DOI:
10.1039/C6RA02635D
(Paper)
RSC Adv., 2016,
6, 47434-47442
Aromatic-like behavior of germanium nanocrystals
Received
29th January 2016
, Accepted 9th May 2016
First published on 9th May 2016
Abstract
Based on density functional theory calculations, aromatic germanium nanocrystals Ge19H12, which are composed of three parallel, planar hexagons with one additional Ge atom close to the cluster center, have been predicted. By exploring electron deficiency and electron delocalization in the proposed nanocrystals, the definition of the electron-deficiency aromaticity concept is extended to germanium nanocrystals. The natural bond orbital analysis and nucleus independent chemical shifts have been used to investigate the electron delocalization in the nanocrystals. We show new patterns for the optical spectrum of Ge19H12 which may be attributed to the special electronic structure of the studied nanoparticles. This study reports relevant details for the synthesis and structuring of overcoordinated germanium nanocrystals and deserves further work for the development of germanium-based nanoparticles for energy and microengineering purposes.
1. Introduction
Aromaticity, which was originally developed to account for the properties of planar organic molecules, has been recognized as a useful concept in structure stability of aromatic hydrocarbons. In general, aromaticity is not an observable and measurable quantity. Despite the lack of a clear and unambiguous definition of aromaticity, however, the conventional carbon aromaticity is referring to the electronic cyclic delocalization in aromatic species. This concept has nowadays extended to involve inorganic systems.1–5 More recently, based on first principal calculations, Vach proposed a novel approach to induce electron delocalization and, hence, aromatic-like behavior in silicon nanocrystals Si NCs.6,7 The ultrastable Si19H12 is composed of three parallel, planar hexagons with one additional Si atom close to the cluster center causing its hypervalent and electron-deficient character. We like to remind you here that “ultrastable” does not necessarily mean global minimum energy structure, but rather that the electron deficiency induces strong electron delocalization that leads to non-tetrahedral entities that are more stable than their tetrahedral counterparts.8 Consequently, the use of those non-tetrahedral nanocrystals for the fabrication of nanodevices in realistic plasma reactors has been predicted. In an extensive quantum chemical study, we considered a series of modifications on hydrogenated silicon nanocluster, with additional insertion of one silicon Si or one Ge atom close to the center of the selected Si18X12 nanostructures, where X = F, Cl, Br, NH2, COOH, and OH.9 Our calculations shown a series of new patterns for optical, chemical, electronic, and conductive properties, which can be attributed to the special structure and modifications by various substitutional molecular groups of silicon nanoparticles.
Silicon and germanium nanocrystals have emerged as one of the most intriguing candidates to control quantum phenomena at the nanoscale. This control leads to outstanding properties and make them to a wide field of applications ranging from memory devices, solar cells, optoelectronic devices and the use as biomarkers.10–14 Compared with Si, Ge has larger excitonic Bohr radius which implies more prominent quantum confinement effects than that of Si.15 In general, the pure Ge clusters are known to be chemically reactive and not quite stable and suitable as building blocks of self-assembly materials.16 By using an appropriate dopant, it is however possible to modify the cluster chemical properties but those metals are often expensive or toxic.
The presence of aromaticity in silicon nanocrystals raises the question whether Ge can be able to form such aromatic-like behavior nanocrystals? In other words, the proposed new concept of electron-deficiency aromaticity might be expended to germanium nanocrystals Ge NCs showing a tendency to overcoordination? In this study, we predict the existence of hydrogenated germanium nanocrystals Ge19H12; see Fig. 1, which has remarkably similar properties to Si19H12. The studied overcoordinated nanocrystals exhibits a highly aromatic character and potential applications in nanotechnological systems. By using density functional theory DFT calculations, we will explore the origin of electron delocalization and aromaticity properties of studied germanium nanocrystals. In addition, Ge19H12 nanocrystals exhibits enhanced stability compared to empty Ge18H12 that can be rationalized by the presence of aromaticity in overcoordinated Ge19H12. The results obtained in this paper extend the bonding capacity of germanium to planar hexacoordination in pure Ge NCs with new pattern for optical properties without the need for expensive or toxic metal atoms. Due to the technological importance of aromatic derivatives, we believe that the proposed Ge NCs with aromatic-like behavior will show an attractive future studies. Therefore, such aromatic NCs, if successfully synthesized, could have application areas ranging from photovoltaic, light-emitting and memory devices to photothermal cancer treatment.
 |
| Fig. 1 Optimized structures of (A) Ge18H12; (B) Ge19H12, peripheral germanium atoms (GeP), middle-layer germanium atoms (GeM) and the central germanium atom (GeC). | |
2. Computational details
The Ge18H12 and Ge19H12 nanocrystals were optimized in Gaussian 03 (ref. 17) and using the B3LYP functional18 with the 6-31G* basis set.19 Frequency calculations of the optimized structures were carried out to find the true minima on the potential energy surface and did not show any imaginary frequency for optimized nanocrystals. Single point energy calculations were carried on the optimized geometry at MP2/6-311++G** level.20,21 The same package was used to calculate optical absorption spectra using time-dependent density-functional theory (TDDFT) at B3LYP/6-31G* level. For each structure, 150 singlet excited states were computed for the absorption spectra. Natural population analysis (NPA) was conducted using the NBO3.1 module embedded in the Gaussian package. For each donor–acceptor NBO interactions, the stabilization energy E(2) associated with electron delocalization between donor and acceptor is estimated as: |
 | (1) |
where qi is the orbital occupancy, εj, εi are diagonal elements and Fi,j is the off-diagonal NBO Fock matrix element.
Mayer bond order calculations were performed using Multiwfn 3.3.8.22 Density of states diagram analysis were done via GaussSum 3.0 program.23
The optimized structures of molecules were used as input to calculation of magnetic susceptibility Λ and nuclear independent chemical shift NICS parameters using gauge included atomic orbital, GIOA, method24 by DFT/B3LYP level and 6-311++G** basis set. Negative NICS values indicate aromaticity and positive NICS values denote antiaromaticity of the system.
3. Results and discussion
To obtain all necessary information regarding the properties of proposed Ge NCs, it is desirable to perform a complete characterization that includes structural, geometry, electronic and optical specifications. In the following subsections we will do as is desired.
3.1. Geometry, stability, DOS
The empty Ge18H12 and the endohedrally filled Ge19H12 were optimized without any symmetry constrained at B3LYP/6-31G(d) level of theory, see Fig. 1. The calculated first frequencies 94.42 cm−1, 73.38 cm−1 for Ge18H12 and Ge19H12, respectively, indicate that the clusters to be stable. In order to better understand the properties of the studied Ge NCs, we first compare their geometry and stability. As shown in Fig. 1, both Ge NCs are composed of three parallel and planer hexagons with one additional Ge atom close to the cluster center in the case of Ge19H12. Also, three types of Ge atoms can be classified; GeP and GeM are related to the peripheral and middle-layer germanium atoms, respectively, and GeC which is the central germanium atom in Ge19H12 cluster. The optimized geometrical parameters including bond distances and bond angles are summarized in Table 1. By introducing 19th Ge atom inside the cage, it is found that the GeM–GeP and GeM–GeM are slightly elongated. In contrast, GeP–GeP is decreased about 0.1 Å. Furthermore, the GeC is positioned in the center of the Ge19H12 nanotube with bond lengths of 2.512 Å connected to six neighbors GeM atoms.
Table 1 Calculated bond distances (R/Å) and angles (∠/deg) for Ge18H12 and Ge19H12 nanocrystals
Geometry |
Ge18H12 |
Ge19H12 |
R(GeM–GeP) |
2.438 |
2.532 |
R(GeM–GeM) |
2.383 |
2.512 |
R(GeP–GeP) |
2.461 |
2.399 |
R(GeP–H) |
1.552 |
1.558 |
R(GeC–GeM) |
— |
2.512 |
∠GeP–GeM–GeM |
90.9 |
88.7 |
∠H–GeP–GeP |
112.6 |
113.4 |
∠H–GeP–GeM |
131.6 |
124.9 |
∠GeP–GeP–GeP |
120.0 |
120.0 |
∠GeM–GeM–GeM |
120.0 |
120.0 |
By introducing 19th Ge atom in the Ge18H12 nanostructure, we found the optimized Ge19H12 structure that gains 4.5 eV in stability. This value is comparable with the reported 5.1 eV value by Vach for similarly silicon nanostructure.7 As we will see in below this unusual stability is related to the aromaticity of the nanocrystals as indicated by some aromaticity criterion.
To confirm the effect of the inserted 19th Ge atom on the electronic structure, density of state DOS calculations and highest occupied molecular orbitals (HOMO) and the lowest-lying unoccupied molecular orbitals (LUMO) analysis carried out for studied nanotubes, Fig. 2 and 3. The energy change between the highest occupied molecular orbital EHOMO and the lowest unoccupied molecular orbital ELUMO decreased from 2.24 eV value for Ge18H12 to 1.72 eV for Ge19H12 nanocrystals. The inserted 19th Ge atom produces a new state into the forbidden region of Ge18H12, and hence narrows the HOMO–LUMO gap. The HOMO and LUMO distributions of Ge18H12 and Ge19H12 nanotubes are illustrated in Fig. 3. In general, the HOMO represents the electron donating ability, whereas LUMO indicates electron accepting ability of the molecule, which are important in chemical reactions and optical properties of nanocrystals. The LUMO for both Ge18H12 and Ge19H12 is distributed over the molecule. On the other hand, the HOMO of Ge19H12 is mainly focused on the GeC atom and the HOMO of Ge18H12 is predominantly focused around the middle-layer germanium atoms. This HOMO feature may be related to the ability of Ge18H12 to host additional Ge atom as over-coordinated in the center.
 |
| Fig. 2 Density of states maps for (a) Ge18H12 and (b) Ge19H12. | |
 |
| Fig. 3 Molecular orbitals of Ge18H12 and Ge19H12 at B3LYP/6-31G* level. | |
3.2. NBO analysis, Mayer bond order and electron-deficiency aromaticity
In order to make the effect of inserted atom on the charges of the atoms, we calculated Natural Population Analysis (NPA) charges of the atoms in the Ge18H12 and Ge19H12, the results are listed in Table 2. For both nanostructures GeM atoms have more electron than GeP atoms, and hydrogen atoms are partially negatively charged. This phenomenon is somewhat expected, because hydrogen has a higher electronegativity than germanium. By insertion of a Ge atom into the center of the Ge18H12 cluster, while each peripheral germanium atoms become only 0.018ē more negative; conversely, each middle-layer germanium atoms become 0.167ē more positive. Natural charge calculations also indicate that GeC atom in Ge19H12 serves as the negative charge center by −0.778ē, to keep the overall charge balance.
Table 2 NPA charges of the qH, peripheral silicon atoms qGe-peri, middle-layer silicon atoms qGe-mid and the inserted atom qGe-ins of Ge18H12 and Ge19H12, the Mayer bond orders of Ge–H and G–Ge bonds
|
qH |
qGe-peri |
qGe-mid |
qGe-ins |
GeP–H |
GeP–GeP |
GeP–GeM |
GeM–GeM |
GeM–GeC |
|
Ge18H12 |
−0.091 |
0.163 |
−0.143 |
— |
0.889 |
0.861 |
0.789 |
0.862 |
— |
|
Ge19H12 |
−0.092 |
0.145 |
0.024 |
−0.778 |
0.886 |
0.897 |
0.660 |
0.602 |
0.492 |
|
Mayer bond order25 is a quantity frequently used to study multiplicity of bonds and compare strength of the same kind of bonds. To study the characteristics of the Ge–Ge and Ge–H bonds and to reveal the influence of atomic insertion, we computed Mayer bond order for all Ge–Ge bonds and listed them in Table 2. Calculated bond orders are smaller than 1.0 and in the range of 0.492–0.897, which suggest that, less than one pair of electrons is involved in each of these bonds. This observation is in line with ref. 9, which showed that most Si–Si and Si–Ge bonds in Si19H12 and Si18GeH12 are electron-deficient. Based on the calculated bond order, the strengths of Ge–Ge bonds in Ge18H12 follow the trends of GeM–GeM ≅ GeP–GeP > GeM–GeP. After inserting a Ge atom in the cluster center, the bond order of GeP–GeP are not considerably changed, and bond orders of GeP–GeM and GeM–GeM are decreased by about 0.129 and 0.260 values, respectively. The bonding between inserted Ge atom and six middle-layer Ge atoms has the bond order 0.492. This value suggest GeC atom is able to bonded to GeM atoms as covalent “half-bonds” which strengthen idea of an electron deficient nanostructure.
In order to more investigate the electron deficient and detailed information regarding the electronic conjugation between the Ge–Ge bonds, the occupancies and stabilization energies E2 values of the most important donor–acceptor interactions were analyzed for Ge19H12 structure. For similar silicon nanocrystals structure, Vach demonstrated the aromatic-like properties of Si19H12 due to the strong electron delocalization.6 Owing to this phenomenon, Si19H12 shows more stability than any other known silicon nanocrystals and proposed their experimental synthesis is very likely. In the NBO method, electron delocalization between Lewis-type (donor) and non-Lewis (acceptor) corresponds to stabilizing donor–acceptor interactions. The stabilization energy E(2) associated with these interactions can be estimated by eqn (1). These NBOs results are illustrated in Fig. 4 and 5 and Table 3. The interaction between σGeP–GeP orbital as donor and
,
and
orbitals as an acceptor are depicted in Fig. 4. While natural bond orbitals σGeP–GeP (1.920ē) and σGeP–H (1.972ē) in up and down rings are almost fully electron occupied, the σGeM–GeM and σGeM–GeP bonds are founded to have considerable depletion in electron with 1.6326ē and 1.779ē, respectively. Furthermore, the two lone-pairs, LP1 (GeG) with 1.588ē and LP2 (GeG) with 1.201ē, of the center germanium atom GeG, see Fig. 5, show strong depletion and lone-pair orbitals LP3*(GeG) with 0.990ē and LP4*(GeG) with 0.9879ē are highly occupied. Large electron deficiency of σGeM–GeM bonds in the middle layer are related to their strong interactions with LP*(3) and LP*(4) of the central GeC atom, see Table 3. Two binding orbitals σGeA–GeB and σGeD–GeE in the middle layer have been interacted with LP*(4) GeG by E(2) values about 193 kcal mol−1. On the other hand, four other bonds in the middle layer have also strong interactions with both LP*(3) GeG and LP*(3) GeG, where the total resulted E(2) values are yet again about 193 kcal mol−1. The donor–acceptor interactions in the middle layer of Ge19H12 nanocrystals are shown in Fig. 5. These strong interactions indicate that electron-deficiency of Ge–Ge bonds in Ge19H12 nanocrystals are resolved by delocalizing participating electrons. This electron delocalization phenomenon is similar to what is known for aromatic compounds and Vach termed it as “electron-deficiency aromaticity”. Therefore, the results from NBO calculations suggest the strong electron delocalization caused by electron deficiency in Ge19H12 can be related to calculated stabilization energy shown in Fig. 1 by creation of aromaticity.
 |
| Fig. 4 Stabilization energies E(2) resulting from delocalization between donor σGeP–GeP and acceptor sites (a) (b) and (c) . Energies are given in kcal mol−1. | |
 |
| Fig. 5 Some of stabilization energies E(2) resulting from delocalization between donor and acceptor sites in the middle layer of Ge19H12 nanocrystals. Energies are given in kcal mol−1. | |
Table 3 Calculated occupancy and the second-order perturbation energy E(2) at B3LYP/6-31G* level for Ge19H12
Lewis-type NBOs (donor) |
Non-Lewis NBOs (acceptor) |
E(2) (kcal mol−1) |
Type |
Occupancy |
Type |
Occupancy |
Ge18H12 |
Ge19H12 |
Ge18H12 |
Ge19H12 |
σGeA–GeB |
1.897 |
1.633 |
LP*(4) GeG |
— |
0.989 |
193.12 |
σGeA–GeF |
1.897 |
1.633 |
LP*(3) GeG |
— |
0.990 |
129.37 |
σGeA–GeF |
1.897 |
1.633 |
LP*(4) GeG |
— |
0.989 |
65.37 |
σGeB–GeC |
1.897 |
1.633 |
LP*(3) GeG |
— |
0.990 |
160.98 |
σGeB–GeC |
1.897 |
1.633 |
LP*(4) GeG |
— |
0.989 |
33.91 |
σGeC–GeD |
1.897 |
1.633 |
LP*(3) GeG |
— |
0.990 |
129.3 |
σGeC–GeD |
1.897 |
1.633 |
LP*(4) GeG |
— |
0.989 |
65.17 |
σGeD–GeE |
1.897 |
1.633 |
LP*(4) GeG |
— |
0.989 |
192.64 |
σGeE–GeF |
1.897 |
1.633 |
LP*(3) GeG |
— |
0.990 |
160.72 |
σGeE–GeF |
1.897 |
1.633 |
LP*(4) GeG |
— |
0.989 |
33.67 |
3.3. NICS and aromaticity
As we discussed in previous section, introducing one Ge atom in the center of Ge18H12 leading to stable overcoordinated Ge19H12 with strong electron delocalization and aromatic like properties. Due to the presence of delocalized electrons in the germanium cluster studied here, aromaticity criterion is an interesting theoretical property and highly worthy to be investigated. Nucleus independent chemical shifts NICS, proposed by Schleyer and co-workers have become the most popular index for measuring aromaticity due to its simplicity and efficiency.26,27 The negative isotropic chemical shielding −σiso at some selected point is defined as NICS. Negative NICS values indicate aromaticity and positive NICS values denote antiaromaticity of the system. In some literature, it was shown that NICSZZ is a better index than the original NICS.28 Actually, NICS is an average of its components NICSXX, NICSYY and NICSZZ. However, in this study NICSZZ component and NICS were considered to evaluate the aromaticity of the studied nanocrystals. For cylindrical clusters such as Ge18H12 and Ge19H12, a more thorough way to characterize aromaticity may be plotting the NICS curve along the symmetry axis of the cluster. The NICS results, calculated at the B3LYP/6-311++G** level of theory, along the longitudinal direction in the center of the Ge18H12 and Ge19H12 nanoparticle are shown in Fig. 6. For comparison, the NICS curves, obtained at the same level, of benzene (the reference of aromatic species) and Si19H12 are also given in Fig. 6. Notice that the occurrence of very negative values in the middle region of the NICS curve of Si19H12 and Ge19H12 is because the sample points exactly cross the inserted atom, whose inner-core electrons shield external magnetic fields significantly. However, this does not hamper fair comparison of aromaticity, as long as the middle region of the NICS curves is ignored. As it can be seen from this Fig. 6, Ge18H12 with negative NICSZZ index values show aromaticity character and antiaromaticity property from positive NICS index values around the cage center. The positive NICS values are caused by the positive NICSXX and NICSYY components and thus one may conclude that this cluster possesses a weak antiaromaticity character. However, after inserting one Ge atom into the Ge18H12 cluster center, the aromaticity becomes significant and even stronger than benzene and slightly more than Si19H12.
 |
| Fig. 6 (a) NICS (b) NICSZZ values along the longitudinal distance from the center. | |
The magnetic susceptibility Λ resulting from the presence of cyclic delocalization of electrons is another important aromaticity criterion which is used here. Unlike NICS parameters which are local properties, the magnetic susceptibility is a global property of the molecule. The more negative values of magnetic susceptibility Λ exhibit more aromaticity character of the molecule. Magnetic susceptibility Λ of Ge18H12 and Ge19H12 nanocrystals are calculated to be −321.2 cgs ppm and −428.4 cgs ppm, increasing the aromaticity property of about 100 cgs ppm from empty to overcoordinated Ge NCs. The corresponding magnetic susceptibility value for benzene was also obtained Λ (C6H6) = −56.4 cgs ppm, which indicate that the global aromaticity properties of studied nanocrystals is significantly larger than benzene.
3.4. Absorption spectra
In our previous work, it has been shown that due to the insertion of the center silicon atom to the silicon nanocrystals Si18X12 (X = H, F, Cl, Br, NH2, COOH, OH) the absorption spectrum are extended from UV-VIS region to infrared IR region.9 Likewise, we obtained the adsorption spectrum of studied Ge18H12 and Ge19H12 nanocrystals and are plotted in Fig. 7. While optical absorption spectrum of Ge18H12 is only limited in UV region by wavelengths lower than 400 nm, the overcoordinated germanium nanocrystals Ge19H12 can absorb light in the three spectral UV-VIS and IR regions. Two nanocrystals exhibit similar spectrum for excitation energies with wavelengths < 400 nm, but Ge19H12 show additional features in the spectrum with strong peaks in the range of 800–850 nm. Thus, this finding is similar to the previous results of silicon nanocrystals. In general, standard tetrahedral clusters of this size show absorption only in the UV and may absorb light in the VIS and IR only when toxic or expensive metal are incorporated.29–32 Such an extension of spectral response by overcoordinated germanium nanocrystals might it be used to the environmentally safe replacement of other expensive or toxic elements in nanocrystals with application in solar cell and electronic devices. In particular the strong adsorption at interval 800–850 nm may open potential applications in cancer treatment where the use of gold and silver nanoparticles has been proposed where the resulting laser light absorption at about 800 nm yields optimal efficiency.33,34 If successfully synthesized, the experimental electronic properties derived for Ge19H12 structure are expected to be coherent with the data derived in this text.
 |
| Fig. 7 Absorption spectra of (a) Ge18H12 and (b) Ge19H12. | |
4. Conclusions
In the present work, DFT calculations have been indicated that Ge18H12 nanocrystals can by more stabilized by insertion of one Ge atom in the center and makes overcoordinated Ge19H12 nanocrystals. We extend the definition of the electron-deficiency aromaticity concept to Ge nanocrystals by exploiting the electron-deficiency of a system to induce electron delocalization and, thus, enhanced stability. The NICS and magnetic susceptibility aromaticity indexes show that Ge19H12 nanocrystals exhibit greater aromaticity than conjugated organic molecules such as benzene. The NBO analysis have demonstrated that Ge–Ge bonds in Ge19H12 have strongly delocalized nature and are strongly interacted with the lone pair of central Ge atom. It has shown that pure hydrogenated germanium nanocrystals Ge19H12 with overcoordinated ring structures can absorb light not only in the ultraviolet, but also in the visible and infrared spectral region without the need for expensive or toxic metal atoms. This new pattern for optical spectrum of Ge19H12 can be attributed to the special electronic structure of studied nanoparticles. Finally, given the technological importance of aromatic derivatives, we believe that the proposed Ge NCs with aromatic-like behavior will have a significant technological impact and deserves further extensive study in various emerging nanotechnological domains.
Acknowledgements
We gratefully acknowledge discussions with Holger Vach. Allotment of computer time provided by David Vander Spoel from the Biomedical Center in Uppsala University is gratefully acknowledged. The authors also thank Kharazmi University for financial support.
References
- X. Li, A. E. Kuznetsov, H. F. Zhang, A. I. Boldyrev and L. S. Wang, Observation of All-Metal Aromatic Molecules, Science, 2001, 291, 859–861 CrossRef CAS PubMed.
- N. G. Szwacki, V. Weber and C. J. Tymczak, Aromatic Borozene, Nanoscale Res. Lett., 2009, 4, 1085–1089 CrossRef PubMed.
- R. Islas, T. Heine, K. Ito, P. v. R. Schleyer and G. Merino, Boron Rings Enclosing Planar Hypercoordinate Group 14 Elements, J. Am. Chem. Soc., 2007, 129, 14767–14774 CrossRef CAS PubMed.
- J. O. C. Jimenez-Halla, E. Matito, J. Robles and M. Sola, Nucleus-independent chemical shift (NICS) profiles in a series of monocyclic planar inorganic compounds, J. Organomet. Chem., 2006, 691, 4359–4366 CrossRef CAS.
- T. R. Galeev and A. I. Boldyrev, Aromaticity and Antiaromaticity in Inorganic Chemistry, Reference Module in Chemistry, Molecular Sciences and Chemical Engineering Comprehensive Inorganic Chemistry II, From Elements to Applications, 2nd edn, 2013, pp. 245–275 Search PubMed.
- H. Vach, Electron-Deficiency Aromaticity in Silicon Nanoclusters, J. Chem. Theory Comput., 2012, 8, 2088–2094 CrossRef CAS PubMed.
- H. Vach, Ultrastable Silicon Nanocrystals due to Electron Delocalization, Nano Lett., 2011, 11, 5477–5481 CrossRef CAS PubMed.
- H. Vach, Terahertz and Gigahertz emission from an all silicon nanocrystal, Phys. Rev. Lett., 2014, 112, 197401–197405 CrossRef PubMed.
- S. Manzetti, T. Lu, H. Behzadi, M. D. Esrafili, H.-L. Thi Le and H. Vach, Intriguing properties of unusual silicon nanocrystals, RSC Adv., 2015, 5, 78192–78208 RSC.
- C. Bonafos, M. Carrada, G. Benassayag, S. Schamm-Chardon, J. Groenen, V. Paillard, B. Pecassou, A. Claverie, P. Dimitrakis, E. Kapetanakis, V. Ioannou-Sougleridis, P. Normand, B. Sahu and A. Slaoui, Si and Ge nanocrystals for future memory devices, Mater. Sci. Semicond. Process., 2012, 15, 615–626 CrossRef CAS.
- V. Svrcek, S. Cook, S. Kazaoui and M. Kondo, Silicon Nanocrystals and Semiconducting Single-Walled Carbon Nanotubes Applied to Photovoltaic, J. Phys. Chem. Lett., 2011, 2(14), 1646–1650 CrossRef CAS.
- H. C. Weissker, N. Ning, F. Bechstedt and H. Vach, Luminescence and absorption in germanium and silicon nanocrystals: The influence of compression, surface reconstruction, optical excitation, and spin-orbit splitting, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83, 125413–125416 CrossRef.
- M. B. Javan, Ge–Si and Si–Ge core–shell nanocrystals: Theoretical study, Thin Solid Films, 2015, 589, 120–124 CrossRef CAS.
- N. C. Forero-Martinez, H.-L. Thi Le, N. Ning, H. Vach and H. C. Weissker, Temperature dependence of the radiative lifetimes in Ge and Si nanocrystals, Nanoscale, 2015, 7, 4942–4948 RSC.
- M. J. Jawad, M. R. Hashim, N. K. Ali, E. P. Corcoles and V. K. Arora, Photoluminescence of Ultraviolet Initiated Green Emission from Electrochemically Deposited Germanium Films on (100) Silicon, J. Electrochem. Soc., 2014, 161, D801–D805 CrossRef CAS.
- M. F. Jarrold and J. E. Bower, Mobilities of silicon cluster ions: The reactivity of silicon sausages and spheres, J. Chem. Phys., 1992, 96, 9180–9190 CrossRef CAS.
- M. J. Frisch, G. W. Trucks, H. B. Schlegel, H. B. Scuseria, M. A. Robb, J. R. Cheeseman, V. G. Zakrzewski, J. A. Montgomery, R. E. Stratmann, J. C. Burant, S. Dapprich, J. M. Millam, A. D. Daniels, K. N. Kudin, M. C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R. Cammi, C. Mennucci, C. Pomelli, C. Adamo, S. Clifford, J. Ochterski, G. A. Petersson, P. Y. Ayala, Q. Cui, K. Morokuma, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. Cioslowski, J. V. Ortiz, A. G. Baboul, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. Gomperts, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, C. Gonzalez, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, J. L. Andres, J. Gonzalez, M. Head-Gordon, E. S. Replogle and J. A. Pople, Gaussian 03, 2003 Search PubMed.
- C. Lee, W. Yang and R. G. Parr, Development of the Colle–Salvetti Correlation-Energy Formula into a Functional of the Electron Density, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789 CrossRef CAS.
- R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople, Self-Consistent Molecular Orbital Methods. Xx. A Basis Set for Correlated Wave Functions, J. Chem. Phys., 1980, 72, 650–654 CrossRef CAS.
- C. Moller and M. S. Plesset, Note on an Approximation Treatment for Many-Electron Systems, Phys. Rev., 1934, 46, 618–622 CrossRef CAS.
- M. J. Frisch, J. A. Pople and J. S. Binkley, “Self-Consistent Molecular Orbital Methods 25: Supplementary Functions for Gaussian Basis Sets,”, J. Chem. Phys., 1984, 80, 3265–3269 CrossRef CAS.
- T. Lu and F. Chen, Multiwfn: A Multifunctional Wavefunction Analyzer, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed.
- N. M. OBoyle, A. L. Tenderholt and K. M. Langner, cclib: A Library for Package-Independent Computational Chemistry Algorithms, J. Comput. Chem., 2008, 29, 839–845 CrossRef CAS PubMed.
- K. Wolinski, J. F. Hilton and P. Pulay, Efficient Implementation of the Gauge-Independent Atomic Orbital Method for NMR Chemical Shift Calculations, J. Am. Chem. Soc., 1990, 112, 8251–8260 CrossRef CAS.
- I. Mayer, Charge, Bond Order and Valence in the Ab Initio Scf Theory, Chem. Phys. Lett., 1983, 97, 270–274 CrossRef CAS.
- P. V. R. Schleyer, C. Maerker, A. Dransfeld, H. Jiao and N. J. R. v. Eikema Hommes, Nucleus-Independent Chemical Shifts: A Simple and Efficient Aromaticity Probe, J. Am. Chem. Soc., 1996, 118, 6317 CrossRef CAS.
- Z. Chen, C. S. Wannere, C. Corminboeuf, R. Puchta and P. V. R. Schleyer, Nucleus-Independent Chemical Shifts (NICS) as an Aromaticity Criterion, Chem. Rev., 2005, 105, 3842–3888 CrossRef CAS PubMed.
- H. F. B. Shaidaei, C. S. Wannere, C. Corminboeuf, R. Puchta and P. v. R. Schleyer, Which NICS Aromaticity Index for Planar π Rings Is Best?, Org. Lett., 2006, 8, 863–866 CrossRef PubMed.
- A. J. Nozik, Nanoscience and Nanostructures for Photovoltaics and Solar Fuels, Nano Lett., 2010, 10, 2735–2741 CrossRef CAS PubMed.
- K. M. Noone, E. Strein, N. C. Anderson, P.-T. Wu, S. A. Jenekhe and D. S. Ginger, Broadband Absorbing Bulk Heterojunction Photovoltaics Using Low-Bandgap Solution-Processed Quantum Dots, Nano Lett., 2010, 10, 2635–2639 CrossRef CAS PubMed.
- H. Lee, M. Wang, P. Chen, D. R. Gamelin, S. M. Zakeeruddin, M. Gratzel and M. K. Nazeeruddin, Efficient CdSe Quantum Dot-Sensitized Solar Cells Prepared by an Improved Successive Ionic Layer Adsorption and Reaction Process, Nano Lett., 2009, 9, 4221–4227 CrossRef CAS PubMed.
- D. Palagin and K. Reuter, Evaluation of Endohedral Doping of Hydrogenated Si Fullerenes as a Route to Magnetic Si Building Blocks, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 86, 045416 CrossRef.
- A. Csaki, F. Garwe, A. Steinbrück, G. Maubach, G. Festag, A. Weise, I. Riemann, K. Konig and W. Fritzsche, A Parallel Approach for Subwavelength Molecular Surgery Using Gene-Specific Positioned Metal Nanoparticles as Laser Light Antennas, Nano Lett., 2007, 7, 247–253 CrossRef CAS PubMed.
- J. Yang, J. Choi, D. Bang, E. Kim, E. K. Lim, H. Park, J. S. Suh, K. Lee, K. H. Yoo and E. K. Kim, Convertible Organic Nanoparticles for near-Infrared Photothermal Ablation of Cancer Cells, Angew. Chem., 2011, 123, 461–464 CrossRef.
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.