Atazaz Ahsinac,
Aamna Qamarbc,
Sadegh Kaviani
d,
Safi Ullah Majide,
Jianwei Cao
*a and
Wensheng Bian
*ac
aBeijing National Laboratory for Molecular Sciences, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China. E-mail: caojw@iccas.ac.cn; bian@iccas.ac.cn
bBeijing National Laboratory for Molecular Sciences, State Key Laboratory of Polymer Physics and Chemistry, Chinese Academy of Sciences, Beijing 100190, China
cSchool of Chemical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
dInstitute of Physics, Kazan Federal University, 420008 Kazan, Russia
eDepartment of Chemistry, Abbottabad Campus, COMSATS University Islamabad, Abbottabad 22060, KPK, Pakistan
First published on 12th August 2025
Recognizing the increased stability of metalide complexes over alkali anions, research on transition metal-based metalides continues to grow. Here, employing various quantum chemical methods, we investigate novel superalkali-M (where M = Sc, Ti, V, Cr, and Mn) pairings theoretically, and our results indicate that the excited-state dynamic properties of these complexes are enhanced due to their metalide character and excess electrons. In the designed Li3O@[12-crown-4]M complexes, Li3O demonstrates excellent charge transfer through van der Waals (vdW) interactions. Ab initio molecular dynamic simulations demonstrate their kinetic and dynamic stabilities, and the thermodynamics of physisorption is also unveiled. A remarkable hyperpolarizability value of 7.6 × 105 a.u. is observed for the Sc complex, while the hyper-Rayleigh scattering value is highest for the Mn complex. The influence of different solvents on the hyperpolarizability response is thoroughly investigated. Moreover, dynamic NLO parameters are computed using various externally applied frequencies. The nature of bonding and the role of vdW interactions in triggering NLO responses are unveiled. In addition to the infrared analysis, the effect of vertical excitation energy has also been investigated in enhancing dynamic NLO features. We hope that these findings hold promise for advancing NLO applications of organic functional materials.
Alkalides containing alkali metal anions (e.g., Li−, Na−, K−, and Cs−) and electrides housing trapped electrons (anions) intercalated within an organic complexant are exemplary representatives of the excess electron family. Alkalides can be used as excellent nonlinear optical (NLO) materials9 and offer a significantly narrow band gap for condensed matter physics.10 Various conventional alkalides based on single alkali metals have been documented in the literature.12–17 On the other hand, non-traditional alkalides containing coinage metals (Cu, Ag, and Au),11 superalkali (Li3O)-based alkalides,12 and alkaline earth metal-based earthides (Be, Mg, and Ca) have been investigated and were grouped into excess electron compounds. However, substituting alkali metals with more stable d-block metals or clusters of atoms is a promising strategy for achieving thermodynamically stable complexes.12,13,20–23 For instance, Sun et al.11 proposed using coinage metals to design metalides (transition metal as an anion) as alternatives to alkali anions. Due to the higher ionization potential of coinage metals compared to alkali, the proposed metalide complexes have good thermodynamic and electronic stability. Doping of transition metals (particularly those in the 3d series) is expected to be a promising strategy for designing complexes with metalide character.15–21
For designing metalide and alkalide complexes, superalkali clusters are potential candidates.22 For example, a significant charge transfer in superalkali-M (where M = Li, Na, and K) pairing was observed from superalkali clusters to alkali metals, demonstrating alkalide characteristics.22 Superalkali clusters, such as Li3O, have lower ionization potentials (IPs) than alkali metals23,24 and strong reduction potential, making them intriguing substitutes for alkali metals in constructing alkalide or metalide complexes. In particular, Li3O stands out as an experimentally validated system with distinctive electronic properties and remarkably low ionization energy. Obviously, superalkali clusters very actively display electropositive properties and non-covalent interactions, promoting the physiochemical properties of the obtained complexes.12 Furthermore, it is known that the formation and stabilization of metalide complexes require stringent chemical environments and involve either a disproportionation or the reduction of one elemental metal by another metal or cluster of atoms.25 Generally, the reduction process and charge transfer are facilitated by strong stabilization of the metal cation by complexants such as crown ethers and cryptands.26 Crown ethers are heterocyclic polyethers capable of forming host–guest complexes, especially from the atomic scale to large molecular clusters.27,28 To our knowledge, the exploration of superalkali–transition metal pairing using [12-crown-4] has not yet been undertaken for optical and NLO applications. Although conventional metalides, based on single metal doping on complexes, have been reported,13,20,21 the effect on charge transfer in long-range interactions using superalkali–transition metal pairing would improve the stability of metalides and thus be a good choice for designing molecular functional materials.22,29 In addition, van der Waals (vdW) interactions are very important in many chemical systems and reactions,30–35 making it necessary to investigate the role of vdW interactions in the present designed complexes.
In this work, we present a detailed theoretical study on Li3O–M (M = Sc, Ti, V, Cr, and Mn) pairing on both sides of [12-crown-4]. In contrast to traditional alkalides and metalides based on single metals, a significant metalide characteristic has been observed in Li3O@[12-crown-4]M complexes. The designed metalide complexes show excellent charge transfer between Li3O and M, which is crucial for boosting NLO properties. The excited-state dynamic properties of these complexes are improved with the help of metalide character and excess electrons. The formation and stability of these complexes are also investigated via molecular dynamics calculations, and the important role of vdW interactions is underscored in triggering NLO features.
Energetics of complexes were investigated by computing several parameters, such as the interaction energy (ΔEint), Gibbs free energy of formation (ΔGf), adsorption enthalpy (ΔH), and adsorption entropy (ΔS) at 1 atm and 298.15 K. In particular, the complex formation strength is validated by computing ΔEint, which was computed using the following equations:
ΔEint = ELi3O@[12-crown-4]M − (E[12-crown-4] + ELi3O + EM) | (1) |
ΔECPint = ΔEint + BSSE | (2) |
ΔG = GLi3O@[12-crown-4]M − (G[12-crown-4] + GLi3O + GM) | (3) |
ΔH = HLi3O@[12-crown-4]M − (H[12-crown-4] + HLi3O + HM) | (4) |
![]() | (5) |
To further understand the electronic properties and superalkali-like nature of the designed complexes, the vertical ionization potential (VIP) and vertical electron affinity (VEA) were calculated using the following equations:
VIP = (E)+ − (E)0 | (6) |
VEA = (E)− − (E)0 | (7) |
To analyse the simultaneous interaction of Li3O and M on [12-crown-4] and the kinetic stability of complexes, ab initio molecular dynamics (AIMD) simulations were conducted using the ORCA 5.0 package.37 Due to the high computational cost for AIMD simulation at the ωB97xd/def2-tzvp level, the simulations were performed at the B3LYP-D3/def2-SVP level of theory,46 with the optimized geometry at the ωB97xd/def2-tzvp level used as the initial geometry. In AIMD simulations, each trajectory was simulated for 750 fs with a step size of 0.5 fs at 298.15 K. Furthermore, the metalide character of the Li3O@[12-crown-4]M complexes and charge transfer were explored using the frontier molecular orbital (FMO), natural bond orbital (NBO), electron localizing function (ELF), and localized orbital locator (LOL) analyses. Charge decomposition analysis (CDA) is also exploited using the same method to determine charge donation (Li3O → TM) and back donation (TM → Li3O) in present complexes. To determine the static NLO characteristics, the dipole moment (μo), polarizability (αo), first-order hyperpolarizability (βo), static second hyperpolarizability (γo), and scattering hyperpolarizability (βHRS) were calculated using the following equations:
μo = (μx2 + μy2 + μz2)½ | (8) |
αo = 1/3(αxx + αyy + αzz) | (9) |
![]() | (10) |
βx = βxxx + βxyy + βxzz, βy = βyyy + βyzz + βyxx and βz = βzzz + βzxx + βzyy |
〈γ〉 = 1/5(γxxxx + γyyyy + γzzzz + γxxyy + γxxzz + γyyxx + γyyzz + γzzxx) | (11) |
Here, βijk (i, j, k = {x, y, z}) is the tensor component of βo and γijkl (i, j, k, l = {x, y, z}) is the tensor component of second hyperpolarizability (γo). The second hyperpolarizability (γ) is the fourth rank tensor of 3 × 3 × 3 × 3 form, which exhibits both static and dynamic nature,
![]() | (12) |
DR = 〈βzzz2〉/〈βzxx2〉 | (13) |
In addition, the implicit solvation model based on density (SMD) has been considered to examine the effect of dielectric constants and protic solvent nature on hyperpolarizability response, with water, ethanol and N,N-dimethylformamide (DMF) considered.
Frequency-dependent (dynamic) nonlinear optical parameters were calculated at 532, 1064, 1300, and 1500 nm. Frequency-dependent first hyperpolarizability electro-optic Pockel's effects βEOPE(ω) = β(−ω;ω,0) and βESHG(ω) = β(−2ω;ω,ω) were obtained. The second hyperpolarizability γ(ω), dc-Kerr γdc-Kerr(ω) = γ(−ω;ω,0,0) and second harmonic generation γESHG(ω) = γ(−2ω;ω,ω,0) were also calculated similarly. Excited state studies were conducted by considering the time-dependent density functional theory (TD-DFT) approach at the same level of theory. In the present study, 60 electronic states were considered, including both singlets and triplets in determining the excited state parameters and absorbance wavelengths. To observe changes in vibrational modes after doping Li3O and M, we compared the vibrational frequency of complexes and pristine [12-crown-4] at the ωB97xd/def2-tzvp level. Furthermore, total density of states (TDOS) and projected density of states (PDOS) studies were conducted to unveil the orbital contribution and electrical properties of complexes.47
The topological features of complexes were analysed through non-covalent interaction (NCI) and quantum theory of atoms in molecules (QTAIM) analyses as implemented in Multiwfn48 and visual molecular dynamics (VMD) program.49 The NCI index relies primarily on the reduced density gradient (RDG) and electron density (ρ), and their correlation can be described using the equation given below:
![]() | (14) |
The stabilization energy (E2) during the electron delocalization between the donor (i) and acceptor (j) interacting orbitals was estimated through the NBO analysis. The following equation was used to calculate the E2 of the present complexes:
![]() | (15) |
Here, Fi,j is the off-diagonal NBO Fock matrix element and qi is the donor orbital occupancy. Ei and Ej are the diagonal elements (orbital energies).
Complexes | Symmetry | Spin | Interaction distances | Energies | ||||||
---|---|---|---|---|---|---|---|---|---|---|
dM–H (Å) | dM–O (Å) | dLi–O (Å) | ΔEint kcal mol−1 | ΔECPint kcal mol−1 | ΔGf kcal mol−1 | ΔH kcal mol−1 | ΔS kcal (mol K)−1 | |||
Li3O@[12-crown-4]Sc | C1 | Triplet | 3.23 | 3.59 | 2.52 | −48.74 | −34.93 | −26.09 | −44.47 | −0.06 |
Li3O@[12-crown-4]Ti | C1 | Doublet | 3.31 | 4.01 | 2.54 | −86.97 | −41.95 | −64.61 | −82.66 | −0.06 |
Li3O@[12-crown-4]V | C1 | Quintet | 3.43 | 4.80 | 2.54 | −98.25 | −98.36 | −68.12 | −93.69 | −0.08 |
Li3O@[12-crown-4]Cr | C1 | Sixet | 3.25 | 4.91 | 2.60 | −197.98 | −149.86 | −178.12 | −193.50 | −0.05 |
Li3O@[12-crown-4]Mn | C1 | Septet | 3.34 | 5.58 | 2.55 | −131.39 | −114.75 | −72.51 | −90.43 | −0.06 |
We can also see from Table 1 that the values of ΔEint and ΔECPint are lower than 23 kcal mol−1, indicating the physical interactions between the three components during complex formation.30 The much negative ΔEint (also see Table S1) justifies a strong complexation among three components of present complexes. Our computed interaction energies are slightly smaller than those reported for coinage-based metalides Cu–F6C6H6–Cu and Cu–F6C6H6–Ag.11 The strong binding of anions and cations with F6C6H6 could be the reason for their increased interaction energy. However, in the case of single alkali metal-based metalides Li–F6C6H6–Sc and Na–F6C6H6–Sc, and K–F6C6H6–Sc,20 the interaction energies of present complexes are significantly high, signifying the strong complexation of Li3O with transition metals. Hence, Li3O pairing with transition metals might be a better strategy for enhancing dynamic NLO features in the present complexes.
Table 1 also presents thermodynamic parameters, in which the thermodynamic feasibility and spontaneity of complex formation can be evidenced by the negative change in ΔGf of formation (ΔG < 0). A ΔG298.15 value of −21.35 kcal mol−1 is obtained for the Sc complex, indicating exothermic adsorption processes, which also agrees with its strong interaction energy. The feasibility of complex formation decreases progressively from the Sc to Mn complexes (with increased atomic number). This is reasonable since the increase of the number of d-electrons from Sc to Mn in the 3d-series leads to strong binding and less distortion due to the moderate interaction distance (dM–O). Excellent correlation is observed between the thermodynamic parameters (ΔG and ΔH) and the computed interaction energies. A similar pattern of reduction in ΔH298.15 can also be observed in Table 1.
The AIMD approach is reliable and versatile to unveil the thermodynamic and kinetic stability of molecular complexes at the atomic scale, offering significantly reduced computational costs relative to quantum dynamics simulations.51–54 In the obtained 1500 distinct geometries associated with the trajectory of the complex, no dissociation or fragmentation pattern is observed, indicating the stability of complex formation (adsorption) and thermodynamic stability at room temperature. The calculated root-mean-square deviation (RMSD) and kinetic energy are shown in Fig. 2. As can be seen from Fig. 2a, the root-mean-square distances increase up to 1.0 Å, and equilibrium during simulations is attained after 500 fs except for [12-crown-4], which relaxes after 600 fs. Although RMSD values increase in the beginning, no distortion in complexes is observed. The fluctuations are higher for Sc and Ti complexes, but decrease for V to Mn complexes. This can be correlated with nuclear charge and variable spin multiplicities of 3d-transition metals in complexes. The higher interaction energy of Cr and Mn complexes is consistent with its reduced RMSD trend, probing the relatively high dynamic stability among complexes. Also, a small fluctuation in the RMSD curve for Cr and Mn complexes might be correlated with strong binding of M having less distortion effect with [12-crown-4]. An increasing trend in total kinetic energy has been recorded up to 200 fs, where equilibrium is achieved after 300 fs. The snapshots of geometries during AIMD simulations are provided in Fig. S2. The conformations of each complex were taken after every 300 fs to ensure structural integrity during AIMD simulations.
Complexes | QM | QLi | Qsum | QMM | EHOMO | ELUMO | Eg | μo | VIP | VEA |
---|---|---|---|---|---|---|---|---|---|---|
Li3O | — | 0.45 | — | — | −7.17 | −0.14 | 7.03 | 0 | 5.14 | 4.44 |
[12-crown-4] | — | — | — | — | −8.74 | −3.06 | 5.68 | 0.84 | 5.72 | 4.25 |
Li3O@[12-crown-4]Sc | −0.93 | 0.71 | 0.18 | −0.22 | −3.61 | −0.31 | 3.30 | 5.19 | 0.27 | 3.48 |
Li3O@[12-crown-4]Ti | −0.35 | 0.92 | 0.92 | −0.20 | −3.71 | −0.88 | 2.83 | 5.70 | 1.09 | 1.36 |
Li3O@[12-crown-4]V | −0.23 | 0.96 | 0.92 | −0.82 | −3.03 | −0.35 | 2.68 | 19.24 | 2.45 | 1.18 |
Li3O@[12-crown-4]Cr | −0.10 | 0.93 | 0.91 | −0.87 | −2.49 | −0.59 | 1.90 | 18.55 | 3.27 | 0.82 |
Li3O@[12-crown-4]Mn | −0.01 | 0.94 | 0.90 | −0.02 | −2.67 | −1.03 | 1.64 | 10.10 | 4.08 | 0.35 |
Furthermore, the charge decomposition analysis (CDA) offers comprehensive insights into charge transfer and bonding interactions within the Li3O@[12-crown-4]M complexes by evaluating the primary contributions from the fragments ([12-crown-4], Li3O, and M) involved in charge transfer. Notably, a substantial charge transfer is observed from Li3O to M. Thus, the analysis focuses on the electron density contributions for donation (Li3O → M) and back donation (M → Li3O), which are summarized in Table S5. The findings correlate well with the aforementioned charge analysis (NBO) of the complexes, suggesting that the most prominent electron transfer occurs in the Li3O@[12-crown-4]Sc complex, which exhibits a value of 0.44 |e|. A similar trend of charge donation was observed in Li3O@[12-crown-4]Ti and Li3O@[12-crown-4]Mn complexes. However, the charge donation (Li3O → M) is slightly reduced in Li3O@[12-crown-4]V and Li3O@[12-crown-4]Cr complexes. The negative NBO charge on M decreases from Sc to Mn, which is also consistent with charge donation (Li3O → M) in CDA analysis. Furthermore, the negative repulsive polarization (r) observed in all complexes indicates a high charge density localized in the interaction region.
To further support the metalide character and excess electron nature, frontier molecular orbital (FMO) results are illustrated in Fig. 3. As can be seen, the highest occupied molecular orbital (HOMO) on the transition metals depicts the excess electronic cloud, indicating the negative NPA charges on transition metals. The HOMO and HOMO−1 densities have distribution across metal sites and on Li3O. This distribution suggests that excess electrons are available to participate in charge driven optical properties. HOMOs are similar to s-orbitals except for the V complex, which represents the d-orbital shape, while the LUMO varies in shape from dxy to dz2. The LUMOs of the V and Mn complexes have a shape similar to the dz2 orbital. The electron in HOMO of the Ti complex is associated with the dxy orbital, and the LUMO shape is reminiscent of the diffused s-orbital. On the other hand, molecular materials with significantly reduced HOMO–LUMO gaps (Eg) are highly desirable due to their potential optical and electrical properties. The [12-crown-4] has an Eg of 5.68 eV. The complexation of Li3O with M reduced the Eg of the resulting complexes, indicating the strong interactions through vdW forces. After simultaneous adsorption of M and Li3O on the complexant, a significant reduction in Eg values is observed, ranging from 1.64 to 3.30 eV (see Table 2). Among the complexes, the lowest Eg (1.64 eV) is observed for the Mn complex, while the highest one (3.30 eV) is found in the Sc complex. As shown in Table 2, the Eg decreases with the increased atomic number of M, which could be attributed to the gradually increased HOMO energies from the Sc complex to the Mn complex.
![]() | ||
Fig. 3 The plotted (a) ELF spectra, (b) HOMO and LUMO orbitals, and (c) localized orbital locator maps of the designed Li3O@[12-crown-4]M complexes (isovalue = 0.025) at the ωB97xd/def2tzvp level. |
Fig. 3 shows the outcomes of the electron localization function (ELF) and localized orbital locator (LOL) analyses. The ELF explains the electron pair density to manifest the surface topology of complexes with high ELF values indicated by red color and blue color manifesting delocalized regions. Examining the excess electron nature reveals high electron localization around transition metals. From Fig. 3a, we can see that a strong negative charge appears on the M atom in the Sc and Ti complexes, and the intensity of the red color decreases and delocalization increases on moving from left to right (Fig. 3). This can be attributed to the growing number of unpaired electrons in the d-orbital and the repulsion between them. Additionally, the increased interaction distance results in reduced electron localization. The localized orbital locator (LOL) is also given in Fig. 3c. Mathematically, the ELF and the LOL exhibit similar chemical mapping because they depend on the kinetic energy density. ELF is a dimensionless quantity, in the range of 0–1, to provide a visual description of the chemical bond in the present complexes. The ELF includes the Pauli kinetic energy density, while the LOL analysis does not include Pauli repulsion. The LOL further unveils the excess electron localization in d-orbitals of M in the present complexes. In addition, VIP and VEA values are computed for the present complexes, as shown in Table 2. The reduced VIP values suggest an increased tendency for electron donation, enhancing interactions among complex fragments. Notably, these values are considerably lower than those of individual Li3O and [12-crown-4]. Interestingly, the VIP values of these complexes are also lower than those of alkali metals. In fact, the VIP values for Sc, Ti, V, and Cr complexes are even smaller than the ionization potential of caesium (3.89 eV), the lowest among all elements. Readily electron loosing tendency of present complexes also indicates the formation of more stable anionic species. This can also be correlated with the potential interaction and stability of metalide complexes. For instance, VIP values of present complexes are significantly lower than those reported for coinage metalides M–F6C6H6–M (where M = Cu, Ag, and Au),11 M′–F6C6H6–M (where M′ = Li, Na, and K, and M = transition metals),20 and M–Janus–M (where M = Sc–Zn).21 In these metalides, the ns1 electrons of alkali metals are polarized and transferred to transition metals (M′), where negative NBO charges are observed on transition metals. Consequently, the present metalide complexes based on superalkali are intriguing candidates for further exploration of their NLO properties.
Complexes | αo | αx | αy | αz | βo | βx | βy | βz | βvec |
---|---|---|---|---|---|---|---|---|---|
[12-crown-4] | 1.0 × 102 | 1.1 × 102 | 1.1 × 102 | 8.7 × 101 | 60.21 | 1.2 × 10−2 | 6.0 × 102 | 60.01 | 60.02 |
Li3O@[12-crown-4]Sc | 4.3 × 102 | 3.4 × 103 | 7.5 × 102 | 9.0 × 102 | 7.6 × 105 | 1.6 × 104 | 4.4 × 104 | 4.3 × 104 | 6.2 × 103 |
Li3O@[12-crown-4]Ti | 4.2 × 102 | 1.8 × 103 | 6.5 × 102 | 8.6 × 102 | 5.9 × 104 | 1.9 × 104 | 4.2 × 104 | 3.7 × 104 | 3.6 × 103 |
Li3O@[12-crown-4]V | 2.0 × 102 | 2.6 × 103 | 1.5 × 103 | 1.9 × 102 | 8.3 × 103 | 8.2 × 103 | 8.2 × 102 | 8.5 × 102 | 8.5 × 102 |
Li3O@[12-crown-4]Cr | 2.0 × 102 | 2.5 × 103 | 1.9 × 102 | 1.8 × 102 | 7.9 × 103 | 7.9 × 103 | 6.5 × 102 | 1.6 × 102 | 7.9 × 103 |
Li3O@[12-crown-4]Mn | 4.1 × 102 | 7.3 × 101 | 4.5 × 102 | 1.5 × 102 | 1.7 × 104 | 7.6 × 103 | 1.5 × 104 | 7.0 × 101 | 8.2 × 103 |
![]() | ||
Fig. 4 Components of NLO properties: (a) polarizability (αo) and (b) hyperpolarizability (βo) of the Li3O@[12-crown-4]M (where M = Sc, Ti, V, Cr, and Mn) complexes using the ωB97xd/def2tzvp method. |
The nonlinear optical effect is estimated through the calculated static first hyperpolarizability (βo) values, ranging from 7.6 × 105 to 1.7 × 104 a.u., where the highest value is observed for the Sc complex (also see Fig. 5a). The remarkable βo value of the Sc complex is 12622 times higher than that of pristine [12-crown-4] (see Table 3). The significant NLO response of the Sc complex may be attributed to the strong metalide character and the reduced VIP value. Similarly, the presence of excess electrons decisively decreases the excitation energy ΔE (2.76 eV) of the Sc complex. The remaining Ti, V, Cr, and Mn complexes have βo values of 5.9 × 104, 8.3 × 103, 7.9 × 103, and 1.7 × 104 a.u., respectively. The βo value monotonically decreases from Sc to Cr, while the Mn complex shows a slightly higher response. The rising trend of βo can be linked to the electronic properties such as VIP, HOMO–LUMO gaps, metalide character, and the nuclear charge on M. Complexes with strong metalide character, reduced VIP, and low excitation energies, resulting in excellent NLO features. In αo, the αx component is dominant, but βz contributes significantly to the total hyperpolarizability response. Furthermore, values of projection of hyperpolarizability on dipole moment (βvec) are also given in Table 3. The results of βvec are consistent with those of βo, which confirms that the designed metalides are remarkable NLO materials.
Table 4 lists the static second hyperpolarizability (γo), which is the third-rank NLO response for molecular materials, and the average contribution to hyperpolarizability. The γo value of [12-crown-4] is 9.9 × 103 a.u., while the values dramatically increase up to 5.6 × 106 a.u. for the V complex. The significant γo response of the V complex might be attributed to its noticeable μo and reduced Eg. Overall, the γo values range from 1.7 × 105 to 5.6 × 106 a.u., demonstrating excellent third-order NLO properties. A comparison of the NLO properties is made with reported metalides and transition metals containing simple excess electron compounds. The exceptional NLO features of current metalides reveal why they are chosen over traditional alkalides and simple excess electron complexes. In the present study, βo of the Sc complex is higher than those of coinage metalides19 and transition metal-based Janus metalides.21 Likewise, the unprecedented involvement of Li3O in regulating vdW forces and charge transfer also makes them interesting candidates over single metal-based alkalides and metalides. The involvement of vdW forces and charge transfer to transition metals render them appealing competitors over single metal-based alkalides and metalides for NLO applications.58,59
Complexes | γo | βHRS | 〈βj=1〉 | 〈βj=3〉 | Φ(βj=1) | Φ(βj=3) | DR |
---|---|---|---|---|---|---|---|
[12-crown-4] | 9.9 × 103 | 46.33 | 34.13 | 1.4 × 102 | 0.19% | 0.80% | 1.75 |
Li3O@[12-crown-4]Sc | 5.7 × 105 | 9.1 × 103 | 1.6 × 104 | 1.4 × 104 | 0.53% | 0.47% | 4.69 |
Li3O@[12-crown-4]Ti | 9.6 × 105 | 8.7 × 103 | 1.6 × 104 | 1.3 × 104 | 0.54% | 0.45% | 4.88 |
Li3O@[12-crown-4]V | 5.6 × 106 | 7.1 × 103 | 1.2 × 104 | 1.3 × 104 | 0.48% | 0.51% | 4.02 |
Li3O@[12-crown-4]Cr | 1.7 × 105 | 8.4 × 102 | 1.6 × 103 | 1.2 × 103 | 0.57% | 0.42% | 5.34 |
Li3O@[12-crown-4]Mn | 4.7 × 106 | 2.1 × 104 | 4.2 × 104 | 2.3 × 104 | 0.64% | 0.35% | 6.42 |
The effect of solvent on the NLO properties was investigated at the ωB97xd/def2tzvp level. An implicit solvent effect using the universal solvent based on solute electron density (SMD) is considered,60 and the results are reported in Table S3. In the implicit solvent effect, water with a dielectric constant of 78.35 exhibits a slightly higher βo response as compared to ethanol, but Cr and Mn complexes hold significant βo values in ethanol (see Fig. 6). Overall, entire complexes show a better hyperpolarizability response in the solvent medium as compared to the gas phase. The enhancement in βo is attributed to structural changes in complexes influenced by solvent polarity. In water and ethanol, an excellent increase in the dipole moment is recorded, influencing the hyperpolarizability response. Hence, strong polarization effects and charge fluctuations are key parameters in enhancing the static NLO properties using the SDM model.
![]() | ||
Fig. 6 The plotted hyperpolarizability (βo) with implicit solvents at the ωB97xd/def2tzvp/SMD (water, ethanol, and DMF) level. |
The calculated hyper-Rayleigh scattering first hyperpolarizability (βHRS) and its components are also given in Table 4 and Fig. 5b. βHRS is a parametric optical effect that occurs when two incoming photons of frequency ω annihilate, resulting in a scattered photon of frequency 2ω. A substantial enhancement of the βHRS is seen for the designed metalides as compared to the pristine complexant. βHRS values range from 8.4 × 102 to 2.1 × 104 a.u., where the Mn complex holds the highest value of scattering and the first hyperpolarizability among the series. The highest value for the Mn complex might be credited to its significant dipole moment and lowest HOMO–LUMO gap (Eg). In addition, the βHRS response could be related to the d5 electron system and delocalization of electrons in the Mn complex. The significant dipolar contribution percentage might be a contributing factor for the increased βHRS of the Mn complex. Unlike previous studies on similar systems, the present complexes show no correlation between the dipole moment and βHRS.61–63 However, a correlation exists between Eg and the dipole moment in the studied complexes. As the interaction distance between the metal (M) and the oxygen/hydrogen atoms of [12-crown-4] increases, the value of Eg gradually decreases. Concurrently, the dipole moment follows an increasing trend from Sc to Mn complexes. The values of 〈βj=1〉 and 〈βj=3〉 are also higher for the Mn complex. From Fig. 5d, an increased dipolar contribution to hyperpolarizability Φ(βj=1) can be observed with the depolarization ratio (DR), suggesting the dipolar nature of complexes.
The NLO response of complexes can be separated into dipolar Φ(βj=1) and octupolar Φ(βj=3) contributions to total hyperpolarizability,59 and the contribution percentages to hyperpolarizability are given in Table 4. The Φ(βj=1) dipole characteristics prevail in cases where charge transfer and dipole moment are almost one-dimensional. The pure [12-crown-4] ether has an octupolar nature as revealed by significant octupolar contribution percentage with Φ(βj=3) ≈ 79% to total hyperpolarizability. In contrast, a notable dipolar contribution to hyperpolarizability is observed in Li3O@[12-crown-4]M complexes, signifying their dipolar nature. Furthermore, the dipolar nature of complexes can also be observed from their computed depolarization ratio (DR). In particular, a small DR value of 1.75 is calculated for the pure complexant, which evidences its octupolar nature, while the designed complexes reflect high DR values. The DR values oscillate between 1.5 (octupolar) and 9 (dipolar). For the present metalides, the DR value is between 4.02 and 6.42, confirming the dipolar nature of these complexes.
Frequency | [12-crown-4] | Li3O@[12-crown-4]Sc | Li3O@[12-crown-4]Ti | Li3O@[12-crown-4]V | Li3O@[12-crown-4]Cr | Li3O@[12-crown-4]Mn | |
---|---|---|---|---|---|---|---|
Frequency-dependent first hyperpolarizability β(ω) (a.u.) | |||||||
0.085 (532 nm) | β(−ω;ω,0) | 49.32 | 1.5 × 105 | 5.0 × 105 | 1.5 × 105 | 6.0 × 104 | 2.6 × 105 |
β(−2ω;ω,ω) | 64.23 | 9.8 × 104 | 3.6 × 104 | 9.8 × 104 | 1.8 × 105 | 2.6 × 104 | |
0.042 (1064 nm) | β(−ω;ω,0) | 49.10 | 1.2 × 105 | 1.2 × 105 | 5.5 × 103 | 4.6 × 103 | 1.4 × 105 |
β(−2ω;ω,ω) | 48.21 | 1.9 × 105 | 2.2 × 105 | 1.1 × 104 | 4.4 × 103 | 1.1 × 105 | |
0.035 (1300 nm) | β(−ω;ω,0) | 46.46 | 2.0 × 105 | 2.2 × 105 | 2.9 × 105 | 9.4 × 103 | 2.5 × 104 |
β(−2ω;ω,ω) | 47.46 | 1.3 × 106 | 9.9 × 106 | 1.3 × 106 | 1.7 × 105 | 1.6 × 105 | |
0.030 (1500 nm) | β(−ω;ω,0) | 46.34 | 7.5 × 104 | 6.4 × 104 | 7.1 × 104 | 2.5 × 103 | 3.2 × 105 |
β(−2ω;ω,ω) | 47.08 | 7.0 × 104 | 8.4 × 104 | 8.1 × 104 | 4.9 × 103 | 1.3 × 105 | |
Frequency-dependent second hyperpolarizability γ(ω) (a.u.) | |||||||
0.085 (532 nm) | γ(−ω;ω,0,0) | 1.0 × 104 | 7.7 × 108 | 3.1 × 1010 | 7.9 × 109 | 8.6 × 107 | 8.0 × 109 |
γ(−2ω;ω,ω,0) | 1.6 × 104 | 7.7 × 107 | 1.6 × 1010 | 6.5 × 107 | 1.1 × 108 | 4.5 × 108 | |
0.042 (1064 nm) | γ(−ω;ω,0,0) | 1.0 × 104 | 1.0 × 104 | 6.6 × 107 | 3.9 × 106 | 1.3 × 107 | 2.1 × 107 |
γ(−2ω;ω,ω,0) | 1.1 × 104 | 1.1 × 104 | 8.3 × 108 | 1.9 × 107 | 1.9 × 106 | 3.2 × 108 | |
0.035 (1300 nm) | γ (−ω;ω,0,0) | 4.3 × 104 | 9.2 × 106 | 1.0 × 107 | 2.6 × 106 | 2.0 × 1011 | 1.8 × 107 |
γ(−2ω;ω,ω,0) | 4.5 × 104 | 1.6 × 107 | 2.4 × 107 | 1.0 × 107 | 1.4 × 1010 | 1.4 × 107 | |
0.030 (1500 nm) | γ (−ω;ω,0,0) | 4.3 × 104 | 2.3 × 106 | 3.9 × 106 | 3.3 × 107 | 2.3 × 1011 | 4.5 × 106 |
γ(−2ω;ω,ω,0) | 4.4 × 104 | 5.5 × 108 | 2.5 × 108 | 7.7 × 105 | 5.0 × 109 | 1.0 × 107 |
Likewise, frequency-dependent second hyperpolarizability γ(ω) values are also provided in Table 5. Overall, the values of γ(ω) are quite higher than the corresponding static ones, indicating the influence of externally applied frequencies on the NLO features. In the series, the Ti and Cr complexes exhibit extraordinary responses for γ(ω). γ(−ω;ω,0,0) is determined for the dc-Kerr effect, and γ(−2ω;ω,ω,0) is determined for the second harmonic generation (SHG) effect.64 The complexes exhibit better dynamic NLO responses at higher dispersion frequency, and the dc-Kerr effect becomes less intense at 1500 nm except for the Cr complex. The dc-Kerr effect is the effect of an instantaneously occurring nonlinear response and plays a crucial role in photonics and optoelectronics. The Cr complex is more responsive at the applied frequency of 1500 nm. Entire metalides show a higher dynamic γ(ω) response at high external frequencies. For these complexes, the dc-Kerr effect is distinct at a small applied frequency (532 nm) as compared to the 1064 nm frequency. Notably, the Mn complex exhibits the highest value of 3.1 × 1010 a.u. for the dc-Kerr effect, while the Cr complex holds the lowest response. In addition, quantitative analysis at 1500 nm reveals an inverse correlation between the VIP and γ(−ω;ω,0,0). A systematic decrease in VIP from Sc to Mn complexes generally corresponds to enhanced γ(−ω;ω,0,0) responses, with the exception of Mn complexes showing a minor reduction. Notably, the trend in γ responses contrasts with the decreasing metalide character across the series.
To further identify vdW forces and their important roles in electronic and NLO properties, the noncovalent interaction (NCI) analysis is considered. The RDG versus sign(λ2)ρ scattered plots and the colour-filled RDG isosurface for Sc and Mn are shown in Fig. 7, while the remaining pictures are shown in Fig. S3. It can be seen from Fig. 7a that the RDG isosurface of Li3O@[12-crown-4]M complexes holds green-coloured layered patches, suggesting vdW interactions. Moreover, the appearance of more spike blue regions in the Sc complex signifies the stronger attraction between the Sc atom and [12-crown-4], which is in agreement with the short interaction distance. The observed significant vdW interactions in the Sc complex can be attributed to its shorter interaction distance (dM–H) and considerable NBO charge transfer, resulting in a noteworthy hyperpolarizability response. It can also be observed that the ionic pairing Li3O+–M− is predominantly dominated by attractive interactions followed by vdW forces. The impact of vdW interaction is highly observed for M lying exactly in the middle of [12-crown-4]. This is reasonable because the charge transferred from Li3O to M is significant, resulting in increased stability for the guest–host system as well as the NLO response of complexes.
The second-order perturbation energy (E2) is also examined to gain a better comprehension of donor–acceptor orbital interactions and their contribution to the stability of the Li3O@[12-crown-4]M complexes. Table S5 lists the corresponding results, detailing intermolecular charge transfers and electron delocalization. The E2 is used to measure the amount of electron delocalization between the donor (i) and acceptor (j) interacting orbitals as described in eqn (15). In the present complexes, the interaction between lone pair (LP) electrons of the oxygen atom of Li3O (donor) and LP* orbitals of M (acceptor) is the most significant donor–acceptor interaction. This delocalization of electrons in Li3O plays a crucial role in maintaining the polar environment and its charge-donating ability to metals. Therefore, a significant E2 value indicates a strong interaction existing between the donor and acceptor orbitals, resulting in notable stability of complexes. From the results collected in Table S5, it can be found the lowest E2 value (0.06 kcal mol−1) obtained for the Mn complex displays that it has the minimum stabilization energy for charge transfer, while a significant value (31.45 kcal mol−1) is calculated for the Sc complex, demonstrating strong donor–acceptor interactions in this complex. These results are consistent with the calculated NBO charge analysis as well as with the interaction energy of the Sc complex. It should be noted that E2 is not the sole factor enhancing NLO responses. The magnitude of E2 is strongly dependent on both the orbital overlap and the energy difference between the donor and acceptor orbitals at specific interaction distances.
Complexes | ΔE (eV) | fo (a.u.) | λmax (nm) | Major orbital contribution |
---|---|---|---|---|
[12-crown-4] | 9.70 | 0.2273 | 128 | HOMO → LUMO+6 (24.5%) |
HOMO → LUMO+5 (12.5%) | ||||
Li3O@[12-crown-4]Sc | 2.76 | 2.7685 | 447 | HOMO → LUMO (30%) |
HOMO → LUMO+3 (34%) | ||||
HOMO → LUMO+2 (9%) | ||||
Li3O@[12-crown-4]Ti | 3.00 | 0.3283 | 411 | HOMO → LUMO+2 (37%) |
HOMO → LUMO+3 (24%) | ||||
HOMO → LUMO+5 (18%) | ||||
Li3O@[12-crown-4]V | 3.07 | 0.4554 | 402 | HOMO → LUMO+6 (68%) |
HOMO → LUMO+7 (24%) | ||||
Li3O@[12-crown-4]Cr | 3.57 | 0.2543 | 346 | HOMO → LUMO+6 (10%) |
HOMO → LUMO+8 (11%) | ||||
Li3O@[12-crown-4]Mn | 3.69 | 0.3312 | 335 | HOMO → LUMO+6 (16%) |
HOMO → LUMO+8 (22%) |
In the gas phase, vibrational modes associated with the pristine crown-ether and the designed Li3O@[12-crown-4]M complexes have been investigated. The FT-IR spectra of [12-crown-4] and the Li3O@[12-crown-4]M complexes are presented in Fig. S5, while their vibrational frequencies are listed in Table S6. νC–H stretching modes in [12-crown-4] are observed in the range of 3073–2976 cm−1 in FT-IR spectra, closely matching the experimental values of 2947–2865 cm−1.68–70 Furthermore, three vibrations associated with νC–H bending modes, with the calculated intensities of 1496, 1401, and 1134 cm−1 are also in agreement with experimental results.69 Characteristic peaks associated with crown ethers are C–O and C–C stretching. Strong peaks observed at 1227 cm−1 are associated with C–O stretching, while the medium peak at 1075 cm−1 belongs to C–C stretching. The calculated vibrational frequencies for the designed complexes correspond to stretching and bending vibrations in the range of 3029–1270 cm−1. The highest C–H stretching and bending for the Cr complex are observed at 3029 cm−1 and 1270 cm−1, respectively. Additionally, the calculated C–O stretching modes for Sc, Ti, V, Cr, and Mn are 1203, 1207, 1216, and 1323 cm−1. In conclusion, after the interaction of Li3O and transition metals with [12-crown-4], the vibrational peaks become more intense, indicating the change in the vibrational level after complexation. In the pure complexant, the C–O stretching vibration (1206 cm−1) is smaller as compared to that in the designed complexes. The vibrational frequencies align in the order of Cr > V > Ti > Mn > Sc. The obtained values are 1068, 1069, 1072, 1093, and 1067 cm−1, which are consistent with the experimental results.67,68
Furthermore, to better understand the electronic properties of the Li3O@[12-crown-4]M complexes and orbital contributions concerning energy, the projected density of state (PDOS), as well as total density of state (TDOS) spectra, are plotted and given in Fig. S6 and S7. PDOS provides us with information about the number of states concerning energy. The energy in the PDOS plot spreads from −21 to 5.44 eV. A significant shift to high energy in the HOMO can be observed compared to [12-crown-4]. Newly generated HOMOs are observed after the interaction of Li3O and M with the complexant. Degenerated states are higher in energy and have a contribution from the d-orbital from transition metals. A reduction in the HOMO–LUMO gap is witnessed for complexes, promoting electron transfer from occupied to unoccupied orbitals, which leads to imparting optical NLO characteristics.
Supplementary information available: Optimized molecular geometries, snapshot of geometries in AIMD calculations, RDG scattered plots, absorption spectra, IR spectra, PDOS and TDOS spectra, the computed interaction energies, optimized geometric coordinates, static NLO parameters, topological parameters, and vibrational frequencies of the studied complexes. See DOI: https://doi.org/10.1039/d5nj00827a
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2025 |