Next Article in Journal
Chemical Characterization and Sensory Evaluation of Scottish Malt Spirit Aged in Sherry Casks®: Comparison Between Static and Dynamic Aging Systems
Next Article in Special Issue
Hydroperoxyl Radical Scavenging Activity of Bromophenols from Marine Red Alga Polysiphonia urceolata: Mechanistic Insights, Kinetic Analysis, and Influence of Physiological Media
Previous Article in Journal
Novel Fe(II)-Based Supramolecular Film Prepared by Interfacial Self-Assembly of an Asymmetric Polypyridine Ligand and Its Electrochromic Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural, Electronic, and Nonlinear Optical Characteristics of Europium-Doped Germanium Anion Nanocluster EuGen (n = 7–20): A Theoretical Investigation

1
Inner Mongolia Key Laboratory of Theoretical and Computational Chemistry Simulation, School of Chemical Engineering, Inner Mongolia University of Technology, Hohhot 010051, China
2
School of Resources and Environmental Engineering, Inner Mongolia University of Technology, Hohhot 010051, China
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(6), 1377; https://doi.org/10.3390/molecules30061377
Submission received: 25 February 2025 / Revised: 14 March 2025 / Accepted: 16 March 2025 / Published: 19 March 2025

Abstract

:
Doping rare-earth metals into semiconductor germanium clusters can significantly enhance the stability of these clusters while introducing novel and noteworthy optical properties. Herein, a series of EuGen (n = 7–20) clusters and their structural and nonlinear optical properties are investigated via the ABCluster global search technique combined with the double-hybrid density functional theory mPW2PLYP. The structure growth pattern can be divided into two stages: an adsorption structure and a linked structure (when n = 7–10 and n = 11–20, respectively). In addition to simulating the photoelectron spectra of the clusters, their various properties, including their (hyper)polarizability, magnetism, charge transfer, relative stability, and energy gap, are identified. According to our examination, the EuGe13 cluster exhibits a significant nonlinear optical response of the βtot value of 7.47 × 105 a.u., and is thus considered a promising candidate for outstanding nonlinear optical semiconductor nanomaterials.

1. Introduction

Nanomaterials have emerged as key materials with multidisciplinary potential in various fields, like energy and optics, driven by their unique structure–property relationships [1,2]. In the field of advanced semiconductors, germanium and germanium-based nanomaterials play an essential role [3]. Since 1958—when it was used to develop the world’s first integrated circuit—germanium has gradually been used in various advanced manufacturing fields [4]. For example, germanium has high transparency towards infrared light, allowing it to be used for fabricating infrared lenses and windows [5], narrow-emitting quantum dots for high-color-index displays [6], non-silicon solid-state transistors with high-mobility charge carriers [7], solar cells for high photoelectric conversion efficiency [8], and so on. As next-generation nanomaterials, germanium and germanium-based semiconductors are gradually becoming irreplaceable in cutting-edge technologies and high-precision manufacturing. Additionally, more and more unique physical and chemical properties await further discovery, so it is urgent to develop novel germanium-based semiconductor materials to further satisfy current industrial and societal needs.
Pure germanium nanoclusters provide excellent semiconductor performance and are already widely applied in mainstream advanced optoelectronic devices. However, pure germanium has a large number of dangling bonds, leading to a highly sensitive device service environment [9,10]. To overcome this situation and prolong the devices’ lifetimes, one effective and easy-to-implement strategy is doping transitional metal (TM) into pure germanium nanoclusters. There are several examples of this in the literature. Manganese-doped anionic germanium nanoclusters MnGe6 and Mn2Ge7 were studied by means of joint photoelectron spectroscopy (PES) with theoretical calculations, and the results showed that MnGe6 and Mn2Ge7 have the same symmetry of C2v and vertical electron detachment energies of 2.58 ± 0.08 and 2.88 ± 0.08 eV, respectively [11]. A series of transitional and lanthanide metals, including Sc, Ti, V, Y, Zr, Nb, Lu, Hf, and Ta, were examined by means of anion PES at a 213 nm pulse using a Nd3+:YAG laser, which revealed a Ge16 cage structure with a large cavity, allowing the metal to be encapsulated [12]. In a theoretical investigation, bimetallic Mo2Gen (n = 9–15) clusters were studied via the B3LYP density functional approach, and the Mo2Ge9 theoretical predicted structure was analogous to the one that was experimentally observed [13]. Group 4 and 5 elements in germanium clusters TMGe8–17 were studied by applying a genetic algorithm coupled with density functional theory. The calculated PES highly agreed with the measured spectra, in which the complete closed cage motif showed great stability [14]. Monoanionic Ru2Gen (n = 3–20) was examined by using a genetic algorithm associated with PBE functionality. The ground-state structure study showed that, when extra electrons were added in, the cluster structure was significantly influenced [15]. Regarding noble metal-doped germanium nanoclusters, AuGe12 clusters with a bicapped pentagonal prism geometry were investigated via density functional theory calculations. The results showed that the hybridization between the transitional metal Au and the germanium cage enhanced the chemical stability of the whole cluster [16]. Such conclusions were also made for IrGen (n = 1–20) clusters [17]. Because of the similar chemical and physical properties of silicon and germanium, Kumar and Kawazoe, using the ab initio pseudopotential plane wave method, calculated the structures of transitional metal Ti-, Zr-, and Hf-doped Ge16, forming the Frank–Kasper structure MGe16 and capped decahedral or cubic MGe16 and MGe14; moreover, they found that the growth behaviors were different from those for the MSin nanocluster [18]. Another research study by Han and Hagelberg proved that TMSin with n = 12 or n = 16 and TMGen with n = 10, n = 12, or n = 16 are the most promising candidates for modular nanomaterials [19]. Besides doping with d-block transition metals, another way to achieve such goals is through doping with rare-earth metals. A scandium atom doped into an anion germanium cluster can vastly enhance the stability of the cluster, owing to the form of high-symmetry Frank–Kasper ScGe16 with 68 valence electron-filled shell configurations [20,21,22,23]. The different cage types of ScGe12 were also theoretically studied [24]. Li et al. systematically explored the structural evolution, relative stability, and electronic properties of LaGen (n = 3–14), and their results indicated that a lanthanum atom doped into a pure germanium cluster increased the interactions between germanium atoms and enhanced the stability of the whole system; such building blocks have potential for the design of molecular chains [25]. A series of germanium anionic clusters ReGe6 doped with the heavy rare-earth elements Gd, Tb, Dy, Ho, Er, Tm, Yb, and Lu were identified using the PBE0 functional method combined with the def2-TZVP basis set. The calculations showed that rare-earth atoms doped into center pentagonal ring structures have greater stability and aromaticity than other types, making them potential building blocks for novel rare-earth-based semiconductor materials [26]. The lowest energy structures of rare-earth element Yb-, Pm-, and Dy-doped neutral and cationic germanium clusters GenM (n = 9, 10; M = Yb, Pm, and Dy) were determined using a genetic algorithm coupled with the PBE functional and projected augmented wave method. The results showed that rare-earth metal doping into germanium is favorable in terms of the energy direction [27]. With the great number of potential applications in the current semiconductor industry and the huge demand in today’s society, there is an urgent need to discover novel, stable, and ultra-efficient nanomaterials.
The element europium (Eu) is a typical and unique luminescent activator, widely used in a variety of luminescent materials. Depending on the coordination environment and electron transition of its 4f-4f and 4f-5d orbitals, it can exhibit different optical properties, and there is high demand for its application in different situations. In view of this, we selected Eu as a dopant in pure germanium clusters to design novel semiconductor building blocks. The ground-state structures of the EuGen (n = 7–20) clusters were systematically determined through stepwise optimization, and their nonlinear optical properties were investigated. Moreover, their photoelectron spectra, magnetic behavior, relative stability, energy gap, and density of states were all studied. The results reveal the large polarizability and hyperpolarizability of the EuGe13 clusters, which may serve as potential nonlinear optical materials for the next-generation semiconductor industry.

2. Results and Discussion

2.1. Ground-State Structure and Evolution Pattern of EuGen (n = 7–20) Nanoclusters

The ground-state structures, symmetry, and energy differences in EuGen (n = 7–20) optimized at the mPW2PLYP level are shown in Figure 1, and the Cartesian coordinates of each cluster are collected in Table S1. To further ensure the ground-state structures, the AIMD results are shown in Figure S1 (Supporting Information), and the total energies at mPW2PLYP with zero-point energy at TPSSh are collected in Table S2. Each cluster is noted as nA-x, where n is the number of Ge atoms, A is the anionic cluster, and x is the sequence number from small to large for increasing energy differences compared to the energy of the ground-state structure. The spin multiplicities of EuGen (n = 7–20) clusters were all predicted for nonets. For EuGe7, 7A-1 and 7A-2 have the same symmetry of Cs, which can be seen as one Eu atom replacing one Ge atom in the pure Ge8 cluster [28]. The structure of 7A-3 is two Ge atoms forming a capped tetragonal bi-pyramid, of which the Eu atom is the vertex. The energies of 7A-2 and 7A-3 are higher than that of 7A-1, by 0.29 and 0.33 eV, respectively. For EuGe8, 8A-1 is one Eu atom adsorbed on the top of the square frustum of Ge8, and 8A-2 is a typical tricapped triangular prism (TTP) structure where the Eu atom is located on one side of the triangular prism. The structure of 8A-3 can be described as one extra Ge atom adsorbed onto 7A-3. The energy differences for 8A-2 and 8A-3 with regard to 8A-1 are 0.32 and 0.49 eV, respectively. For EuGe9, all isomers have a similar structure, depicted as one Eu atom and two Ge atoms capped on the TTP Ge6 basic framework, where the difference is only in the three atoms’ cap location. 9A-1 is more stable than 9A-2 and 9A-3, with energy differences of 0.37 and 0.76 eV, respectively. For EuGe10, 10A-1 and 10A-2 are similar types, with one Eu atom site on the side of two capped anti-prisms Ge10. Due to the similarity of their structures, the energies of 10A-1 and 10A-2 are close, with a difference of only 0.07 eV. 10A-3 is derived by adding one Ge atom to the ground-state structure of EuGe9. The energy of 10A-3 is higher than that of 10A-1 by about 0.67 eV. For EuGe11, a linked structure emerges. The isomers of 11A-1 and 11A-2 are all depicted with Eu atoms as linkers between two parts of TTP Ge9 and Ge2. The structure of 11A-3 is one Eu atom and one Ge atom on the plane of Cs symmetry, co-connected to two trigonal bi-pyramids Ge5. 11A-4 can be seen as one extra Ge atom added onto the side of 10A-1. The energy differences for 11A-2, 11A-3, and 11A-4 with regard to 11A-1 are 0.20, 0.27, and 0.39 eV, respectively. For EuGe12, the lowest-energy 12A-1 is a linked type with a Eu atom connected to Ge3 as one part and one capped anti-prism Ge9 as the other part. The structure of 12A-2 exhibits a center Eu atom joining two tetragonal bi-pyramids Ge6. 12A-1 is more stable than 12A-2, with an energy difference of 0.18 eV. For EuGe13, the minimum-energy structure is a Eu atom associated with TTP Ge9 and a four-membered ring Ge4. 13A-2 and 13A-3 are similar motifs viewed as a four-membered ring Ge4 sub-cluster adsorbed onto two capped anti-prism basic frameworks; their energies are higher than that of the 13A-1 linked structure by 0.21 and 0.31 eV, respectively. For EuGe14, all selected isomers are of linked forms. 14A-1, with Cs symmetry, consists of a Eu atom connected to one unit of TTP Ge9 and one unit of trigonal bi-pyramid Ge5. 14A-2 is an energy-degenerate form of 14A-1 showing one sub-structure of TTP Ge9 and another of square-pyramid Ge5. The structure of 14A-3 shows mirror symmetry with that of 14A-2, but with a slight difference. The energy difference between 14A-1 and 14A-2 is only 0.04 eV, and that between 14A-1 and 14A-3 is 0.16 eV. For EuGe15, the lowest-energy structure of 15A-1 with C2V symmetry can be divided into one Eu atom combined with three Ge atoms forming a co-planar quaternary ring that connects two mirror-symmetric pentagonal pyramids Ge6. 15A-2 and 15A-3 can both be described as a Eu atom binding one sub-unit of TTP Ge9 and one sub-unit of Ge6. The energies of 15A-2 and 15A-3 are higher than that of 15A-1 by 0.22 and 0.48 eV. For EuGe16, the architectures of 16A-1, 16A-2, and 16A-3 can all be noted as a Eu atom as a connector joining one sub-cluster of TTP Ge9 and another sub-cluster of Ge7. The energy of 16A-1 is lower than those of 16A-2 and 16A-3 by 0.34 and 0.46 eV, respectively. For EuGe17, the skeleton of 17A-1 can be described as a center Eu atom bridging one capped anti-prism Ge9 and one anti-prism Ge8. 17A-2 is of a chain type with two parts of two Ge atoms capping an anti-prism Ge10 and two Ge atoms capping a tetragonal bi-pyramid where the Eu atom is not located at the middle position as a linker. The energy of 17A-2 is 0.26 eV higher than that of 17A-1. The cage structure of 17A-3 has a larger energy difference relative to the linked type of 17A-1 of about 1.02 eV. For EuGe18, 18A-1 and 18A-2 are analogous structures portrayed as a Eu atom anchoring two TTP Ge9 sub-clusters. 18A-3 is an endohedral cage structure where a Eu atom is encapsulated in a Ge18 cage skeleton. The energies of 18A-2 and 18A-3 are 0.16 and 1.54 eV higher than the energy of 18A-1. For EuGe19, 19A-1 and 19A-2, with the same Cs symmetry, can be viewed as a Eu atom bridging one Ge atom capping a Ge9 anti-prism on one side and two Ge atoms capping a Ge10 anti-prism on the other. The energies of 19A-1 and 19A-2 are degenerative with a small energy difference of 0.04 eV. The 19A-3 cage structure originates from the cage motif of 17A-3 with the addition of two extra Ge atoms into the Ge bones. The energy of 19A-3 is 1.67 eV larger than that of 19A-1. For EuGe20, the energy-minimum pattern 20A-1 can be deconstructed as a middle triangle composed of one Eu and two Ge atoms that connects symmetrical bi-lateral TTP Ge9 blocks. The geometry of 20A-2 is a Eu atom located at the center, bridging two capped anti-prisms. The energies of 20A-2 and the cage structure 20A-3 are 0.24 and 1.06 eV larger than that of 20A-1.
To sum up, the structural evolution of EuGen (n = 7–20) can be classified into two stages: (i) for small-sized clusters with n = 7–10, the configurations mainly present an adsorbed mode, where a Eu atom is adsorbed onto the Gen cluster; (ii) for medium-sized clusters with n = 11–20, the structures exhibit a pronounced tendency to form elongated link structures. Specifically, Eu atoms serve as connection points, bridging sub-clusters with TTP structures of Ge9 and Gen-9. Consequently, the EuGe9 sub-cluster, which is formed via linkage between a Eu atom and a TTP-structured Ge9 unit, constitutes the fundamental structural unit for the growth of medium-sized clusters.

2.2. Predicted Photoelectron Spectroscopy of EuGen (n = 7–20)

Photoelectron spectroscopy (PES) is a common technique for indirectly testing the geometry and electronic structure of nanoclusters. Because nanoclusters have multiple isomers and states that are difficult to accurately determine, PES characterization of nanoclusters is absolutely crucial. In view of this, the theoretical predicted PES of EuGen (n = 7–20) was calculated via the method of Koopmans’ theorem and the first peak equivalent to vertical detachment energy (VDE) calculated under the mPW2PLYP level. The first ionization potential (IP) is defined as E(anion) − E(neutral). The first peak is set to the first IP + E(HOMO); then, the second peak is obtained by adding the absolute value of the energy difference between the occupied orbital below the HOMO and the HOMO orbital as the second IP, and so on. The number of peaks and their position on the PES reveal the energy differences between diverse N-1 states and the original N-electron state. By default, all orbitals have a strength equal to 1.0, where peak broadening can be plotted by means of strength stacking and Gaussian broadening. A plot for each cluster is shown in Figure 2, with a Gaussian broadening of 0.30 eV for the energy region smaller than 5 eV. The peaks of each EuGen (n = 7–20) cluster with energy from low to high are marked as X and A to E in alphabetical order. For EuGe7, the PES results exhibit four peaks at 2.00 (X), 2.47 (A), 3.55 (B), and 4.71 (C) eV. For EuGe8, in the energy region smaller than 5 eV, there are only two peaks located at 3.43 (X) and 4.87 (A) eV. For EuGe9, there are four peaks (X, A to C) located at 2.33, 3.28, 3.69, and 4.29 eV (respectively). For EuGe10, the PES results for the ground-state structure 10A-1 and its energy-degenerate isomer 10A-2 are shown. They have similar shapes and positions of their X, A, and B peaks due to the similarity of their structures. The PES results for 10A-1 show five peaks in the energy region from 0 to 5 eV, sited at 2.10, 3.49, 3.85, 4.51, and 4.87 eV, compared with those for 10A-2 showing only four peaks at 2.42, 3.60, 4.06, and 4.71 eV; thus, they can be distinguished by the number of peaks. For EuGe11, the six peaks X and A to E are located at 2.63, 2.99, 3.77, 4.28, 4.62, and 4.92 eV, respectively. For EuGe12, three obvious peaks, X, A, and B, are located at 2.61, 3.21, and 4.23 eV, with peaks C and D as tail peaks of B at 4.64 and 4.98 eV. For EuGe13, three clear spikes, X, B, and D, are located at 2.05, 3.92, and 4.66 eV. A and C are shoulder peaks of B at 3.50 and 4.22 eV. For EuGe14, the PES curves of the two near-energy structures 14A-1 and 14A-2 have different numbers of peaks and shapes, making them easy to distinguish. The peaks X, A, and B of 14A-1 are positioned at 3.21, 4.23, and 4.76 eV. The PES results for 14A-2 show two more peaks than those for 14A-1, at 2.76 eV for A, 3.20 eV for B, 3.77 eV for C, 4.10 eV for D, and 4.55 eV for E. For EuGe15, there are three independent peaks at 2.07 eV (A), 4.09 eV (B), and 4.83 eV (C). For EuGe16, four consecutive peaks, X and A to C, appear at positions of 3.20, 3.78, 4.43, and 4.74 eV. For EuGe17, five peaks, X and A to D, are located at 3.02, 3.51, 3.79, 4.28, and 4.75. Among them, the peaks A and B are very close to each other. For EuGe18, three evident peaks, X, B, and C, are sited at 3.19, 4.48, and 4.87 eV. A is a satellite peak in the middle of X and B at 3.84 eV. For EuGe19, the PES results for 19A-1 present two clear peaks, X at 3.30 eV and A at 4.26, with tailed peaks at the end of the energy region denoted as C at 4.52 and D at 4.84 eV. The PES results for 19A-2, compared with those for 19A-1, only show three peaks at energy positions of 3.05, 3.72, and 4.73 eV, so they can be discerned by the curve shape and number of peaks. For EuGe20, in the selected energy region, there are three peaks at 2.77, 4.09, and 4.58 eV. Because there are no comparable experimental results, to some extent, these predicted PES results can provide a reference and instructions for future experimental investigations.

2.3. Magnetic Moments and Charge Transfer of EuGen (n = 7–20)

The Eu atom possesses remarkable magnetism as a result of the spin magnetic moment of electrons in its 4f orbitals. When a Eu atom is doped into a pure germanium cluster, it brings this novel magnetic property to the nanocluster. Considering this, a natural population analysis (NPA) in the NBO 3.1 software was performed to discover the charge, valence electron configuration, and magnetic moments of each orbital and the total structure of EuGen (n = 7–20). The data are collected in Table 1. From Table 1, it can be seen that the Eu atoms in all ground-state structure clusters have a positive charge of from 0.51 a.u. to 0.81 a.u., which means that the Eu atoms in the clusters serve as electron donors and the Ge frameworks serve as electron acceptors. Between the Eu and Ge atoms, ionic bonds are formed. When n = 7–10, the ground-state structure is replaced by a structure in which the Eu atom has a valence electron configuration of 6s0.80–1.004f6.985d0.18–0.416p0.14–0.23. When the ground-state structure changes to a linked type (n = 11–20), the valence electron configuration of Eu becomes 6s0.24–0.514f6.975d0.63–0.856p0.15–0.32. From the valence electron configuration of the Eu atom of Xe [6s24f7], it is known that the 4f electrons of the Eu atom in EuGen (n = 7–20) do not participate in bonds, mainly providing magnetic moments for the cluster. Without bonding involving the 4f electrons of Eu, the bonding between Eu and Ge is mainly attributed to electrons in Eu’s 5d and 6s orbitals hybridizing with electrons in Ge’s 4s and 4p orbitals.
To further explore the magnetism of EuGen (n = 7–20), the spin density and spin population were examined. The spin density three-dimensional contour maps of EuGen (n = 7–20) shown in Figure 3 were obtained using the following equation:
ρ s ( r ) = ρ α ( r ) ρ β ( r )
The equation definition is the electron density of Alpha minus the electron density of Beta in three-dimensional space. It can be easily seen that the iso-surfaces of spin density are all located in the Eu atom, and the values of ρ s ( r ) are all close to or equal to 8, indicating that the magnetism of EuGen (n = 7–20) originates from the Eu atom. A further analysis was carried out via the spin population method, as shown in Table 1. For n = 7–10, the adsorption structure stage, the total magnetic moment contribution by the Eu atom is 7.86 to 7.95 μB, with a contribution ratio exceeding 99.10%. For n = 11–20, the linked stage, the total magnetic moment provided by the Eu atom is 7.11 to 7.48 μB, with a contribution ratio exceeding 88.96%—a slight decrease compared to that of the adsorption structure.

2.4. Relative Stability of EuGen (n = 7–20)

The average binding energy (ABE) is an important parameter used to describe the relative stability with the change in the size of a cluster. The second-order energy difference (Δ2E) is another critical parameter used to evaluate the stability within one cluster of its left and right sides. The ABE and Δ2E are calculated through following Equations (2) and (3):
A B E EuGe n = n 1 E Ge + E ( Ge ) + E ( Eu ) E EuGe n n + 1
Δ 2 E ( EuGe n ) = E ( EuGe n 1 ) + E ( EuGe n + 1 ) 2 E ( EuGe n )
The values of ABE and Δ2E are plotted in Figure 4. From Figure 4a, in the adsorption pattern phase, the ABE rises sharply, with a peak observed for EuGe10 of 4.16 eV. At the linked structure stage, two obvious sections can be seen. When n = 11–17, the increasing trend is smaller than that for the adsorption stage, tending toward a flat curve for n = 18–20. The maximum and minimum values of the ABE within the cluster size were found for EuGe19 at 4.35 eV and EuGe7 at 3.90 eV. In Figure 4b, the Δ2E curve presents four peaks at 0.81, 0.48, 0.45, and 0.37 eV, belonging to EuGe9, EuGe13, EuGe15, and EuGe18; this means that these clusters are more stable than their adjacent ones.
The energy gap (Eg) is a very important physical quantity that denotes the energy difference between the lowest unoccupied molecular orbital and the highest occupied molecular orbital, described in Equation (4):
E g = E ( LUMO ) E ( HOMO )
The Eg values in Figure 4c were calculated using the mPW2PLYP and PBE0 methods. In general, the number of Hartree–Fock components in a functional strongly affects the value of Eg. The mPW2PLYP functional has more Hartree–Fock components and usually over-estimates Eg. However, having fewer Hartree–Fock components in a functional mostly leads to under-estimation of Eg, such as with the pure PBE functional. In view of this, the hybrid density functional PBE0 was implemented to calculate Eg. The trends in Eg values calculated using mPW2PLYP and PBE0 are generally similar, being only slightly different at n ≤ 11; when n > 11, the two Eg curves present the same trend. The maximum Eg values pertain to EuGe14, with 3.67 eV at the mPW2PLYP level and 2.40 eV at the PBE0 level. The minimum Eg values belong to EuGe15, with 2.00 eV at the mPW2PLYP level and 1.24 eV at the PBE0 level. The above results show that the EuGen (n = 7–20) clusters are all semiconductors and potential semiconductor building block candidates for next-generation optoelectronic materials.

2.5. Nonlinear Optical Properties of EuGen (n = 7–20)

The Eu atom is a typical optical activator. In previous works [29,30], lone-pair electrons helped to enhance the nonlinear optical effects of optical materials. The Eu atom has seven single electrons in its 4f orbitals; when an extra electron is brought into the cluster, the whole cluster changes to have eight single electrons. In view of this, the dipole moment (μ0), static polarizability (α0), static first hyperpolarizability (βtot), components of βtot in the x, y, and z directions, and projection of βtot in the direction of the dipole moment (βprj) for EuGen (n = 7–20) clusters were calculated at the level of long-range corrected density functional theory CAM-B3LYP, which is commonly used to evaluate nonlinear optical responses [31,32]. The results are summarized in Table 2. For clusters with n = 7–10, the dipole moment varies between 0.91 and 2.02 a.u. When n increases to the range of 11 to 20, the dipole moment exhibits a significant increase, ranging from 2.27 to 3.94 a.u., which can be attributed to changes in the cluster configuration. The static polarizability of the ground-state structures of EuGen (n = 7–20) demonstrates an overall increasing trend, with particularly notable values at n = 13 and 20, reaching 803.99 and 860.61 a.u., respectively. The second-order nonlinear optical coefficient of the clusters, i.e., the first hyperpolarizability, can reflect the nonlinear optical response of the materials. Thus, the static hyperpolarizability and its components along the three directions of the 14 ground-state clusters were simulated and compared. It was observed that the βtot value for EuGe13 was markedly higher than those of the other clusters, reaching 747,032.61 a.u., with βyyy being the predominant contributing component. These findings suggest that EuGe13 holds significant potential for applications in the development of nonlinear optical materials.
To further elucidate the nonlinear optical properties of EuGen (n = 7–20), the electronic spatial extent R2 of the clusters serves as an effective descriptor quantifying their polarizability and hyperpolarizability [33]. Figure 5a,b, respectively, illustrate the relative relationship between α0 and R2 and that between βtot and R2 as the number of Ge atoms increases. From Figure 5a, it can be seen that the α0 value increases as the cluster size and R2 increase. However, for βtot, there is a linear trend only for small-sized clusters with n < 12 (shown in Figure 5b). Notably, the polarizability reaches a significant peak when n = 13.
Besides the descriptor R2, the relative relationships between the polarizability, hyperpolarizability, and energy gap are illustrated in Figure 6a,b. In the figure, the HOMO–LUMO gap exhibits a negative correlation with α0 and βtot. Smaller Eg values could account for the larger α0 and βtot observed for EuGe13 and EuGe15. Given that the symmetry of EuGe15clusters is higher than that of EuGe13 clusters, the electron cloud distribution in the EuGe15 clusters is more uniform, rendering them less susceptible to polarization by external electric fields. Consequently, the α0 and βtot values for EuGe15 clusters are lower than those for EuGe13 clusters. Additionally, the calculated βprj values can serve as a reference and comparison for hyperpolarizability values obtained from future experiments on electric field-induced second harmonic generation (EFISH).
To obtain a more comprehensive insight into the relationships among the parameters, we performed additional data processing and analysis. Figure 7a reflects the relationship between lg(βtot) and the cluster size. Figure 7b shows the relative relationship between α0 and R2, as well as the fitting curve. In Figure 7a, the hyperpolarizability of EuGe13 is clearly higher than that of any other type. The second peak is observed for EuGe15, which is consistent with the discussion above. In Figure 7b, α0 and R2 exhibit a better linear relationship, with a fitting R-squared value equal to 0.78. Finally, Figure 7c shows a volcano plot to further identify the best hyperpolarizability, which belongs to EuGe13. Overall, the static polarizability exhibits a distinct positive correlation with R2, while the hyperpolarizability shows a trend of first increasing and then decreasing with R2. This suggests that as the electron cloud becomes more extended, the electrons become more responsive to external electric fields, thereby enhancing the likelihood of polarization. Furthermore, the volume and morphology of the clusters influence the outcome.
To avoid basis set effects, the hyperpolarizability was also calculated with aug-cc-pVTZ for Ge and ma-def2-TZVP for Eu [34]. The outcomes were consistent with the discussion above (seen Table S3, Supporting Information).
In summary, the nonlinear optical properties of EuGen (n = 7–20) clusters are strongly correlated with their growth patterns and structural symmetry. As the cluster size increases, the (hyper)polarizability of the cluster exhibits a positive correlation with the overall electron space R2 and a negative correlation with the energy gap. The smaller Eg values might be the reason why the clusters have larger (hyper)polarizability values. Among the investigated clusters, EuGe13 had the highest (hyper)polarizability value of 7.47 × 105 a.u.; that is, R2 and Eg are two important descriptors for evaluating the nonlinear optical properties of EuGen (n = 7–20).

2.6. Nonlinear Optical Properties of EuGe13 from TD-DFT

Whether compared within the same system or with some other materials [31,32,35], EuGe13 displays significant βtot values; thus, its (hyper)polarizability density was targeted for further examination. Figure 8 presents the local contributions of EuGe13 polarizability and hyperpolarizability along the y direction in a static electric field. Positive contributions are represented by blue isosurfaces, whereas negative contributions are depicted by white isosurfaces. This visualization method facilitates a clearer understanding of the three-dimensional spatial contributions to the (hyper)polarizability. From Figure 8a, it is evident that the region of positive contribution substantially exceeds that of negative contribution. This discrepancy in areas provides a clear rationale for the cluster’s relatively large positive polarizability. For the hyperpolarizability, despite the presence of distinct positive and negative contribution areas in Figure 8b, the negative contribution regions are more extensive. This is consistent with the significantly negative hyperpolarizability of EuGe13. In general, both the polarizability and hyperpolarizability exhibit a significant contribution region surrounding the Eu atom, which indicates that Eu atoms in the system have a stronger response to the external electric field, enhancing the nonlinear optical properties of the cluster.
The two-level expression (5) is widely employed to qualitatively investigate the NLO property of a system in depth [36,37]. Utilizing this formula, the crucial excitation energies of the EuGe13 cluster were calculated within the framework of TD-DFT:
β tot Δ μ f 0 Δ E 3
where Δμ is the difference in dipole moment between the ground state and the crucial excited state, f0 is the oscillator strength, and ΔE is the transition energy of the key transition. Figure 9 shows the simulated UV–vis spectrum and the molecular orbital diagram of the main transitions of the EuGe13 cluster. The strongest absorption peak at 2710.63 nm is dominated by N0 → N1, where N stands for a nonet. ΔE is 0.457a.u. The dominant electronic excitation mode in this transition is HOMO(α) → LUMO(α) with a contribution rate of 97%. N0 → N37 mainly contributes to the absorption peak at 474 nm. The f0 value of this crucial excitation is 0.0214 a.u., ΔE is 2.51 eV, and Δμ is 0.348 a.u. Multiple electronic excitation modes are observed in this transition, with significant contributions from HOMO(α) → LUMO+12(α) (35%) and HOMO-3(α) → LUMO+1(α) (14%). The relatively large βtot value of the cluster can be attributed to factors such as the large Δμ and small ΔE values associated with the transitions. The relevant parameters for other clusters are collected in Table S4.

2.7. Density of States of EuGe13

The density of states (DOSs) of a cluster can provide a deep understanding of the bonding characteristics within it. Figure 10 displays the DOS of the EuGe13 cluster, where the black curve is the total DOS and the colored curves are the partial DOS curves of different orbits. As illustrated in the figure, the HOMO and LUMO are predominantly contributed by the Ge-s, p, and Eu-s orbitals; for the HOMO level, these contributions account for 25.91%, 35.13%, and 36.83%, respectively, and for the LUMO level, the corresponding contributions are 46.02%, 25.96%, and 27.11%. In the energy region below the HOMO, the most prominent peak in the DOS is observed at approximately −7.05 eV, which is predominantly attributed to the 4f orbitals of the Eu atom (89.18%). The remainder is mainly contributed by the Ge-s and p orbitals. It is evident that there is a pronounced asymmetry between the spin-up Alpha and spin-down Beta, especially the spin-up peak at −7.05 eV. Based on the NPA, it is once again demonstrated that the magnetic property of the cluster is mainly provided by the 4f orbitals of Eu and that the cluster exhibits spin polarization effects and magnetism.

3. Computational Details

All density functional theory calculations were performed using the Gaussian 09 quantum chemistry software [38]. The initial structures were obtained (1) using the ABCluster global searching software, version 3.1 and (2) from previously reported structures [39,40,41]. By way of (1), each cluster generated at least 400 isomers to ensure that the global minimum structure was found. The initial structures were optimized with TPSSh [42] combined with BS-1 basis sets (ECP28MWB [43] for Ge and ECP53MWB [44,45] for Eu). Then, the structures with an energy difference of within 0.8 eV were selected and further optimized using TPSSh combined with BS-2 basis sets (cc-pVTZ-PP [46,47] for Ge and ECP28MWB [48] for Eu). At the same time, frequency calculations were carried out to rule out imaginary frequencies, indicating that the structure is a true local minimum. Then, the above structures were further optimized at the mPW2PLYP [49] level with BS-2 basis sets, but frequency calculations were not carried out in order to reduce the calculation time. Finally, the single-point energy was calculated at the mPW2PLYP level with BS-3 basis sets (aug-cc-pVTZ [50] for Ge and ECP28MWB for Eu).
The PES spectra of EuGen (n = 7–20) nanoclusters were obtained by applying Koopmans’ theorem [51,52] at the level of mPW2PLYP, and ultraviolet–visible (UV–vis) spectra were simulated via the time-dependent density functional theory method of the PBE0 functional with BS-3 basis sets [53,54]. The dipole moment, static polarizability, and first hyperpolarizability of each EuGen (n = 7–20) cluster were calculated by means of the CAM-B3LYP functional [55] with BS-3 basis sets. The results were visualized using Multifwn [56,57] and VMD software, VMD 1.9.4 [58].
The formulas for the dipole moment (μ0), static polarizability (α0), and static first hyperpolarizability (βtot) are presented as follows:
μ 0 = μ x 2 + μ y 2 + μ z 2 1 / 2
α 0 = 1 / 3 α x x + α y y + α z z
β t o t = β x 2 + β y 2 + β z 2 1 / 2
where β i = 1 / 3 j β i j j 2 + β j j i 2 + β j i j 2 i , j = x , y , z .
The ab initio molecular dynamics (AIMD) of EuGen (n = 7–20) were obtained using Orca quantum chemistry code [59] at the level of the B97-3c functional [60] with def2-mTZVP for Ge and def2-mTZVP for Eu combined with Def2-ECP basis sets [48]. The B97-3c functional contains the Becke–Johnson damping scheme (D3BJ) [61,62].
mPW2PLYP displays high reliability in predicting the ground-state structures of rare-earth elements doped into silicon or germanium clusters. The different density functionals mPW2PLYP, TPSSh, PBE, wB97X, and B3LYP were compared with the ROCCSD(T) method for ScSin0/− (n = 4–9) in a previous work [63], and the results showed that all the ground-state structures predicted using mPW2PLYP were in line with those obtained using ROCCSD(T). Besides those of rare-earth-element-doped silicon clusters, the ground-state structures of ScGen (n = 6–17) and CeGen (n = 5–17) predicted using the mPW2PLYP approach agreed with those predicted via closed-shell DLPNO-CCSD(T1) and open-shell DLPNO-CCSD(T) methods [23,64], which further proved that the mPW2PLYP functional is suitable for such systems, combining accuracy and speed. The optimization scheme and TD-DFT calculation were also based on the above previous works.

4. Conclusions

This study systematically investigated the structural evolution and nonlinear optical properties of europium-doped germanium anionic clusters EuGen (n = 7–20), utilizing an approach that combined global search techniques and double-hybrid density functional theory mPW2PLYP. The growth patterns were categorized into two distinct stages: (i) for n = 7–10, the structures were mainly characterized by an adsorption configuration with a Eu atom adsorbed onto a pure Ge cluster, and (ii) for n = 11–20, the clusters exhibited linked structures wherein a Eu atom was individually linked to a Ge9 sub-cluster with a TTP configuration and a Gen-9 sub-cluster. The PES spectra, charge transfer, magnetic moments, relative stability, energy gaps, and (hyper)polarizability of the clusters were simulated and analyzed. EuGe13, which exhibits notable NLO properties and a βtot value of 7.47 × 105 a.u., was subjected to a further investigation. The findings indicate that EuGe13 holds potential as a nonlinear optical semiconductor material for application in multifunctional nanomaterials.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30061377/s1, Figure S1: The ab initio molecular dynamics of EuGen (n = 7–20); Table S1: The Cartesian coordinates of EuGen (n = 7–20); Table S2: Total energies at mPW2PLYP with zero-point energy at TPSSh (in Hartree); Table S3: The dipole moment (μ0, in a.u.), static polarizability (α0 in a.u.), static first hyperpolarizability (βtot, in a.u.), the components of hyperpolarizability in the x, y, and z directions (βxxx βyyy βzzz, in a.u.) as well as the projection of the hyperpolarizability in the direction of the dipole moment (βprj, in a.u.) of EuGen (n = 7–20) with BS4 (aug-cc-pVTZ for Ge and ma-def2-TZVP for Eu); Table S4: The main transitions, oscillator strengths (f0, in a.u.), difference in dipole moment (Δμ, in a.u.) transition energies (ΔE, in eV), wavelengths (λ, in nm) and orbital contribution of the crucial excited states for the EuGen (n = 7–20) cluster.

Author Contributions

C.H.: Investigation, Writing—original draft. X.D.: Investigation, Data curation. C.L. Software, Visualization, Formal analysis. C.D.: Conceptualization, Writing—review & editing, Funding acquisition. Z.Y.: Conceptualization, Writing—review & editing, Supervision. J.Y.: Writing—review & editing, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 21863007), by the Inner Mongolia Natural Science Foundation (Grant No. 2023QN02023), by the Inner Mongolia Autonomous Region High Level Talent Introduction and Research Support Project (Grant No. DC2400002178), by the Doctoral Initiation Fund Project of Inner Mongolia University of Technology (Grant Nos. DC2300001273 and BS2024059), by Central Guidance for Local Scientific and Technological Development Funding Projects (Grant No. 2024ZY0134), by Special Programs for Research in First-Class Disciplines (Grant No. YLXKZX-NGD-016), by Talent Revitalization of Inner Mongolia, and by the Key Laboratory of Environmental Pollution Control and Remediation at Universities of Inner Mongolia Autonomous Region.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. El Kassaoui, M.; Labrousse, J.; Loulidi, M.; Benyoussef, A.; Mounkachi, O. Ni/V-MgnHm nanoclusters: Recent advances toward improving the dehydrogenation thermodynamics for efficient hydrogen storage. Int. J. Hydrogen Energy 2024, 89, 272–278. [Google Scholar] [CrossRef]
  2. Boubkri, M.; El Kassaoui, M.; Razouk, A.; Balli, M. Computational investigation of NLi4-cluster decorated phosphorene for reversible hydrogen storage. Int. J. Hydrogen Energy 2024, 72, 1–8. [Google Scholar] [CrossRef]
  3. Pillarisetty, R. Academic and industry research progress in germanium nanodevices. Nature 2011, 479, 324–328. [Google Scholar] [CrossRef]
  4. Kilby, J.S.C. Turning potential into realities: The invention of the integrated circuit (Nobel lecture). Chem. Phys. Chem. 2001, 2, 482–489. [Google Scholar] [CrossRef]
  5. Abyar, F.; Bamdadi, F.; Behjatmanesh-Ardakani, R. Geometric, electronic and spectral properties of germanium and Eu-doped germanium clusters. Comput. Theor. Chem. 2022, 1214, 113783. [Google Scholar] [CrossRef]
  6. Trivedi, R.; Banerjee, A.; Bandyopadhyay, D. Insight into stabilities and magnetism of EuGen (n = 1–20) nanoclusters: An assessment of electronic aromaticity. J. Mater. Sci. 2022, 57, 19338–19355. [Google Scholar] [CrossRef]
  7. Jing, Q.; Ge, G.; Cao, H.; Huang, X.; Yan, H. Structural, electronic and magnetic properties of the GenEu (n = 1–13) clusters. Acta Phys. Chim. Sin. 2010, 26, 2510–2514. [Google Scholar]
  8. Politano, G.G. Optical properties of Graphene Nanoplatelets on amorphous Germanium substrates. Molecules 2024, 29, 4089. [Google Scholar] [CrossRef]
  9. Jena, P.; Sun, Q. Super atomic clusters: Design rules and potential for building blocks of materials. Chem. Rev. 2018, 118, 5755–5870. [Google Scholar] [CrossRef] [PubMed]
  10. Zhao, J.; Du, Q.; Zhou, S.; Kumar, V. Endohedrally doped cage clusters. Chem. Rev. 2020, 120, 9021–9163. [Google Scholar] [CrossRef]
  11. Zhao, L.J.; Xu, H.G.; Xu, X.L.; Zheng, W.J. Investigation of the structures and chemical bonding of Mn2Ge6 and Mn2Ge7 clusters via anion photoelectron spectroscopy and theoretical calculation. Inorg. Chem. 2023, 6, 2033–2039. [Google Scholar] [CrossRef] [PubMed]
  12. Atobe, J.; Koyasu, K.; Furuse, S.; Nakajima, A. Anion photoelectron spectroscopy of germanium and tin clusters containing a transition-or lanthanide-metal atom; MGen (n = 8–20) and MSnn (n = 15–17) (M = Sc–V, Y–Nb, and Lu–Ta). Phys. Chem. Chem. Phys. 2012, 14, 9403–9410. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, J.; Han, J.G. Geometries, stabilities, and vibrational properties of bimetallic Mo2-doped Gen (n = 9−15) clusters: A density functional investigation. J. Phys. Chem. A 2008, 112, 3224–3230. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, K.; Jia, Z.Z.; Wang, R.Y.; Zhu, X.D.; Moro, R.; Ma, L. TMGe8–17 (TM = Ti, Zr, Hf, V, Nb, Ta) clusters: Group determined properties. Eur. Phys. J. Plus 2022, 137, 949. [Google Scholar] [CrossRef]
  15. Liang, X.; Li, X.; Gao, N.; Wu, X.; Zhao, Z.; Shi, R.; Su, Y.; Zhao, J. Theoretical prediction for growth behavior and electronic properties of monoanionic Ru2Gen (n = 3–20) clusters. Inorg. Chim. Acta 2022, 542, 121141. [Google Scholar] [CrossRef]
  16. Zhang, Y.; Li, X.; Lu, J.; Li, S.; Zhang, Y. Endohedral group-14 clusters Au@X12 (X = Ge, Sn, Pb) and their anions: A first-principles study. J. Mol. Liq. 2023, 376, 121477. [Google Scholar] [CrossRef]
  17. Lasmi, M.; Mahtout, S.; Rabilloud, F. Growth behavior and electronic and optical properties of IrGen (n = 1–20) clusters. J. Nanopart. Res. 2021, 23, 26. [Google Scholar] [CrossRef]
  18. Kumar, V.; Kawazoe, Y. Metal-encapsulated caged clusters of germanium with large gaps and different growth behavior than silicon. Phys. Rev. Lett. 2002, 88, 235504. [Google Scholar] [CrossRef]
  19. Han, J.G.; Hagelberg, F. Recent progress in the computational study of silicon and germanium clusters with transition metal impurities. J. Comput. Theor. Nanosci. 2009, 6, 257–269. [Google Scholar] [CrossRef]
  20. Nguyen, H.T.; Cuong, N.T.; Lan, N.T.; Tung, N.T.; Nguyen, M.T.; Tam, N.M. First-row transition metal doped germanium clusters Ge16M: Some remarkable superhalogens. RSC Adv. 2022, 12, 13487–13499. [Google Scholar] [CrossRef]
  21. Bandyopadhyay, D.; Kaur, P.; Sen, P. New insights into applicability of electron-counting rules in transition metal encapsulating Ge cage clusters. J. Phys. Chem. A 2010, 114, 12986–12991. [Google Scholar] [CrossRef] [PubMed]
  22. Borshch, N.A.; Pereslavtseva, N.S.; Kurganskii, S.I. Spatial structure and electron energy spectra of ScGen (n = 6–16) clusters. Russ. J. Phys. Chem. B 2015, 9, 9–18. [Google Scholar] [CrossRef]
  23. Wang, H.; Yang, J.; Lin, L. Geometric evolution, electronic structure, and vibrational properties of ScGen (n = 6–17) anion clusters: A DFT insight. Int. J. Quantum Chem. 2022, 122, e27002. [Google Scholar] [CrossRef]
  24. Tang, C.; Liu, M.; Zhu, W.; Deng, K. Probing the geometric, optical, and magnetic properties of 3d transition-metal endohedral Ge12M (M = Sc–Ni) clusters. Comput. Theor. Chem. 2011, 969, 56–60. [Google Scholar] [CrossRef]
  25. Li, A.; Li, H.; Li, Z.; Qin, L.; Mei, X.; Zhang, J.; Zhang, Y.; Zheng, H.; Jiang, K.; Wu, W.; et al. Structural evolution and electronic properties of the La-doped germanium clusters. Mol. Phys. 2024, 123, e2356191. [Google Scholar] [CrossRef]
  26. Zhang, J.; Wang, H.; Li, H.; Mei, X.; Zeng, J.; Qin, L.; Zheng, H.; Zhang, Y.; Jiang, K.; Zhang, B.; et al. Aromatic and magnetic properties in a series of heavy rare earth-doped Ge6 cluster anions. J. Comput. Chem. 2024, 45, 1087–1097. [Google Scholar] [CrossRef]
  27. Qin, W.; Lu, W.C.; Xia, L.H.; Zhao, L.Z.; Zang, Q.J.; Wang, C.Z.; Ho, K.M. Structures and stability of metal-doped GenM (n = 9, 10) clusters. AIP Adv. 2015, 5, 067159. [Google Scholar] [CrossRef]
  28. Tai, T.B.; Nguyen, M.T. A stochastic search for the structures of small germanium clusters and their anions: Enhanced stability by spherical aromaticity of the Ge10 and Ge122− systems. J. Chem. Theory Comput. 2011, 7, 1119–1130. [Google Scholar] [CrossRef]
  29. Zhong, Q. Lone pair electrons with weak nuclear binding inducing sensitive nonlinear optical responses in phosphorus clusters. J. Phys. Chem. Lett. 2023, 14, 6361–6367. [Google Scholar] [CrossRef]
  30. Yin, R.; Hu, C.; Lei, B.; Pan, S.; Yang, Z. Lone pair effects on ternary infrared nonlinear optical materials. Phys. Chem. Chem. Phys. 2019, 21, 5142. [Google Scholar] [CrossRef]
  31. Li, X. Design of novel graphdiyne-based materials with large second-order nonlinear optical properties. J. Mater. Chem. C 2018, 6, 7576–7583. [Google Scholar] [CrossRef]
  32. Li, X.; Li, S. Investigations of electronic and nonlinear optical properties of single alkali metal adsorbed graphene, graphyne and graphdiyne systems by first-principles calculations. J. Mater. Chem. C 2019, 7, 1630–1640. [Google Scholar] [CrossRef]
  33. Xu, H.; Li, Z.; Wang, F.; Wu, D.; Harigaya, K.; Gu, F.L. What is the shape effect on the (hyper)polarizabilities? A comparison study on the Mobius, normal cyclacene, and linear nitrogen-substituted strip polyacenes. Chem. Phys. Lett. 2008, 454, 323–326. [Google Scholar] [CrossRef]
  34. Zheng, J.; Xu, X.; Truhlar, D.G. Minimally augmented Karlsruhe basis sets. Theor. Chem. Acc. 2011, 128, 295–305. [Google Scholar] [CrossRef]
  35. Li, X.; Lu, J. Giant enhancement of electronic polarizability and the first hyperpolarizability of fluoride-decorated graphene versus graphyne and graphdiyne: Insights from ab initio calculations. Phys. Chem. Chem. Phys. 2019, 21, 13165–13175. [Google Scholar] [CrossRef]
  36. Oudar, J.L. Optical nonlinearities of conjugated molecules. Stilbene derivatives and highly polar aromatic compounds. J. Chem. Phys. 1977, 67, 446–457. [Google Scholar] [CrossRef]
  37. 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]
  38. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian~09, Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  39. Zhang, J.; Dolg, M. ABCluster: The artificial bee colony algorithm for cluster global optimization. Phys. Chem. Chem. Phys. 2015, 17, 24173–24181. [Google Scholar] [CrossRef]
  40. Zhang, J.; Dolg, M. Global optimization of clusters of rigid molecules using the artificial bee colony algorithm. Phys. Chem. Chem. Phys. 2016, 18, 3003–3010. [Google Scholar] [CrossRef]
  41. Zhang, J.; Glezakou, V.A.; Rousseau, R.; Nguyen, M.T. NWPEsSe: An adaptive-learning global optimization algorithm for nanosized cluster systems. J. Chem. Theory Comput. 2020, 16, 3947–3958. [Google Scholar] [CrossRef]
  42. Staroverov, V.N.; Scuseria, G.E.; Tao, J.; Perdew, J.P. Comparative assessment of a new nonempirical density functional: Molecules and hydrogen-bonded complexes. J. Chem. Phys. 2003, 119, 12129–12137. [Google Scholar] [CrossRef]
  43. Bergner, A.; Dolg, M.; Küchle, W.; Stoll, H.; Preuß, H. Ab initio energy-adjusted pseudopotentials for elements of groups 13–17. Mol. Phys. 1993, 80, 1431–1441. [Google Scholar] [CrossRef]
  44. Dolg, M.; Stoll, H.; Preuss, H. A combination of quasirelativistic pseudopotential and ligand field calculations for lanthanoid compounds. Theor. Chim. Acta 1993, 85, 441–450. [Google Scholar] [CrossRef]
  45. Dolg, M.; Stoll, H.; Savin, A.; Preuss, H. Energy-adjusted pseudopotentials for the rare earth elements. Theor. Chim. Acta 1989, 75, 173–194. [Google Scholar] [CrossRef]
  46. Metz, B.; Stoll, H.; Dolg, M. Small-core multiconfiguration-Dirac–Hartree–Fock-adjusted pseudopotentials for post-d main group elements: Application to PbH and PbO. J. Chem. Phys. 2000, 113, 2563–2569. [Google Scholar] [CrossRef]
  47. Peterson, K.A. Systematically convergent basis sets with relativistic pseudopotentials. I. Correlation consistent basis sets for the post-d group 13–15 elements. J. Chem. Phys. 2003, 119, 11099–11112. [Google Scholar] [CrossRef]
  48. Dolg, M.; Stoll, H.; Preuss, H. Energy-adjusted ab initio pseudopotentials for the rare earth elements. J. Chem. Phys. 1989, 90, 1730–1734. [Google Scholar] [CrossRef]
  49. Schwabe, T.; Grimme, S. Towards chemical accuracy for the thermodynamics of large molecules: New hybrid density functionals including non-local correlation effects. Phys. Chem. Chem. Phys. 2006, 8, 4398–4401. [Google Scholar] [CrossRef]
  50. Wilson, A.K.; Woon, D.E.; Peterson, K.A.; Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. IX. The atoms gallium through krypton. J. Chem. Phys. 1999, 110, 7667–7676. [Google Scholar] [CrossRef]
  51. Tozer, D.J.; Handy, N.C. Improving virtual Kohn–Sham orbitals and eigenvalues: Application to excitation energies and static polarizabilities. J. Chem. Phys. 1998, 109, 10180–10189. [Google Scholar] [CrossRef]
  52. Akola, J.; Manninen, M.; Häkkinen, H.; Landman, U.; Li, X.; Wang, L. Photoelectron spectra of aluminum cluster anions: Temperature effects and ab initio simulations. Phys. Rev. B 1999, 60, R11297. [Google Scholar] [CrossRef]
  53. Liu, Z.; Lu, T.; Chen, Q. An sp-hybridized all-carboatomic ring, cyclo [18] carbon: Electronic structure, electronic spectrum, and optical nonlinearity. Carbon 2020, 165, 461–467. [Google Scholar] [CrossRef]
  54. Runge, E.; Gross, E.K.U. Density-functional theory for time-dependent systems. Phys. Rev. Lett. 1984, 52, 997. [Google Scholar] [CrossRef]
  55. Yanai, T.; Tew, D.P.; Handy, N.C. A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef]
  56. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  57. Lu, T. A comprehensive electron wavefunction analysis toolbox for chemists, Multiwfn. J. Comput. Chem. 2024, 161, 082503. [Google Scholar] [CrossRef]
  58. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graphics 1996, 14, 33–38. [Google Scholar] [CrossRef]
  59. Neese, F. The ORCA program system. Wires Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
  60. Brandenburg, J.G.; Bannwarth, C.; Hansen, A.; Grimme, S. B97-3c: A revised low-cost variant of the B97-D density functional method. J. Chem. Phys. 2018, 148, 64104. [Google Scholar] [CrossRef] [PubMed]
  61. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  62. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  63. Liu, Y.; Yang, J.; Cheng, L. Structural stability and evolution of scandium-doped silicon clusters: Evolution of linked to encapsulated structures and its influence on the prediction of electron affinities for ScSin (n = 4–16) clusters. Inorg. Chem. 2018, 57, 12934–12940. [Google Scholar] [CrossRef] [PubMed]
  64. Hao, C.; Dong, C.; Yang, Z. Structural evolution and electronic properties of cerium doped germanium anionic nanocluster CeGen (n = 5–17): Theoretical investigation. Int. J. Quantum Chem. 2024, 124, e27315. [Google Scholar] [CrossRef]
Figure 1. The lowest-energy structures and other isomers of EuGen (n = 7–20) after mPW2PLYP method optimization. Pink and cyan represent Eu and Ge atoms, respectively.
Figure 1. The lowest-energy structures and other isomers of EuGen (n = 7–20) after mPW2PLYP method optimization. Pink and cyan represent Eu and Ge atoms, respectively.
Molecules 30 01377 g001
Figure 2. Simulated PES of EuGen (n = 7–20).
Figure 2. Simulated PES of EuGen (n = 7–20).
Molecules 30 01377 g002
Figure 3. Spin density contour maps of EuGen (n = 7–20) clusters.
Figure 3. Spin density contour maps of EuGen (n = 7–20) clusters.
Molecules 30 01377 g003
Figure 4. The (a) average binding energy (ABE, in eV), (b) second-order energy difference (Δ2E, in eV), and (c) HOMO–LUMO gap (in eV) of EuGen (n = 7–20) clusters.
Figure 4. The (a) average binding energy (ABE, in eV), (b) second-order energy difference (Δ2E, in eV), and (c) HOMO–LUMO gap (in eV) of EuGen (n = 7–20) clusters.
Molecules 30 01377 g004aMolecules 30 01377 g004b
Figure 5. The relative relationships between (a) α0 and R2 and (b) βtot and R2.
Figure 5. The relative relationships between (a) α0 and R2 and (b) βtot and R2.
Molecules 30 01377 g005
Figure 6. The relationships of (a) static polarizability (α0) and (b) static hyperpolarizability (βtot) with the HOMO–LUMO energy gap (Eg) of EuGen (n = 7–20) clusters.
Figure 6. The relationships of (a) static polarizability (α0) and (b) static hyperpolarizability (βtot) with the HOMO–LUMO energy gap (Eg) of EuGen (n = 7–20) clusters.
Molecules 30 01377 g006
Figure 7. The relationship between (a) lg(βtot) and cluster size and the relative relationships between (b) α0 and R2 and (c) lg(βtot) and R2 for different cluster sizes, with corresponding fitting curves.
Figure 7. The relationship between (a) lg(βtot) and cluster size and the relative relationships between (b) α0 and R2 and (c) lg(βtot) and R2 for different cluster sizes, with corresponding fitting curves.
Molecules 30 01377 g007aMolecules 30 01377 g007b
Figure 8. (a) The density of polarizability (isovalue = 0.6 a.u.) and (b) the density of hyperpolarizability (isovalue = 30 a.u.) of EuGe13.
Figure 8. (a) The density of polarizability (isovalue = 0.6 a.u.) and (b) the density of hyperpolarizability (isovalue = 30 a.u.) of EuGe13.
Molecules 30 01377 g008
Figure 9. The simulated UV–vis spectrum and electronic excitation modes of the crucial excited state of EuGe13.
Figure 9. The simulated UV–vis spectrum and electronic excitation modes of the crucial excited state of EuGe13.
Molecules 30 01377 g009
Figure 10. The DOS of the EuGe13 cluster.
Figure 10. The DOS of the EuGe13 cluster.
Molecules 30 01377 g010
Table 1. The NPA valence electron configuration, each valence orbital’s magnetic moment, the charge of the Eu atom in EuGen (n = 7–20), and the total magnetic moment of the whole cluster.
Table 1. The NPA valence electron configuration, each valence orbital’s magnetic moment, the charge of the Eu atom in EuGen (n = 7–20), and the total magnetic moment of the whole cluster.
ClusterCharge
(a.u.)
Electron ConfigurationMagnetic Moment of Eu AtomMolecule
B)
6s4f5d6pTotal
EuGe70.57[core]6s0.964f6.985d0.326p0.180.756.970.050.127.898
EuGe80.51[core]6s0.894f6.985d0.406p0.230.716.970.050.137.868
EuGe90.69[core]6s0.804f6.985d0.416p0.140.646.970.060.077.748
EuGe100.67[core]6s1.004f6.985d0.186p0.170.706.980.030.137.848
EuGe110.81[core]6s0.404f6.975d0.676p0.170.006.970.090.017.078
EuGe120.80[core]6s0.414f6.975d0.696p0.150.016.970.130.017.128
EuGe130.70[core]6s0.514f6.975d0.636p0.210.286.970.110.077.438
EuGe140.69[core]6s0.274f6.975d0.796p0.310.016.970.090.017.088
EuGe150.66[core]6s0.344f6.975d0.826p0.240.106.970.140.027.238
EuGe160.74[core]6s0.274f6.975d0.756p0.300.016.970.090.007.078
EuGe170.68[core]6s0.244f6.975d0.856p0.290.006.970.160.017.148
EuGe180.73[core]6s0.244f6.975d0.826p0.290.046.960.120.017.138
EuGe190.77[core]6s0.264f6.975d0.756p0.280.006.970.090.007.068
EuGe200.66[core]6s0.264f6.975d0.836p0.320.026.960.090.017.088
Table 2. The dipole moment (μ0, in a.u.); static polarizability (α0, in a.u.); static first hyperpolarizability (βtot, in a.u.); components of hyperpolarizability in the x, y, and z directions (βxxx βyyy βzzz, in a.u.); and projection of hyperpolarizability in the direction of the dipole moment (βprj, in a.u.) of EuGen (n = 7–20).
Table 2. The dipole moment (μ0, in a.u.); static polarizability (α0, in a.u.); static first hyperpolarizability (βtot, in a.u.); components of hyperpolarizability in the x, y, and z directions (βxxx βyyy βzzz, in a.u.); and projection of hyperpolarizability in the direction of the dipole moment (βprj, in a.u.) of EuGen (n = 7–20).
n.μ0α0βprjβtotβxxxβyyyβzzz
71.37454.46−2242.882281.641179.931952.850.00
80.91488.37−3911.633911.630.000.00−3911.63
92.02548.018681.478681.48107.488680.810.00
101.40553.61−10,775.9910,801.04−6283.05−8785.540.00
113.94527.99−20,833.6620,895.1317,745.6010,984.98−1014.91
123.47552.37−12,056.4312,066.615700.4010,634.59118.72
133.34803.99−726,411.06747,032.61373,853.70−630,754.60142,967.70
142.27630.10−2737.343113.671204.372871.310.00
153.48680.11−96,087.0496,087.040.000.00−96,087.04
162.98682.23−2359.592551.032538.74−250.110.00
172.84738.93−2682.552776.80137.43−2771.25−109.35
182.47750.25−2526.952557.41−1881.68−1731.940.00
192.70788.58−2690.122736.342618.65793.880.00
202.53860.61−3405.023442.56−3111.181473.710.00
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

Hao, C.; Dong, X.; Li, C.; Dong, C.; Yang, Z.; Yang, J. Structural, Electronic, and Nonlinear Optical Characteristics of Europium-Doped Germanium Anion Nanocluster EuGen (n = 7–20): A Theoretical Investigation. Molecules 2025, 30, 1377. https://doi.org/10.3390/molecules30061377

AMA Style

Hao C, Dong X, Li C, Dong C, Yang Z, Yang J. Structural, Electronic, and Nonlinear Optical Characteristics of Europium-Doped Germanium Anion Nanocluster EuGen (n = 7–20): A Theoretical Investigation. Molecules. 2025; 30(6):1377. https://doi.org/10.3390/molecules30061377

Chicago/Turabian Style

Hao, Chenliang, Xueyan Dong, Chunli Li, Caixia Dong, Zhaofeng Yang, and Jucai Yang. 2025. "Structural, Electronic, and Nonlinear Optical Characteristics of Europium-Doped Germanium Anion Nanocluster EuGen (n = 7–20): A Theoretical Investigation" Molecules 30, no. 6: 1377. https://doi.org/10.3390/molecules30061377

APA Style

Hao, C., Dong, X., Li, C., Dong, C., Yang, Z., & Yang, J. (2025). Structural, Electronic, and Nonlinear Optical Characteristics of Europium-Doped Germanium Anion Nanocluster EuGen (n = 7–20): A Theoretical Investigation. Molecules, 30(6), 1377. https://doi.org/10.3390/molecules30061377

Article Metrics

Back to TopTop