Next Article in Journal
Development of Computational In Silico Model for Nano Lipid Carrier Formulation of Curcumin
Next Article in Special Issue
Properties of Naked Silver Clusters with Up to 100 Atoms as Found with Embedded-Atom and Density-Functional Calculations
Previous Article in Journal
Potential Anti-Alzheimer Properties of Mogrosides in Vitamin B12-Deficient Caenorhabditis elegans
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Geometric, Electronic, and Optoelectronic Properties of Carbon-Based Polynuclear C3O[C(CN)2]2M3 (where M = Li, Na, and K) Clusters: A DFT Study

1
Department of Chemistry, College of Science, King Faisal University, Al-Ahsa 31982, Saudi Arabia
2
Department of Chemistry, Abbottabad Campus, COMSATS University Islamabad, Abbottabad 22060, KPK, Pakistan
3
Chemical Sciences, Faculty of Science, University Brunei Darussalam, Jalan Tungku Link, Gadong BE1410, Brunei
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(4), 1827; https://doi.org/10.3390/molecules28041827
Submission received: 31 December 2022 / Revised: 5 February 2023 / Accepted: 6 February 2023 / Published: 15 February 2023
(This article belongs to the Special Issue Computational Study of Non-metal and Metal Clusters)

Abstract

:
Carbon-based polynuclear clusters are designed and investigated for geometric, electronic, and nonlinear optical (NLO) properties at the CAM-B3LYP/6-311++G(d,p) level of theory. Significant binding energies per atom (ranging from −162.4 to −160.0 kcal mol−1) indicate excellent thermodynamic stabilities of these polynuclear clusters. The frontier molecular orbital (FMOs) analysis indicates excess electron nature of the clusters with low ionization potential, suggesting that they are alkali-like. The decreased energy gaps (EH-L) with increased alkali metals size revael the improved electrical conductivity (σ). The total density of state (TDOS) study reveals the alkali metals’ size-dependent electronic and conductive properties. The significant first and second hyperpolarizabilities are observed up to 5.78 × 103 and 5.55 × 106 au, respectively. The βo response shows dependence on the size of alkali metals. Furthermore, the absorption study shows transparency of these clusters in the deep-UV, and absorptions are observed at longer wavelengths (redshifted). The optical gaps from TD-DFT are considerably smaller than those of HOMO-LUMO gaps. The significant scattering hyperpolarizability (βHRS) value (1.62 × 104) is calculated for the C3 cluster, where octupolar contribution to βHRS is 92%. The dynamic first hyperpolarizability β(ω) is more pronounced for the EOPE effect at 532 nm, whereas SHG has notable values for second hyperpolarizability γ(ω).

1. Introduction

Nonlinear optical (NLO) materials are at the front line of research in interdisciplinary science and laser-based technology due to their fundamental applications in the field of optoelectronics [1,2,3,4]. Photonic devices, laser-based technology, endoscope, and sensors are examples of well known technologies where NLO materials have possible applications [5,6,7,8,9,10,11,12]. To design and synthesize the NLO materials, much efforts are exerted to understand the origin of nonlinearity in molecules and clusters in order to correlate NLO responses to electronic structure and molecular geometry. Polarization, asymmetric charge distribution, asymmetric crystal packing, and π-conjugated electron transport routes are all required for NLO materials. Because of their high thermal stability and transparency, inorganic nonlinear optical materials have been preferred over organic ones [13]. Some inorganic borates crystals, such as KB5 (KB5O8H2O), BBO (-BaB2O4), and LiB3O5 (LBO), have been investigated as good NLO materials, particularly in the ultraviolet range [13].
For obtaining high-performance NLO materials, several strategies were proposed, which include bond length alternation (BLA) [14], doping metal atoms [15], push–pull mechanisms from donor to acceptor [16], modification of sp2 hybridized carbon nanomaterials [17], designing octupolar molecules [18], multidecker sandwich complexes [19], and excess electron induction [20].
The introduction of excess electrons into molecules and clusters is the most viable technique to escalate hyperpolarizability. The availability of loosely bound electrons predominantly decreases the excitation energies for the crucial transition [21,22,23]. Excess electrons in molecules and crystals behave similarly to Rydberg orbitals, which are positioned outside the parent molecules and are held loosely [24,25,26]. Several studies have demonstrated the substantial role of the diffuse excess electrons in compounds for developing NLO materials. Wei Chen et al. investigated the Li@calix[4]pyrrole electride complex, which has a significant static hyperpolarizability (βo) value of up to 7.3 × 103 au, where the presence of excess electrons has a significant role [27].
Theoretically designed compounds having excess electrons that are further classified into [28] alkalides [29], alkaline-earthides [30], metalides [31], and electrides [32]. Alkalides are complex compounds in which alkali metals bear the negative charge (Li, Na, K) [33]. On the other hand, electride complexes have anionic sites occupied by the electron inside the complexes [34]. Furthermore, the alkaline–earthides were recently introduced to excess electrons compounds, where the alkaline earth metals hold a negative charge [35]. Interestingly, superalkali clusters are a new class of materials that can transport electrons, making them useful for the fabrication of electro-optical materials [36].
Superalkalis clusters with lower ionization energy (IE) than alkali metal elements are well known due to their powerful reducing capabilities. The very first report about superalkalis was obtained in 1982 by Gutsev and Boldyrev for Li3O, Li2F, and Li4N clusters [37]. These clusters with unique qualities, such as tuneability in their electrical properties and the ability to function as a bridge between micro and macro materials, are of great interest to cluster science. Recent advances in cluster science show the potential applications of superalkali clusters, i.e., the reductive materials, helium and hydrogen storage, catalysis, supersalt formation, and nonlinear optics [38,39,40,41].
Superalkali clusters are excellent candidates for creating optical and NLO materials because of their excellent tunable electronic and structural properties. The decreased excitation energy may be responsible for electrons shifting from HOMO to LUMO, as they are loosely bound. Based on the intriguing characteristics of superalkali clusters, these were used to fabricate NLO materials. In this regard, two-dimensional materials doped superalkali, and they play an essential role in triggering the hyperpolarizability response. Sun et al. theoretically designed superalkali-based alkalides Li3+(calix[4]pyrrole)M, Li3O+(calix[4]pyrrole)M, and M3O+(calix[4]pyrrole)K (M = Li, Na, and K), where the hyperpolarizability response is recorded up to 34 718 au [42]. Similarly, Faizan Ullah et al. reported a noticeable enhancement in the NLO response of the A112P12 nanocluster by using Li4N, Li2F, and Li3O superalkalis as the source of the excess electrons [43]. Furthermore, macrocyclic oligofurans ring doped with superalkali clusters were also reported as a new kind of nonlinear optical material where a larger hyperpolarizability response is attributed to the presence of loosely bound electrons [44].
Although a larger number of superalkali clusters were theoretically designed, very limited studies have been conducted to show the possibility of using polynuclear superalkali (undoped) clusters as NLO materials. Srivastava et al. investigated the electronic and nonlinear optical properties LinF (n = 2–5) and M2X small clusters as excess electron compounds where the βo increases up to 105 au for Li2F [45,46]. Our group investigated the static and dynamic hyperpolarizability response of M2OCN and M2NCO (M = Li, Na, K) superalkali clusters as excess electrons candidates where the second hyperpolarizability γ(ω) values were calculated up to 2.1 × 108 au [45].
Superalkali clusters can be mononuclear, bimetallic, and polynuclear based on their rational design and elemental composition. We are interested to investigate carbon-based polynuclear clusters for electronic and NLO properties. These clusters are more stable than conventional mononuclear superalkali clusters and might possess better electronic and NLO properties. The previous development in the family of superalkalis and their tunable electronic properties prompted us to further investigate polynuclear clusters for optical and NLO properties. Polynuclear carbon-based clusters C3O[C(CN)2]2M3 (where M = Li, Na, and K) are investigated using DFT.

2. Results and Discussion

2.1. Optimized Geometries and Thermodynamic Stabilities

The optimized geometries of carbon-based polynuclear superalkali clusters C3O[C(CN)2]2M3 (where M = Li, Na, and K) optimized at CAM-B3LYP/6-311++g(d,p) are given in Figure 1. The studied polynuclear structures (C1 to C3) show C2V point group symmetry (Table 1). These clusters are planar with a central carbon core. The determined bond distances between alkali metal and oxygen (d-M-O) are in increasing fashion with the increased size of metals (Li to K). The calculated d-M-O bond distances for C1, C2, and C3 are 1.82, 2.24, and 2.58 Å, respectively (Table 1). The observed geometric parameters (Supplementary Materials) are very consistent with the previously reported study in the literature. Furthermore, these polynuclear clusters also show the increased bond distance between metals and nitrogen (dM-N). The observed bond lengths (dM-N) are 1.89, 2.26, and 2.62 Å for C1, C2, and C3 clusters. The observed monotonic increase in bond lengths from Li to K may be attributed to increased metal size. The performed frequency calculation shows that there is no imaginary frequency associated with these clusters, and these are true minima on the potential energy surface.
The thermodynamic stability of the studied polynuclear clusters is evaluated through calculated binding energy per atom (Eb). Overall, the binding energies range from −160.1 to −162.1 kcal mol−1 (Table 1), where the highest energy is found for C1, while the lowest is observed for the C2 cluster. The obtained significant binding energies per atom suggests their thermodynamic stabilities. The calculated binding energies are higher in comparison to previously reported superalkali clusters NM’M (where M = Li, Na and K), C3X3Y3 (X = O, S, and Y = Li, Na and K) and bimetallic superalkali clusters [47,48]. The trend of binding energies per atom for studied clusters is also shown in Figure 2. Compared to clusters C2 and C3, cluster C1 has a greater binding energy value. The computed binding energies show high thermal stability of these clusters, which demonstrate that they can be synthesized experimentally.

2.2. Electronic Properties and Stability

The electronic stability and superalkali nature of these clusters can be observed from calculated ionization potential and electron affinity. The obtained vertical ionization potential values are smaller than Cs-atom (3.89 eV), which shows the superalkali characteristics of these clusters. These values are also significant and account for the electronic stabilities of these clusters. The highest VIE value of 3.65 eV is found for C1, while the lowest value (3.0 eV) is indicated for the C3 cluster (Table 1). A gradual decrease in VIE values with the increased size of alkali metals can be seen in these clusters. On the other hand, the vertical electron affinity (VEA) values range from 0.27 to 0.89 eV, where C3 shows the lowest value. The reduced values of EA indicate the electropositive nature of these clusters.
To obtain reactivity and charge distribution, the computed NBO charges are given in Table 1. The NBO charges (positive) on alkali metals slightly increase from Li to K metals. The charge is transferred from alkali metals to electronegative atoms (oxygen and nitrogen) within clusters. The NBO charges on alkali metals (QM) lie in the range of 0.58 to 0.62 e, where C1 shows higher charge (positive magnitude) on metals. The charge transferred from alkali to O-atom is more pronounced as compared to the alkali to N-atom transition, which may be attributed to the higher electronegativity of the oxygen atom. The calculated NBO charges on O-atom (QO) lie in the range of −0.83 to −0.96 e and are higher for small-sized metals.

2.3. Global Reactivity Descriptor

To characterize the reactivity of these clusters, we calculated global reactivity descriptor, chemical hardness, softness, and chemical potential (Table 2). The chemical hardness is measured as resistance to change in electronic distribution within clusters. The results obtained show that the C3 cluster has the highest value (1.839 eV) of hardness, whereas the C1 has the lowest value. The size of alkali metals is an obvious factor in controlling the hardness of clusters. The decreased values show a correlation with increased atomic size (Li to K), which guarantees soft nature and reactivity (Table 2). Similarly, the values of chemical softness (S) increase from C1 to C3 and reach the maximum of 0.33 eV.
The chemical potential values are also calculated and given in Table 2. The higher chemical potential (χ) values show the escaping tendency of the electrons in clusters and molecules. Obtained significant values (negative) indicate the stability of these polynuclear clusters. These values also suggest that the clusters do not decompose spontaneously into atoms and possess reasonable electronic stability.

2.4. FMO Analysis and Excess Electron Nature of Clusters

To provide deep insight into the electronic structures of the studied clusters, the densities of the highest occupied molecular orbitals (HOMO) and virtual orbitals are plotted, and their energy values are given in Table 2. The HOMO and LUMO are quite important in quantum chemistry, as they allow the prediction of chemical stability and reactivity of molecules. Imperatively, the small difference between HOMO-LUMO (EH-L) is crucial for the description of reactivity of molecules. The smaller EH-L gaps depict greater chemical reactivity with a high tendency to be polarized, as well as low kinetic stability. The HOMO-LUMO gap values lie in the range of 4.08 to 1.96 eV, where the highest value corresponds to C1 clusters, while the lowest values correspond for C3. One can note that EH-L decreases with increased metals size (Li to K) within clusters. Furthermore, decreased EH-L gaps for the studied clusters can be attributed to increased energies of occupied orbitals where the energy of virtual orbitals goes on decreasing.
The reactivity and conducting qualities of these clusters are revealed by a significant reduction in HOMO-LUMO gaps; there are excitable valence electrons (excess electrons) with transition HOMO → LUMO. The excess electron nature is further justified by the distribution of HOMO densities, and the electronic density cloud is mainly spread over alkali metals, which indicates the excess electron character of these superalkali clusters. The three-dimensional HOMO density of C1 is shaped as a s-orbital, while for C2 and C3, its look like a diffuse p-orbital (Figure 3). The LUMO densities that are generated are spherical and resemble s-orbitals.

2.5. Electrical Conductivity (σ)

The electrical conductivity is also a crucial aspect to demonstrate the NLO properties of molecules. The electrical conductivity (σ) is the function of energy gaps (EH-L); thus, narrowing HOMO-LUMO gaps more will lead to higher electrical conductivity of materials. In our designed clusters, the HOMO–LUMO gaps are significantly reduced from 4.08 to 1.96 eV. The electrical conductivity increases with increased size of alkali metals, which might be attributed to ease in excitation of electrons (HOMO to LUMO).

2.6. TD-DFT Analysis

In the transparent region, the applications of nonlinear optical materials can be better understood. The obtained TD-DFT parameters of crucial transitions and first allowed transitions are given in Table 3. The percentage contribution of particular orbitals of these clusters for both transitions are also given in Table 3, whereas spectra are shown in Figure 4. The higher value of ϵ shows strong absorption at particular wavelength. Additionally, a higher value of fo reveals the strong transition probability. The studied cluster C3 has significant value of ϵ and oscillator strength at higher wavelength. The absorption maxima (λmax) during crucial transition for C1, C2, and C3 are 758, 688, and 995 nm, respectively, where the redshifted (i.e., bathochromic sift) in λmax is observed for C3 (Table 3). The obtained excitation energies of crucial transition are 1.63, 0.92, and 1.24 eV for C1, C2, and C3 clusters. On the other hand, the obtained optical gaps during allowed transitions are 1.63, 0.92, and 0.86 eV. The C1 cluster has same value for crucial excitation and optical gap, while for C2 and C3, optical gaps (allowed transition) values are significantly reduced. The excitation energies of allowed transition are decreasing monotonically from C1 to C3 with increased metal size (Li to K). The absorption maxima (λmax) of allowed transition are observed at longer wavelength as compared to absorption during crucial transition. As a result, bigger alkali metals have a stronger influence on absorptions shift to higher wavelengths. Furthermore, these clusters are completely transparent under the deep-UV region and have broadband absorption in the near-Visible region (Figure 4). The highest energy state TD-DFT parameters also reveal transparency in the deep-UV region, while absorption is mostly in the UV-visible region (Table 3). Likewise, the gradual increase in oscillator strength (fo) can also be seen for C1 to C3 clusters for crucial transition and allowed transition, which suggest increased quantum chemical excitation probabilities in higher-sized clusters.

2.7. Dipole Moment (µo) and Change in Dipole Moment (Δµ)

For better comprehension of the electronic properties in these clusters, the dipole moment (µo) and change in dipole moments (Δµ) values are also calculated. Overall, the dipole moment and change in dipole moment (Δµ) values are quite significant, which reveal asymmetric electronic distribution in these clusters (Table 4). The measured total dipole moment indicate polarity in clusters and the values of µo are significant and range from 1.49 to 4.11 au, where the highest value is observed for the C3 cluster. On the other hand, the total change in dipole moment (Δµ) values are slightly smaller than those of dipole moment, but C2 shows a significant value of 4.69 au (Table 4).

2.8. Linear and Nonlinear Optical (NLO) Properties

To investigate the influence of excess electrons on triggering the NLO properties of studied polynuclear clusters, hyperpolarizability (βo) and second hyperpolarizability (γo) are two crucial evaluation indices. The presence of excess electrons greatly increases the hyperpolarizability and second hyperpolarizability values, as shown in a number of studies [43,47,48,49,50,51,52,53,54]. We are interested in exploring the role of excess electrons in decreasing excitation energies, which ultimately escalates hyperpolarizabilities. The calculated linear and NLO parameters for the C3O[C(CN)2]2M3 (where M = Li, Na and K) at CAM-B3LYP/6-311++g(d,p) clusters are given in Table 4. The αo values lie in the range of 2.5 × 102 to 6.62 × 102, and there is a slight increase with the increased size of alkali metals. These values show liner optical properties of polynuclear clusters, and the presence of polarizabilities is due to asymmetric electronic density distribution in these clusters.
The hyperpolarizability values of studied clusters range from 2.37 × 103 to 5.78 × 103 au, where the highest value is obtained for C3, while the lowest value is for the C1 cluster. βo values are increasing from Li to K metals within these clusters, which shows size dependence. It can be seen that electronic properties significantly contribute to hyperpolarizability response, and the larger the change in dipole moment, the higher the hyperpolarizabilities are. Thus, βo values follow the increasing trend in these clusters, C1 < C2 < C3. Furthermore, the increased βo values have a good match with reduced ionization potential and HOMO–LUMO gaps. The trend of size-dependent βo is also shown in Figure 5.
In addition, the static second hyperpolarizability (γo) values are also calculated and lie in the range of 2.9 × 105 to 5.5 × 106 au (Table 4). Overall, γo values are significant where the highest value (5.5 × 106 au) is obtained for the C1 cluster, while the lowest is for C3. It is found that, with the increased size of alkali metals, the γo values decrease slightly from Li to Na and then dramatically for K. These values follow decreasing trend of γo values in order of C1 > C2 > C3. The calculated significant γo values guarantee the superior NLO properties of polynuclear clusters. The calculated values of βo and γo are quite significant as compared to previously reported M2OCN superalkalis [47], M2X (where M = Li, Na and X = F, Cl) superalkali clusters, and lithium-based superalkalis Lin (n = 3, 5, and 7) [55].
Furthermore, the βvec values are strongly correlated with total hyperpolarizability (βo). The calculated βvec values are given in Table 4. These values range from 2.37 × 103 to 5.78 × 103 au. The βvec is the projection of hyperpolarizability on dipole moment vector and shows close resemblance βo. However, good agreement between βo and βvec shows that the direction of the dipole moment vector and the projection of hyperpolarizability are in the same direction. The factor affecting βvec values might be the same for βo, where the highest βvec values are obtained for higher-sized alkali metals (Table 4).

2.9. Scattering Hyperpolarizability (βHRS)

Density functional theory calculations have been carried out to find scattering hyperpolarizability (βHRS), and values range from 1.34 × 103 to 1.62 × 104 au, where values are increasing steadily from the C1 to C3 cluster. The computed highest value is (1.62 × 104 au), found for C1 cluster, whereas the lowest value of 1.34 × 103 au is for the C1 cluster (Table 4). The βHRS is the most viable parameter to calculate the hyperpolarizability of centrosymmetric molecules and clusters, even with zero change in dipole moment. There is an excellent agreement of βHRS with βo where the βHRS show dependence on the size of alkali metals (M). The increased size of alkali metals (Li to K) favors the excellent electronic properties. Therefore, it also causes significantly enhanced βHRS values. Additionally, average dipolar and octupolar hyperpolarizability, which are more prominent in C2 and C3 clusters, provide a notable contribution to βHRS. Moreover, these clusters are of octupolar molecules, which can be seen by their highest octupolar contribution Φβ(j = 3) of 92 % for C3 (Table 4).

2.10. Frequency Dependent NLO Properties

We theoretically examined the incident–frequency (ω) effect on the first and second hyperpolarizability at applied frequencies of 532 and 1064 nm. The frequency-dependent first hyperpolarizability β(ω) is calculated with the electro–optical Pockel’s effect (EOPE) and second harmonic generation (SHG), whereas the γ(ω) is expressed in terms of dc-Kerr effect and second harmonic generation (SHG). Overall, the dynamic hyperpolarizabilities values are higher than those of static hyperpolarizabilities. The significant EOPE effect β(−ω; ω,0) was observed for the C3 cluster at 532 nm, while its SHG value increased up to 1.7 × 106 au (Table 5). It can be demonstrated that the dynamic hyperpolarizabilities are higher at the smaller incident frequency (ω = 532 nm) and slightly decreased at the higher dispersion frequency (1064 nm). Strikingly, the β(ω) values are much more pronounced for the EOPE effect at both frequencies.
Furthermore, the γ(ω) values are higher than γo, and the highest dc-Kerr value increased up to 2.6 × 109 au for C3 at 1064 nm (Table 6). The γ(ω) response becomes significant at the higher dispersion frequency (1064 nm), where SHG values are notable at both frequencies. The obtained higher values of the dc-Kerr effect reveal the nonlinear change in the refractive index of studied clusters. Hence, studied clusters have excellent NLO properties and can be used to design high-performance SHG devices.

3. Computational Details

All density functional theory (DFT) calculations are performed in the gas phase with Gaussian 09 software, whereas visualization is achieved using the GaussView 5.0 program [56,57]. Geometries of all polynuclear C3O[C(CN)2]2M3 (where M = Li, Na, and K) clusters are optimized at CAM-B3LYP/6-311++G(d,p) functionality [58]. The quantum mechanics-based Coulomb attenuating method (CAM-B3LYP) is a hybrid exchange-correlation functional that combines B3LYP’s hybrid features with the CAM functional’s long-range corrected parameter. It was found that this long-range corrected density functional substantially reduces the overestimation seen with conventional techniques and typically provides results that are comparable to those of coupled cluster calculations. Previous research has demonstrated that this method is well recognized for examining molecules and clusters, as well as for determining NLO properties [59,60]. Besides, the choice of a suitable basis set is crucial for obtaining reliable results. Thus, the CAM-B3LYP method with 6-311+G(d,p) split valence basis set is a reliable level of theory for geometry optimization and accuracy in results for electronic properties [61,62,63,64,65].
To determine whether the presented structures are true minima on the potential energy surface, frequency calculations are carried out. For thermodynamic stability, we calculated binding energy per atom for these clusters. Electronic stability and superalkali nature are validated through computed ionization energies (IE) and electron affinities (EA). To further explore the electronic properties, we performed frontier molecular orbital (FMO) analysis. Natural bonding orbitals (NBO) study is carried out to explore the charge distribution on atoms within superalkali clusters [66]. The binding energy per atom (EB) is given by the following relations:
E B = [ E T ( X ) E A ( X ) 0 ] / n  
where ET is the total electronic energy of studied (X) superalkali clusters, EA(X) is the total energy of individual atoms within clusters, and n is the total number of atoms. The vertical ionization energy, electron affinity, and electrical conductivity (σ) can be represented by the equation:
VIE = EX+ − EX0
VEA = EX0 − EX
σ     exp E G 2 kT
where VIE and VEA are vertical ionization energies and electron affinities of studied clusters. In Equation (4), σ, EG, k, and T represent the electrical conductivity, energy gap, Boltzmann constant, and temperature, respectively. To further explore the electronic properties of studied clusters, we also performed total density of state (TDOS) analysis at the same method by using the GaussSum software [67]. The following equation can be used to explain the molecules under the static electric field.
E ( F ) =   E 0   µ i F i 1 2 α ij F i F j 1 6 β ijk F i F j F k 1 24 γ ijkl F i F j F k F l  
where F is an external applied electric field, Fi is the component of field along i direction, E0 is the total energy of the superalkali clusters without a static electric field, and µi, αij, βijk, and γijkl are dipole moment, polarizability, hyperpolarizability, and second-order hyperpolarizability, respectively. The mean dipole moment (µo), change in dipole moment (Δµ), static polarizability (αo), and static first hyperpolarizability (βo) are calculated to illustrate the NLO response and associated responsible factors.
αo = 1/3 (αxx + αyy + αzz)
βo = (βx2 + βy2 + βz2)1/2
where βx = βxxx + βxyy + βxzz, βy = βyyy + βyzz + βyxx and βz = βzzz + βzxx + βzyy.
µo = (µx2 + µy2 + µz2) ½
To obtain absorption behaviors and excited state parameters of studied clusters, we performed TD-DFT simulations. We considered 30 states for getting excited states parameters. The Gaussian band shape and the absorption spectra were obtained by using the following relation, ϑ :
ε o ( ϑ ¯ ) = ε i m a x   exp   [ ( ϑ ¯ ϑ ¯ i σ ) 2 ]
where the i subscript represents the electronic excitation of interest. The other symbols in the equation have the following meanings:
  • ϑ ¯ i , shows the excitation energy (in wavenumbers) corresponding to the required electronic excitation in TD-DFT
  • ε i m a x is the value of at the maximum of the band shape
  • Sigma (σ) is a wavenumber representation of the standard deviation that is related to the simulated band’s width.
The second static hyperpolarizability (γo) and the projection of hyperpolarizability on the dipole moment vector (βvec) are also calculated for our studied superalkali clusters at the same level of theory. Static second hyperpolarizability (γo) and vector part of hyperpolarizability (βvec) are expressed as:
< γ > = 1/5 (γxxxx + γyyyy + γzzzz + γxxyy + γxxzz + γyyxx + γyyzz + γzzxx)
β vec = µ i β i | µ |
Moreover, the molecular parameters relevant to electro-optical Pockel’s effect (EOPE) and second harmonic generation (SHG) are calculated at externally applied frequencies (532 and 1064 nm).

4. Conclusions

In summary, we presented the geometric, electronic, and nonlinear optical properties of polynuclear carbon-based clusters at CAM-B3LYP/6-311++G(d,p) level. These clusters are thermodynamically stable, and their binding energies per atom range from −160.07 to −162.07 kcal mol−1. The electronic stability and superalkali nature are characterized through calculated ionization potential (IP) and FMO analyses. Small ionization potential further suggests their superalkali nature. NBO charge analysis reveals excellent charge separation within clusters. The performed DOS analysis shows size-dependent electronic and conductive properties, where C3 is a potential candidate. The significant first and second hyperpolarizabilities, up to 5.78 × 103 and 5.55 × 106 au, respectively, are calculated for the clusters. The βo response shows dependence on the size of alkali metals. Furthermore, the absorption study shows their transparency in the deep-UV region for NLO applications and absorption at longer wavelengths. The significant scattering hyperpolarizability (βHRS) value is (1.62 × 104), calculated for the C3 cluster, where octupolar contribution to βHRS is 92%. The dynamic first hyperpolarizability β(ω) is more pronounced for the EOPE effect at 532 nm, whereas SHG is more prominent for second hyperpolarizability γ(ω).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28041827/s1, Figure S1: DOS spectra of clusters; Table S1: Optimized geometries of clusters.

Author Contributions

I.B.: Methodology, Conceptualization, Funding acquisition; U.R.: Investigation, data handling; A.A.: Writing, Visualization; S.U.M.: Original draft preparation; N.S.S.: Software Validation, Editing; K.A.: Supervision and Reviewing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Scientific Research, Vice Presidency for Graduate Studies and Scientific Research, King Faisal University, Saudi Arabia [Grant No. 2879].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The author confirms that data supporting finding current study are available within article and in its supporting information. Raw data that supports the finding of this study are available from the corresponding author’s upon request.

Acknowledgments

This work was supported by the Deanship of Scientific Research, Vice Presidency for Graduate Studies and Scientific Research, King Faisal University, Saudi Arabia [Grant No. 2879]. Also, thanks to the Higher Education Commission, Pakistan for the support [Grant No. 2981] and Universiti Brunei Darussalam [UBD/RSCH/1.4/FICBF(b)/2022/049].

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Kanis, D.R.; Ratner, M.A.; Marks, T.J. Design and construction of molecular assemblies with large second-order optical nonlinearities. Quantum chemical aspects. Chem. Rev. 1994, 94, 195–242. [Google Scholar] [CrossRef]
  2. He, G.S.; Tan, L.-S.; Zheng, Q.; Prasad, P.N. Multiphoton Absorbing Materials: Molecular Designs, Characterizations, and Applications. Chem. Rev. 2008, 108, 1245–1330. [Google Scholar] [CrossRef]
  3. Dalton, L.R.; Steier, W.H.; Robinson, B.H.; Zhang, C.; Ren, A.; Garner, S.; Chen, A.; Londergan, T.; Irwin, L.; Carlson, B.; et al. From molecules to opto-chips: Organic electro-optic materials. J. Mater. Chem. 1999, 9, 1905–1920. [Google Scholar] [CrossRef]
  4. Bredas, J.L.; Adant, C.; Tackx, P.; Persoons, A.; Pierce, B.M. Third-Order Nonlinear Optical Response in Organic Materials: Theoretical and Experimental Aspects. Chem. Rev. 1994, 94, 243–278. [Google Scholar] [CrossRef]
  5. Kleinman, D.A. Nonlinear Dielectric Polarization in Optical Media. Phys. Rev. 1962, 126, 1977–1979. [Google Scholar] [CrossRef]
  6. Xiao, D.; Bulat, F.A.; Yang, W.; Beratan, D.N. A Donor−Nanotube Paradigm for Nonlinear Optical Materials. Nano Lett. 2008, 8, 2814–2818. [Google Scholar] [CrossRef]
  7. Palmer, S.; Sokolovski, S.G.; Rafailov, E.; Nabi, G. Technologic Developments in the Field of Photonics for the Detection of Urinary Bladder Cancer. Clin. Genitourin. Cancer 2013, 11, 390–396. [Google Scholar] [CrossRef]
  8. Lee, S.; Park, J.; Jeong, M.; Kim, H.; Li, S.; Song, J.; Ham, S.; Jeon, S.; Cho, B. First hyperpolarizabilities of 1, 3, 5-tricyanobenzene derivatives: Origin of larger β values for the octupoles than for the dipoles. ChemPhysChem 2006, 7, 206–212. [Google Scholar] [CrossRef]
  9. Noori, M.; Soroosh, M.; Baghban, H. Highly efficient self-collimation based waveguide for Mid-IR applications. Photonics Nanostructures Fundam. Appl. 2016, 19, 1–11. [Google Scholar] [CrossRef]
  10. Clough, B.; Dai, J.; Zhang, X.-C. Laser air photonics: Beyond the terahertz gap. Mater. Today 2012, 15, 50–58. [Google Scholar] [CrossRef]
  11. Xu, X.; Hu, C.-L.; Kong, F.; Zhang, J.-H.; Mao, J.-G.; Sun, J. ChemInform Abstract: Cs2GeB4O9: A New Second-Order Nonlinear-Optical Crystal. Inorg. Chem. 2013, 52, 5831–5837. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, C.-G.; Guan, W.; Song, P.; Yan, L.-K.; Su, Z.-M. Redox-Switchable Second-Order Nonlinear Optical Responses of Push−Pull Monotetrathiafulvalene-Metalloporphyrins. Inorg. Chem. 2009, 48, 6548–6554. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, C.; Wu, Y.; Li, R. The development of new NLO crystals in the borate series. J. Cryst. Growth 1990, 99, 790–798. [Google Scholar] [CrossRef]
  14. Kertész, M. Bond length alternation and energy gap in (CH)x. Application of the intermediate exciton formalism. Chem. Phys. 1979, 44, 349–356. [Google Scholar] [CrossRef]
  15. Arshad, Y.; Khan, S.; Hashmi, M.A.; Ayub, K. Transition metal doping: A new and effective approach for remarkably high nonlinear optical response in aluminum nitride nanocages. New J. Chem. 2018, 42, 6976–6989. [Google Scholar] [CrossRef]
  16. Liu, Z.-B.; Zhou, Z.-J.; Li, Y.; Li, Z.-R.; Wang, R.; Li, Q.-Z.; Li, Y.; Jia, F.-Y.; Wang, Y.-F.; Li, Z.-J.; et al. Push–pull electron effects of the complexant in a Li atom doped molecule with electride character: A new strategy to enhance the first hyperpolarizability. Phys. Chem. Chem. Phys. 2010, 12, 10562–10568. [Google Scholar] [CrossRef]
  17. Li, Z.; He, C.; Wang, Z.; Gao, Y.; Dong, Y.; Zhao, C.; Chen, Z.; Wu, Y.; Song, W. Ethylenediamine-modified graphene oxide covalently functionalized with a tetracarboxylic Zn(ii) phthalocyanine hybrid for enhanced nonlinear optical properties. Photochem. Photobiol. Sci. 2016, 15, 910–919. [Google Scholar] [CrossRef]
  18. Maury, O.; Le Bozec, H. Molecular Engineering of Octupolar NLO Molecules and Materials Based on Bipyridyl Metal Complexes. Accounts Chem. Res. 2005, 38, 691–704. [Google Scholar] [CrossRef]
  19. Hatua, K.; Nandi, P.K. Beryllium-Cyclobutadiene Multidecker Inverse Sandwiches: Electronic Structure and Second-Hyperpolarizability. J. Phys. Chem. A 2013, 117, 12581–12589. [Google Scholar] [CrossRef]
  20. Li, X.-H.; Zhang, L.; Zhang, X.-L.; Ni, B.-L.; Li, C.-Y.; Sun, W.-M. Designing a new class of excess electron compounds with unique electronic structures and extremely large non-linear optical responses. New J. Chem. 2020, 44, 6411–6419. [Google Scholar] [CrossRef]
  21. Zhong, R.-L.; Xu, H.-L.; Muhammad, S.; Zhang, J.; Su, Z.-M. The stability and nonlinear optical properties: Encapsulation of an excess electron compound LiCN⋯Li within boron nitride nanotubes. J. Mater. Chem. 2011, 22, 2196–2202. [Google Scholar] [CrossRef]
  22. Ahsan, A.; Ayub, K. Adamanzane based alkaline earthides with excellent nonlinear optical response and ultraviolet transparency. Opt. Laser Technol. 2020, 129, 106298. [Google Scholar] [CrossRef]
  23. Casida, M.E.; Jamorski, C.; Casida, K.C.; Salahub, D.R. Molecular excitation energies to high-lying bound states from time-dependent density-functional response theory: Characterization and correction of the time-dependent local density approximation ionization threshold. J. Chem. Phys. 1998, 108, 4439–4449. [Google Scholar] [CrossRef]
  24. Mulliken, R.S. Rydberg States and Rydbergization. Acc. Chem. Res. 1976, 9, 7–12. [Google Scholar] [CrossRef]
  25. Zhong, R.-L.; Xu, H.-L.; Li, Z.-R.; Su, Z.-M. Role of Excess Electrons in Nonlinear Optical Response. J. Phys. Chem. Lett. 2015, 6, 612–619. [Google Scholar] [CrossRef]
  26. Zhong, R.; Xu, H.; Sun, S.; Qiu, Y.; Su, Z. The Excess Electron in a Boron Nitride Nanotube: Pyramidal NBO Charge Distribution and Remarkable First Hyperpolarizability. Chem.—Eur. J. 2012, 18, 11350. [Google Scholar] [CrossRef]
  27. Chen, W.; Li, Z.-R.; Wu, D.; Li, Y.; Sun, C.-C.; Gu, F.L. The Structure and the Large Nonlinear Optical Properties of Li@Calix [4]pyrrole. J. Am. Chem. Soc. 2005, 127, 10977–10981. [Google Scholar] [CrossRef]
  28. Ma, F.; Li, Z.; Xu, H.; Li, Z.; Li, Z.; Aoki, Y.; Gu, F. Lithium salt electride with an excess electron pair—A class of nonlinear optical molecules for extraordinary first hyperpolarizability. J. Phys. Chem. A 2008, 112, 11462–11467. [Google Scholar] [CrossRef] [PubMed]
  29. Li, B.; Peng, D.; Gu, F.L.; Zhu, C. A nonlinear optical switch induced by an external electric field: Inorganic alkaline–earth alkalide. RSC Adv. 2019, 9, 16718–16728. [Google Scholar] [CrossRef] [Green Version]
  30. Ahsan, A.; Ayub, K. Extremely large nonlinear optical response and excellent electronic stability of true alkaline earthides based on hexaammine complexant. J. Mol. Liq. 2019, 297, 111899. [Google Scholar] [CrossRef]
  31. Ayub, K.; Ahsan, F. Transition metalides based on facially polarized all-cis-1, 2, 3, 4, 5, 6-hexafluorocyclohexane–a new class of high performance second order nonlinear optical materials. Phys. Chem. Chem. Phys. 2023, 25, 4732–4737. [Google Scholar]
  32. Sun, W.-M.; Li, X.-H.; Wu, D.; Li, Y.; He, H.-M.; Li, Z.-R.; Chen, J.-H.; Li, C.-Y. A theoretical study on superalkali-doped nanocages: Unique inorganic electrides with high stability, deep-ultraviolet transparency, and a considerable nonlinear optical response. Dalton Trans. 2016, 45, 7500–7509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Kim, J.; Ichimura, A.; Huang, R.; Redko, M.; Phillips, R.; Jackson, J.; Dye, J. Crystalline Salts of Na- and K- (Alkalides) that Are Stable at Room Temperature. J. Am. Chem. Soc. 1999, 121, 10666–10667. [Google Scholar] [CrossRef]
  34. Ichimura, A.S.; Dye, J.L.; Camblor, M.A.; Villaescusa, L.A. Toward Inorganic Electrides. J. Am. Chem. Soc. 2002, 124, 1170–1171. [Google Scholar] [CrossRef] [PubMed]
  35. Hou, J.; Jiang, D.; Qin, J.; Duan, Q. Alkaline-earthide: A new class of excess electron compounds Li-C6H6F6-M (M = Be, Mg and Ca ) with extremely large nonlinear optical responses. Chem. Phys. Lett. 2018, 711, 55–59. [Google Scholar] [CrossRef]
  36. Sun, W.-M.; Li, X.-H.; Wu, J.; Lan, J.-M.; Li, C.-Y.; Wu, D.; Li, Y.; Li, Z.-R. Can Coinage Metal Atoms Be Capable of Serving as an Excess Electron Source of Alkalides with Considerable Nonlinear Optical Responses? Inorg. Chem. 2017, 56, 4594–4600. [Google Scholar] [CrossRef] [PubMed]
  37. Gutsev, G.; Boldyrev, A. DVM Xα calculations on the electronic structure of “superalkali” cations. Chem. Phys. Lett. 1982, 92, 262–266. [Google Scholar] [CrossRef]
  38. Sikorska, C.; Gaston, N. N4Mg6M (M = Li, Na, K) superalkalis for CO2 activation. J. Chem. Phys. 2020, 153, 144301. [Google Scholar] [CrossRef]
  39. Saedi, L.; Dodangi, M.; Mohammadpanaardakan, A.; Eghtedari, M. Superalkali–Superhalogen Complexes as Versatile Materials for Hydrogen Storage: A Theoretical Study. J. Clust. Sci. 2019, 31, 71–78. [Google Scholar] [CrossRef]
  40. Park, H.; Meloni, G. Activation of Dinitrogen with a Superalkali Species, Li3 F2. ChemPhysChem 2018, 19, 256–260. [Google Scholar] [CrossRef]
  41. Giri, S.; Behera, S.; Jena, P. Superalkalis and Superhalogens as Building Blocks of Supersalts. J. Phys. Chem. A 2014, 118, 638–645. [Google Scholar] [CrossRef] [PubMed]
  42. Percec, V.; Won, B.C.; Peterca, M.; Heiney, P.A. Expanding the Structural Diversity of Self-Assembling Dendrons and Supramolecular Dendrimers via Complex Building Blocks. J. Am. Chem. Soc. 2007, 129, 11265–11278. [Google Scholar] [CrossRef] [PubMed]
  43. Ullah, F.; Kosar, N.; Ayub, K.; Mahmood, T. Superalkalis as a source of diffuse excess electrons in newly designed inorganic electrides with remarkable nonlinear response and deep ultraviolet transparency: A DFT study. Appl. Surf. Sci. 2019, 483, 1118–1128. [Google Scholar] [CrossRef]
  44. Sajid, H.; Ayub, K.; Mahmood, T. Exceptionally high NLO response and deep ultraviolet transparency of superalkali doped macrocyclic oligofuran rings. New J. Chem. 2020, 44, 2609–2618. [Google Scholar] [CrossRef]
  45. Srivastava, A.; Misra, N. Nonlinear optical behavior of LinF (n = 2–5) superalkali clusters. J. Mol. Model. 2015, 21, 1–5. [Google Scholar] [CrossRef] [PubMed]
  46. Srivastava, A.K.; Misra, N. M2X (M = Li, Na; X = F, Cl): The smallest superalkali clusters with significant NLO responses and electride characteristics. Mol. Simul. 2016, 42, 981–985. [Google Scholar] [CrossRef]
  47. Ahsin, A.; Ayub, K. Theoretical investigation of superalkali clusters M2OCN and M2NCO (where M = Li, Na, K) as excess electron system with significant static and dynamic nonlinear optical response. Optik 2020, 227, 166037. [Google Scholar] [CrossRef]
  48. Ahsin, A.; Ayub, K. Extremely large static and dynamic nonlinear optical response of small superalkali clusters NM3M’(M, M’= Li, Na, K). J. Mol. Graph. Model. 2021, 109, 108031. [Google Scholar] [CrossRef]
  49. Li, X.-H.; Zhang, X.-L.; Chen, Q.-H.; Zhang, L.; Chen, J.-H.; Wu, D.; Sun, W.-M.; Li, Z.-R. Coinage metalides: A new class of excess electron compounds with high stability and large nonlinear optical responses. Phys. Chem. Chem. Phys. 2020, 22, 8476–8484. [Google Scholar] [CrossRef]
  50. Ahsin, A.; Ali, A.; Ayub, K. Alkaline earth metals serving as source of excess electron for alkaline earth metals to impart large second and third order nonlinear optical response; a DFT study. J. Mol. Graph. Model. 2020, 101, 107759. [Google Scholar] [CrossRef] [PubMed]
  51. Ahsin, A.; Ayub, K. Remarkable electronic and NLO properties of bimetallic superalkali clusters: A DFT study. J. Nanostructure Chem. 2021, 12, 529–545. [Google Scholar] [CrossRef]
  52. Ahsin, A.; Ayub, K. Zintl based superatom P7M2 (M = Li, Na, K & Be, Mg, Ca) clusters with excellent second and third-order nonlinear optical response. Mater. Sci. Semicond. Process. 2021, 134, 105986. [Google Scholar]
  53. Ahsin, A.; Ayub, K. Superalkali-based alkalides Li3O@[12-crown-4]M (where M = Li, Na, and K) with remarkable static and dynamic NLO properties; A DFT study. Mater. Sci. Semicond. Process. 2021, 138, 106254. [Google Scholar] [CrossRef]
  54. Ahsin, A.; Ayub, K. Oxacarbon superalkali C3X3Y3 (X = O, S and Y = Li, Na, K) clusters as excess electron compounds for remarkable static and dynamic NLO response. J. Mol. Graph. Model. 2021, 106, 107922. [Google Scholar] [CrossRef]
  55. Ahsin, A.; Ayub, K. Theoretical investigation of lithium-based clusters Lin (where n = 3, 5, 7) with remarkable electronic and frequency-dependent NLO properties. Eur. Phys. J. Plus 2022, 137, 803. [Google Scholar] [CrossRef]
  56. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  57. Dennington, R.; Keith, T.A.; Millam, J.M. GaussView, 6th ed.; Semichem Inc.: Shawnee Mission, KS, USA, 2009. [Google Scholar]
  58. Grein, F. Influence of diffuse and polarization functions on the second-order Møller–Plesset optimized dihedral angle of biphenyl. Theor. Chem. Accounts 2003, 109, 274–277. [Google Scholar] [CrossRef]
  59. Okuno, K.; Shigeta, Y.; Kishi, R.; Miyasaka, H.; Nakano, M. Tuned CAM-B3LYP functional in the time-dependent density functional theory scheme for excitation energies and properties of diarylethene derivatives. J. Photochem. Photobiol. A Chem. 2012, 235, 29–34. [Google Scholar] [CrossRef]
  60. Kobayashi, R.; Amos, R.D. The application of CAM-B3LYP to the charge-transfer band problem of the zincbacteriochlorin–bacteriochlorin complex. Chem. Phys. Lett. 2006, 420, 106–109. [Google Scholar] [CrossRef]
  61. Bravo-Pérez, G.; Garzón, I.; Novaro, O. Ab initio study of small gold clusters. J. Mol. Struct. THEOCHEM 1999, 493, 225–231. [Google Scholar] [CrossRef]
  62. Marchetti, O.; Werner, H.-J. Accurate Calculations of Intermolecular Interaction Energies Using Explicitly Correlated Coupled Cluster Wave Functions and a Dispersion-Weighted MP2 Method. J. Phys. Chem. A 2009, 113, 11580–11585. [Google Scholar] [CrossRef] [PubMed]
  63. Castet, F.; Rodriguez, V.; Pozzo, J.-L.; Ducasse, L.; Plaquet, A.; Champagne, B. Design and Characterization of Molecular Nonlinear Optical Switches. Accounts Chem. Res. 2013, 46, 2656–2665. [Google Scholar] [CrossRef] [PubMed]
  64. Maroulis, G. Quantifying the performance of conventional DFT methods on a class of difficult problems: The interaction (hyper)polarizability of two water molecules as a test case. Int. J. Quantum Chem. 2011, 112, 2231–2241. [Google Scholar] [CrossRef]
  65. Karamanis, P.; Maroulis, G.; Pouchan, C. Molecular geometry and polarizability of small cadmium selenide clusters from all-electron ab initio and Density Functional Theory calculations. J. Chem. Phys. 2006, 124, 71101. [Google Scholar] [CrossRef] [PubMed]
  66. Reed, A.E.; Weinstock, R.B.; Weinhold, F. Natural population analysis. J. Chem. Phys. 1985, 83, 735–746. [Google Scholar] [CrossRef]
  67. Allouche, A.-R. Gabedit-A graphical user interface for computational chemistry softwares. J. Comput. Chem. 2010, 32, 174–182. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Optimized structure of Clusters C1 to C3.
Figure 1. Optimized structure of Clusters C1 to C3.
Molecules 28 01827 g001
Figure 2. Binding energies per atom (Eb) of clusters.
Figure 2. Binding energies per atom (Eb) of clusters.
Molecules 28 01827 g002
Figure 3. Generated HOMO and LUMO densities of clusters.
Figure 3. Generated HOMO and LUMO densities of clusters.
Molecules 28 01827 g003
Figure 4. Absorption spectra of C1 to C3 clusters.
Figure 4. Absorption spectra of C1 to C3 clusters.
Molecules 28 01827 g004
Figure 5. The representation of size-dependent hyperpolarizability (βo).
Figure 5. The representation of size-dependent hyperpolarizability (βo).
Molecules 28 01827 g005
Table 1. Computed bond length between metal and O-atom (dM-O in Å), the bond length between alkali metals and N-atom (dM-N in Å), binding energies per atom (Eb in kcal mol−1), NBO charges on metals (QM), NBO charge on nitrogen (QN), NBO charge on oxygen atom (QO), VIE (in eV), and VEA (in eV), of C1 to C3 clusters.
Table 1. Computed bond length between metal and O-atom (dM-O in Å), the bond length between alkali metals and N-atom (dM-N in Å), binding energies per atom (Eb in kcal mol−1), NBO charges on metals (QM), NBO charge on nitrogen (QN), NBO charge on oxygen atom (QO), VIE (in eV), and VEA (in eV), of C1 to C3 clusters.
ClusterdM-OdM-NEbQ(M)Q(N)Q(O)VIEVEA
C11.821.89−162.40.58−0.485−0.9603.650.76
C22.242.26−160.10.61−0.521−0.8393.410.89
C32.582.62−162.10.62−0.523−0.8303.000.27
Table 2. Energies of HOMO (EHOMO in eV), LUMO energies (in eV), HOMO-LUMO gaps (EH-L in eV), chemical hardness (η in eV), chemical, softness (S in eV), chemical potential (χ in eV), oscillator strength (fo in au), excitation energies (eV), and maximum absorption (in nm) of C1 to C3 clusters.
Table 2. Energies of HOMO (EHOMO in eV), LUMO energies (in eV), HOMO-LUMO gaps (EH-L in eV), chemical hardness (η in eV), chemical, softness (S in eV), chemical potential (χ in eV), oscillator strength (fo in au), excitation energies (eV), and maximum absorption (in nm) of C1 to C3 clusters.
ClusterEHOMOELUMOEH-LηSχ
C1−4.70−0.614.081.8390.27−1.81
C2−3.24−0.882.351.7210.29−1.72
C3−2.80−0.831.961.5050.33−1.50
Table 3. TD-DFT parameters of crucial excited sates, first allowed transitions, and highest state for C1 to C3 clusters.
Table 3. TD-DFT parameters of crucial excited sates, first allowed transitions, and highest state for C1 to C3 clusters.
ClustersTD-DFT Parameters from Crucial Transitions
ΔE (eV)λmax (nm)fo (au)Major Orbital Contribution
C11.637580.19HOMO→LUMO+2 (82%)
C21.806880.26HOMO→LUMO+3 (36%)
C31.249950.28HOMO→LUMO+5 (67%)
TD-DFT Parameters from First Allowed Transitions
C11.637580.19HOMO→LUMO+2 (82%)
C20.9213380.23HOMO→LUMO+1 (99%)
C30.8614410.25HOMO→LUMO+1 (96%)
TD-DFT Parameters for Highest Energy States
C15.122420.0018
C24.252910.0005
C33.923150.0002
Table 4. Polarizabilities (αo in au), hyperpolarizabilities (βo in au), second hyperpolarizability (γo in au), scattering hyperpolarizability (βHRS in au), vector part of hyperpolarizability (βvec in au), average hyperpolarizability (<βJ=1> in au), average octupolar hyperpolarizability (<βJ=3> in au), % dipolar contribution to hyperpolarizability Φβ(j = 1), and % octupolar contribution to hyperpolarizability Φβ(j = 3) of C1 to C3 clusters.
Table 4. Polarizabilities (αo in au), hyperpolarizabilities (βo in au), second hyperpolarizability (γo in au), scattering hyperpolarizability (βHRS in au), vector part of hyperpolarizability (βvec in au), average hyperpolarizability (<βJ=1> in au), average octupolar hyperpolarizability (<βJ=3> in au), % dipolar contribution to hyperpolarizability Φβ(j = 1), and % octupolar contribution to hyperpolarizability Φβ(j = 3) of C1 to C3 clusters.
Clusters αo βo γo βHRS βvecJ=1> J=3> Φβ(j = 1)Φβ(j = 3)
C12.5 × 1022.37 × 1035.5 × 1061.34 × 1032.37 × 1031.8 × 1033.37 × 10335%65%
C25.07 × 1024.46 × 1031.2 × 1064.87 × 1034.46 × 1033.28 × 1031.49 × 10418%82%
C36.62 × 1025.78 × 1032.9 × 1051.62 × 1045.78 × 1034.30 × 1035.20 × 10308%92%
Table 5. Hyperpolarizability (β0 in au), frequency-dependent hyperpolarizability β(ω) in terms of electro-optic-Pockel’s effect (EOPE) β (−ω; ω, 0) in au, and electric field induced second harmonic generation (EFSHG) β (−2ω; ω, ω) in au at ω = 532 au.
Table 5. Hyperpolarizability (β0 in au), frequency-dependent hyperpolarizability β(ω) in terms of electro-optic-Pockel’s effect (EOPE) β (−ω; ω, 0) in au, and electric field induced second harmonic generation (EFSHG) β (−2ω; ω, ω) in au at ω = 532 au.
Clusterω = 0ω = 532 nmω = 1064 nm
β (0;0,0)β (−ω; ω,0)β (2-ω;ω,ω)β (−ω; ω,0)β (−2ω; ω, ω)
C12.5 × 1028.1 ×1032.2 × 1058.1 × 1052.9 × 105
C25.0 × 1021.0 × 1054.2 × 1052.7 × 1061.6 × 105
C36.6 × 1021.2 × 1071.7 × 1065.0 × 1054.5 × 103
Table 6. Static second hyperpolarizability (γo in au), frequency-dependent second-hyperpolarizability γ(ω) in term of electro-optic-pockel’s effect (EOPE) γ (−ω; ω, 0) in au, and electric field-induced second harmonic generation (efshg) γ (2-ω; ω, ω) in au at ω = 532 au.
Table 6. Static second hyperpolarizability (γo in au), frequency-dependent second-hyperpolarizability γ(ω) in term of electro-optic-pockel’s effect (EOPE) γ (−ω; ω, 0) in au, and electric field-induced second harmonic generation (efshg) γ (2-ω; ω, ω) in au at ω = 532 au.
Clustersω = 0ω = 532 nmω = 1064 nm
γ (0;0,0,0)γ (−ω; ω,0,0)γ (−2ω;ω,ω, ω)γ (−ω; ω,0,0)γ (−2ω; ω, ω, ω)
C15.5 × 1063.0 × 1086.0 × 1071.0 × 1061.2 × 108
C21.2 × 1062.4 × 1082.3 × 1072.3 × 1075.0 × 107
C32.9 × 1054.9 × 1071.7 × 1082.6 × 1092.0 × 109
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bayach, I.; Ahsin, A.; Majid, S.U.; Rashid, U.; Sheikh, N.S.; Ayub, K. Geometric, Electronic, and Optoelectronic Properties of Carbon-Based Polynuclear C3O[C(CN)2]2M3 (where M = Li, Na, and K) Clusters: A DFT Study. Molecules 2023, 28, 1827. https://doi.org/10.3390/molecules28041827

AMA Style

Bayach I, Ahsin A, Majid SU, Rashid U, Sheikh NS, Ayub K. Geometric, Electronic, and Optoelectronic Properties of Carbon-Based Polynuclear C3O[C(CN)2]2M3 (where M = Li, Na, and K) Clusters: A DFT Study. Molecules. 2023; 28(4):1827. https://doi.org/10.3390/molecules28041827

Chicago/Turabian Style

Bayach, Imene, Atazaz Ahsin, Safi Ullah Majid, Umer Rashid, Nadeem S. Sheikh, and Khurshid Ayub. 2023. "Geometric, Electronic, and Optoelectronic Properties of Carbon-Based Polynuclear C3O[C(CN)2]2M3 (where M = Li, Na, and K) Clusters: A DFT Study" Molecules 28, no. 4: 1827. https://doi.org/10.3390/molecules28041827

Article Metrics

Back to TopTop