low anisotropic polarizability (Δα = 72 and 62 a.u., respectively) but substantial dipole moments (μ = 4.694 and 9.789 Debye, respectively), indicating compact structures with localized polarization. Moderate values for α and Δα, such as those of compound C (α = 332 a.u., Δα = 87 a.u., μ = 6.130 Debye), reflect balanced electronic distributions with considerable permanent dipole moments due to specific polar bonds or functional groups. A comparison of nanocluster G with nanoclusters C, E, and H reveals that the removal of hydride and BH4 ligands results in a comparable polarizability (285 a.u.) and the highest polarizability anisotropy (104 a.u.) observed among the studied [Cu3L2(H)i(BH4)j]+ systems. This outcome underscores the significant influence of ligand removal on enhancing anisotropic electronic properties while maintaining similar overall polarizability.
The results presented in Table 12 and Fig. 12 indicate that the polarizability of the nanocluster [Cu3L3(H)i(BH4)j]+ is higher than that of its counterpart, [Cu3L2(H)i(BH4)j]+ (i = 0, 1 and 2 and j = 0 or 1). Among the studied compounds, D, F, A and B exhibit the highest polarizability values. These findings suggest that the presence of three ligands (L) compared to two ligands in a structure leads to increased polarizability. Additionally, the calculated anisotropic polarizability (Δα), ranging from 62 to 148 a.u., appears to be moderately influenced by the L, BH4 ligand and hydride substituents. Moreover, the effect of the incident wavelength (λ = 1340 and 1064 nm) on polarizability and anisotropic polarizability is determined to be negligible, indicating minimal wavelength dependency in these optical properties.
![]() | ||
Fig. 12 Variation of polarizability and anisotropic polarizability in the static and dynamic regime of nanoclusters A to H. |
![]() | ||
Fig. 13 Variation in static first hyperpolarizability (β0tot) and hyper-Rayleigh scattering hyperpolarizability (β0HRS) for nanoclusters A to H. |
The ligand environment plays a critical role in determining the first hyperpolarizability (β0tot) and HRS (β0HRS) of nanoclusters, as it directly influences the electronic distribution, symmetry and polarizability and its anisotropy of the system. In the studied nanoclusters (A to H), the ligand type, number, and arrangement variation significantly modulate their nonlinear optical properties. The static first hyperpolarizability of the title nanocluster exhibits the following overall ordering: β0tot: E < A < B < D < G < F < C < H (β0HRS: E < A < B < G < F < D < C < H). From a more detailed analysis, the second-order NLO response of the Cu3L3Hi(BH4)j series was found to display the following ordering: β0tot: A < B < D < F and β0HRS: A < B < F < D. The first hyperpolarizability is relatively small in nanocluster A but increases notably upon substitution of the BH4 group with a hydride ligand in nanocluster B or the removal of both BH4 and hydride ligands in nanocluster D. Notably, nanocluster F, containing a single hydride ligand, exhibits the highest first hyperpolarizability among the systems studied. For instance, the β0tot value of nanocluster F is approximately 5 times greater than that of nanocluster D (Fig. 13).
Within the [Cu3L2Hi(BH4)j]+ series, nanocluster H, characterized by the presence of a single hydride ligand, exhibits the highest second-order NLO response. This response is diminished in nanocluster C with the incorporation of a BH4 ligand and further reduced in nanocluster G upon the removal of the hydride group. Nanocluster E, which contains two hydride ligands, demonstrates the lowest second-order NLO response in the series. For example, the β0tot [β0HRS] value of H is 2166 a.u. [843 a.u.], which is 13 [7] times larger than that of E. The ordering of the β0tot and β0HRS value for the nanoclusters Cu3L2Hi(BH4)j is E < G < C < H (Fig. 13). These results underscore the pivotal role of ligand composition and symmetry in determining the NLO performance of [Cu3LkHi(BH4)j]+ hydride nanoclusters.
For example, nanocluster G exhibits a higher static first hyperpolarizability (β0tot = 671 a.u.) compared to D (β0tot = 311 a.u.). This can be attributed to G's larger dipole moment (3.563 D), which enhances its charge redistribution under an applied field. However, D demonstrates a greater β0HRS value (736 a.u.) compared to G (407 a.u.), despite its lower β0tot. This discrepancy arises due to D having a higher polarizability anisotropy (148 a.u.) compared to G (104 a.u.). The larger anisotropy in D reflects a more directional electronic response, amplifying the hyperpolarizability contribution in processes associated with second-order scattering, such as hyper-Rayleigh scattering. Therefore, while G benefits from stronger overall charge redistribution (β0tot), D's enhanced anisotropic polarization drives its superior β∞HRS. A similar trend is observed for nanoclusters D and F, where F exhibits a significantly higher static first hyperpolarizability (β0tot = 1594 a.u.) compared to D (311 a.u.), primarily due to its substantial dipole moment (μ(F) = 11.511 D). Conversely, nanocluster D displays a β0HRS value that is 17% larger than that of F, attributed to its greater polarizability anisotropy.
On the other hand, nanocluster C exhibits a higher first hyperpolarizability than nanocluster A, a distinction that can be attributed to the characteristics of the Cu⋯H–B bonds (BCP′). According to QTAIM analysis, the BCP′ in nanocluster A is stronger and associated with a higher potential energy (V) compared to the corresponding bond in nanocluster C. This indicates that the electrons within the Cu⋯H–B bonds in nanocluster A are more tightly localized. Conversely, in nanocluster C, the Cu⋯H–B bonds exhibit greater electron delocalization and mobility (Tables 2 and 4). The increased electron delocalization observed in nanocluster C contributes significantly to its superior first hyperpolarizability compared to nanocluster A.
In addition, the first hyperpolarizability (β0HRS and β0t) of nanocluster B is marginally higher than that of nanocluster E, with a difference of approximately 20%. This variation can be attributed to the potential energy of the Cu⋯H bond (BCPs) in the two systems. In nanocluster B, the potential energy ranges from −0.09 to −0.05 a.u., whereas in nanocluster E, it spans from −0.11 to −0.07 a.u. (Tables 3 and 6). The more negative potential energy values observed in nanocluster E suggest greater electron localization compared to B, where the potential energy is less negative. This increased electron localization in E reduces its capacity for charge redistribution under an applied electric field, thereby slightly lowering its first hyperpolarizability relative to nanocluster B.
On the other hand, the higher first hyperpolarizability of nanocluster H compared to F can be attributed to the kinetic energy within the Cu–H bonds, which is greater in H (Tables 7 and 9). This indicates that electrons in Cu–H bonds are less localized in H than in F. Notably, the presence of a single hydride atom linking two copper atoms has a more pronounced impact on the electronic properties and first hyperpolarizability than a hydride coordinated to three copper atoms (Fig. 8 and 10).
Based on this analysis, it can be concluded that the structural topology of the title nanocluster, along with the spatial distribution of ligands around the copper atoms, exerts a more significant influence on the modulation and prediction of first hyperpolarizability than the number or chemical nature of the ligands within the system.
ω = 0.0340 a.u. (λ = 1340 nm) | ω = 0.04282 a.u. (λ = 1064 nm) | |||||
---|---|---|---|---|---|---|
β ω EOPE | β ω SHG | β ω HRS | β ω EOPE | β ω SHG | β ω HRS | |
A | 189 | 207 | 234 | 194 | 225 | 243 |
B | 287 | 339 | 259 | 301 | 400 | 276 |
C | 1401 | 4199 | 2083 | 1322 | 7776 | 3492 |
D | 414 | 1631 | 1332 | 478 | 381 | 1202 |
E | 175 | 204 | 138 | 183 | 239 | 148 |
F | 1708 | 1952 | 760 | 1773 | 2253 | 868 |
G | 1032 | 1492 | 733 | 1132 | 2627 | 1225 |
H | 2370 | 2884 | 1121 | 2502 | 3543 | 1377 |
The calculated values for the electro-optical Pockels effect (β(−ω; ω, 0)) for the highest-performing NLO nanoclusters H, F, C, and G at 1340 nm are 2370, 1708, 1401, and 1032 a.u., respectively. In contrast, nanoclusters E, A, B, and D exhibit significantly lower EOPE values at 1340 nm, with β(−ω; ω, 0) values of 175, 189, 287, and 414 a.u., respectively. Interestingly, nanoclusters D, E, G, and H demonstrate higher β(−ω; ω, 0) values at 1340 nm compared to their corresponding β0tot. At 1064 nm, the β(−ω; ω, 0) values for the nanoclusters are generally higher than those observed at 1340 nm, except for nanocluster C, which shows a lower value of 1322 a.u. These results indicate that the electro-optical Pockels effect is more sensitive to shorter wavelengths.
The second harmonic generation (SHG) hyperpolarizability values, β(−2ω; ω, ω), for the highest-performing NLO nanoclusters D, C, and G at 1340 nm are 1631, 4199, and 1492 a.u., respectively. Nanoclusters C, G, and H exhibit significantly greater sensitivity at 1064 nm, with corresponding SHG values of 7776, 2627, and 3543 a.u., respectively, demonstrating a marked increase compared to their static β0tot values. At 1064 nm, the hyper-Rayleigh scattering first hyperpolarizability values (βλHRS) for nanoclusters C, H, G, and F are 7776, 3543, 2627, and 2253 a.u., respectively. Furthermore, the βλHRS values at 1340 nm are consistently lower than their corresponding values at 1064 nm.
It is evident that both βλHRS and βλSHG responses are generally more pronounced at 1064 nm compared to 1340 nm, with the exception of nanocluster D, which deviates from this trend. Furthermore, the first hyperpolarizability in the dynamic regime is significantly larger than in the static regime. For instance, the βλHRS and βλSHG responses for nanocluster C at 1064 nm are approximately four times greater than those in the static regime. Similarly, for nanocluster G, the βλHRS and βλSHG responses at 1340 nm and 1064 nm are approximately two and three times, respectively, larger than their static values. These results emphasize that βλHRS and βλSHG responses are more sensitive to shorter wavelengths, displaying enhanced values at smaller wavelengths. However, the frequency-dependent effect on the electro-optical Pockels effect (EOPE) is comparatively weaker than on SHG, with the notable exception of nanocluster D at 1064 nm, where the EOPE response is relatively pronounced.
To further investigate the factors influencing the first hyperpolarizability of the studied nanoclusters, polarization scans of the hyper-Rayleigh scattering (HRS) intensity (I2ωΨV) were conducted. The relationship between I2ωΨV and the polarization angle ψ was calculated and plotted (Fig. 14), along with the corresponding depolarization ratio (DR) values. The results reveal that the DR values are sensitive to ligand composition and follow the descending order: H > F > G > C > E ≈ B > A ≈ D. Notably, nanoclusters F and H exhibit predominant dipolar symmetry in their NLO responses, with DR values of 5.971 and 6.366, respectively. In contrast, nanoclusters B, E, G, and C are classified as octupolar systems, with DR values of 1.943, 2.060, 2.434, and 3.463, respectively. For nanoclusters A and D, the DR values are 1.736 and 1.546, respectively, as shown in Fig. 3. These values, being close to 1.5, are indicative of typical octupolar symmetry.
![]() | ||
Fig. 14 Relationship between I2wΨV and polarization angle Ψ of nanoclusters A to H (black: static; red: λ = 1340 nm; blue: λ = 1064 nm). |
On the other hand, the results reveal that the depolarization ratio (DR) exhibits slight sensitivity to the wavelength of the incident light. Specifically, DR values follow the descending order: DR (λ = 1064 nm) > DR (λ = 1340 nm) > DR (λ = ∞). Additionally, from the data presented in Fig. 3, nanocluster D displays a notably low DR value (DR = 0.999, less than the threshold of 1.5). This low DR suggests that nanocluster D is near resonance at λ = 1340 nm, indicating enhanced nonlinear optical interactions under this specific wavelength.
To gain deeper insight into the variations in first hyperpolarizability values and their primary determining factors, we employed the conventional two-level model approach proposed by Oudar et al.88,89 The mathematical formulation of this model is expressed as follows:
It is well established that molecular nonlinear optical properties calculated via the sum-over-states (SOS) method strongly depend on the number of excited states considered. In this study, 90 excited states were calculated at the M06-2X/6-31+G(d)/SDD level of theory. The results presented in the ESI† (Fig. S2a and b) show that, in general, 90 excited states are sufficient for converging the βSOS value. On the other hand, the two-level analysis discloses that the excited state S8 predominantly influences the hyperpolarizability of nanocluster A. Furthermore, critical excited states S1, S6, and S2 exhibit the highest hyperpolarizability in nanoclusters C, E, and G, respectively. These electronic transitions are attributable to ICT processes, as depicted in Fig. 15. In the case of hydride nanoclusters B, D, F, and H, the principal contributions to their hyperpolarizability values are attributed to the S5, S6, S7, and S6 excited states, respectively. For both series [Cu3L3(H)i(BH4)j]+ and [Cu3L2(H)i(BH4)j]+, our results reveal a direct correlation between the dipole moment variation (Δμ) of crucial excited states and the first hyperpolarizability of the title nanoclusters, except for E. Notably, nanoclusters H and F exhibit the highest dipole moment variations of 11.317 and 7.883 Debye, respectively. These excitations are characterized by charge transfer from the hydride, copper, and ligand to the second ligand, with a Dindex of approximately 2.4 Å.
Furthermore, our results indicate that a single hydride unit enhances the dipole moment variation (Δμ) of crucial excited states more effectively than two hydride units or a hydride and BH4 fragment. These results are consistent with our previous study, which highlighted the significant role of a single hydride in modulating the electronic topology and second-order NLO responses.
ω | ω = 0.034 a.u. | ω = 0.04282 a.u. | ω = 0.034 a.u. | ω = 0.04282 a.u. | |||
---|---|---|---|---|---|---|---|
γ(0; 0, 0, 0) × 104 a.u. | γ(−ω; ω, 0, 0) × 104 a.u. | γ(−2ω; ω, ω, 0) × 104 a.u. | γ(−ω; ω, 0, 0) × 104 a.u. | γ(−2ω; ω, ω, 0) × 104 a.u. | n 2 × 10−18 cm2 W−1 | n 2 × 10−18 cm2 W−1 | |
A | 7.394 | 7.635 | 8.349 | 7.833 | 9.073 | 6.749 | 6.949 |
B | 8.555 | 8.980 | 10.066 | 9.276 | 18.970 | 8.098 | 10.555 |
C | 173.48 | 19.911 | 43.376 | 22.29 | 39.914 | −17.453 | −18.408 |
D | 17.04 | 19.280 | 28.979 | 20.89 | 34.31 | 20.592 | 22.063 |
E | 5.688 | 5.978 | 6.645 | 6.16 | 7.361 | 5.365 | 5.562 |
F | 9.229 | 9.769 | 11.188 | 10.144 | 12.975 | 8.940 | 9.433 |
G | 10.182 | 11.668 | 15.23 | 12.50 | 23.56 | 11.743 | 14.042 |
H | 7.469 | 8.12 | 9.93 | 8.58 | 12.44 | 7.783 | 8.476 |
Based on our results, it is observed that the second hyperpolarizability γ(0; 0, 0, 0) does not follow the same trend as the static first hyperpolarizability. Nanocluster C exhibits an exceptionally high static second hyperpolarizability (γ(0; 0, 0, 0) = 173.48 × 104 a.u.) compared to the other nanoclusters. Moderate γ(0; 0, 0, 0) values are observed for nanoclusters D (17.04 × 104 a.u.), G (10.182 × 104 a.u.) and F (9.22 × 104 a.u.). Nanocluster H displays a static second hyperpolarizability of 7.469 × 104 a.u., which is comparable to that of nanocluster A (7.394 × 10 4 a.u.). The lowest static hyperpolarizability value is found for nanocluster E (5.688 × 104 a.u.).
However, in the dynamic regime, nanocluster G stands out as the most efficient across both frequencies, particularly at ω = 0.04282 a.u., where it achieves the highest values of γ(−ω; ω, 0, 0) = 12.50 × 104 a.u. and γ(−2ω; ω, ω, 0) = 23.56 × 104 a.u. These values indicate nanocluster G's superior ability to mediate dynamic NLO processes. On the other hand, nanocluster C, despite its exceptional static performance, exhibits a significant drop in the dynamic second hyperpolarizability, with γ(−ω; ω, 0, 0) = 19.911 × 104 a.u. and γ(−2ω; ω, ω, 0) = 39.914 × 104 a.u. at ω = 0.04282 a.u., trailing behind clusters like G and B. Nanoclusters E and A showed the smallest dc-Kerr effect at ω = 0.034 a.u. as well as at ω = 0.04282 a.u., respectively. These results suggest that the third-order NLO response can be effectively enhanced by increasing the frequency of the incident light.
The degenerate four-wave mixing (DFWM) coefficients, γDFWM (γ(−ω; ω, −ω, ω)), were calculated using the static, DC-Kerr, and ESHG second hyperpolarizability coefficients, and the results are presented in Fig. 16. The data reveal distinct NLO responses among the nanoclusters, with most exhibiting positive DFWM values, indicative of self-focusing behavior, except for nanocluster C, which displays significant negative values (−21.08 × 104 and −22.23 × 104 a.u.), corresponding to strong self-defocusing effects.90 The positive DFWM values of nanoclusters D and G are the largest, reaching 26.65 × 104 a.u. and 16.96 × 104 a.u. at ω = 0.042 a.u., respectively, highlighting their potential for NLO applications such as optical switching and signal amplification. Nanoclusters A, E, F, and H show moderate positive values, while B exhibits an intermediate enhancement with increasing frequency. The DFWM values generally increase at higher frequencies, indicating frequency-dependent enhancement of the NLO response, except for nanocluster C, where negative values become slightly more pronounced, reinforcing its self-defocusing nature. The results underscore the variability in nonlinear optical behavior across the nanoclusters, with nanoclusters D and G being strong candidates for dynamic applications, while cluster C offers potential for niche uses requiring beam shaping or dispersion control.
The DFWM values and the quadratic nonlinear refractive index (n2) of the nanoclusters exhibit the same trend (Table 14). At both frequencies, the largest n2 values are observed for nanocluster D (20.592 × 10−18 cm2 W−1 and 22.063 × 10−18 cm2 W−1) followed by G (11.743 × 10−18 and 14.042 × 10−18 cm2 W−1). For nanoclusters A, E, and F, the n2 values at ω = 0.034 a.u. are 6.749 × 10−18, 5.365 × 10−18 and 8.940 × 10−18 cm2 W−1, respectively. These values show a slight increase at the higher frequency (ω = 0.04282 a.u.), demonstrating frequency dependence. Notably, the negative n2 values for nanocluster C at both frequencies suggest a dominant one-photon resonance effect, possibly overshadowing two-photon resonance contributions, distinguishing it from the other clusters.91,92
Water (ε = 78.355) | Methanol (ε = 32.613) | Benzene (ε = 2.270) | |
---|---|---|---|
α (0,0) | 807 | 802 | 605 |
Δα (0,0) | 109 | 99 | 53 |
β 0tot (0;0,0) | 6857 | 6221 | 2130 |
β 0HRS (0;0,0) | 3609 | 3676 | 939 |
γ(0; 0, 0, 0) | 2.678 × 106 | 1.183 × 106 | 1.408 × 106 |
The computed results reveal that the static polarizability (α) of nanocluster A undergoes a significant enhancement upon solvation. The α value increases from 379 a.u. in the gas phase to 807 a.u. in water and 802 a.u. in methanol, reflecting the pronounced effect of polar solvents in promoting electronic delocalization and polarizability. On the other hand, the value obtained in benzene (605 a.u.) shows a moderate increase, which is consistent with the comparatively lower dielectric constant of this apolar solvent. The polarizability anisotropy (Δα) exhibits a trend analogous to that of the static polarizability, with the maximum value observed in water (109 a.u.), indicative of pronounced directional electronic response in a highly polar medium. A significantly lower Δα value is obtained in benzene (53 a.u.), consistent with its reduced polarity and limited capacity to induce anisotropic polarization.
The first-order hyperpolarizability of compound A is significantly enhanced in solvent environments compared to the gas phase. In water, β0tot increases from 203 a.u. to 6857 a.u., representing an enhancement by a factor of 34. Similarly, methanol enhances the first hyperpolarizability by approximately 31 times, while benzene results in an increase of about 10 times relative to the gas-phase value. This trend clearly demonstrates the impact of solvent polarity, with polar solvents exhibiting a greater ability to enhance second-order NLO responses compared to apolar media. Furthermore, the hyper-Rayleigh scattering hyperpolarizability (β0HRS) exhibits a comparable trend, increasing by approximately 17 times in polar solvents and 4 times in the apolar solvent (benzene), relative to its gas-phase value.
Regarding the second hyperpolarizability, the value of γ(0; 0, 0, 0) in water is approximately 36 times higher than that in the gas phase, while in methanol and benzene, it is enhanced by factors of 16 and 19, respectively. These results clearly demonstrate the significant influence of solvation on the NLO behavior of the nanocluster A. We can conclude that solvent polarity plays a critical role in amplifying both second- and third-order NLO responses of nanocluster A. In a future study, we will investigate the effect of polar and nonpolar solvent environments on nanoclusters B to H.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5nj01220a |
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2025 |