Next Article in Journal
Threshold Switching in Forming-Free Anodic Memristors Grown on Hf–Nb Combinatorial Thin-Film Alloys
Next Article in Special Issue
Lanthanide(III) Ions and 5-Methylisophthalate Ligand Based Coordination Polymers: An Insight into Their Photoluminescence Emission and Chemosensing for Nitroaromatic Molecules
Previous Article in Journal
Graphene-Based Adsorbents for Arsenic, Fluoride, and Chromium Adsorption: Synthesis Methods Review
Previous Article in Special Issue
Electronic Transport in Weyl Semimetals with a Uniform Concentration of Torsional Dislocations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electronic Properties of Hexagonal Graphene Quantum Rings from TAO-DFT

1
Department of Physics, National Taiwan University, Taipei 10617, Taiwan
2
Center for Theoretical Physics and Center for Quantum Science and Engineering, National Taiwan University, Taipei 10617, Taiwan
3
Physics Division, National Center for Theoretical Sciences, Taipei 10617, Taiwan
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(22), 3943; https://doi.org/10.3390/nano12223943
Submission received: 21 October 2022 / Revised: 4 November 2022 / Accepted: 6 November 2022 / Published: 9 November 2022
(This article belongs to the Special Issue Theoretical Calculation and Molecular Modeling of Nanomaterials)

Abstract

:
The reliable prediction of electronic properties associated with graphene nanosystems can be challenging for conventional electronic structure methods, such as Kohn–Sham (KS) density functional theory (DFT), due to the presence of strong static correlation effects in these systems. To address this challenge, TAO (thermally assisted occupation) DFT has been recently proposed. In the present study, we employ TAO-DFT to predict the electronic properties of n-HGQRs (i.e., the hexagonal graphene quantum rings consisting of n aromatic rings fused together at each side). From TAO-DFT, the ground states of n-HGQRs are singlets for all the cases investigated (n = 3–15). As the system size increases, there should be a transition from the nonradical to polyradical nature of ground-state n-HGQR. The latter should be intimately related to the localization of active TAO-orbitals at the inner and outer edges of n-HGQR, which increases with increasing system size.

1. Introduction

Since the isolation and characterization of graphene by Geim and Novoselov in 2004 [1,2], graphene has received considerable research attention. Graphene has been found to exhibit interesting properties [3,4,5,6] (e.g., the properties of graphene can be changed by modulating as composite materials [7,8,9]) which can be useful for many technology applications, such as transistors [10,11], batteries [12,13,14], quantum computers [15,16], biocatalysis [17], memory devices and photothermal applications [18]. Graphene is a two-dimensional material constructed of a sequential honeycomb lattice of carbon atoms, each forming s p 2 hybridization bonding with three other carbon atoms. Leaving one delocalized free electron which forms a π -orbital, yielding fascinating properties, such as the quantum Hall effect [19,20,21] and thermal and electronic conductivity [22,23,24]. However, graphene has zero band gap, which can be an obstacle for some electronic applications. Many studies have shown that confining the size of graphene to the nanometer scale, such as graphene nanoribbons (GNRs) [25,26,27,28,29,30], graphene quantum dots (GQDs) [31,32,33,34,35,36,37,38], and graphene quantum rings (GQRs) [34,36,37,39,40,41,42,43], could open the band gap because of finite-size effects [30]. Nonetheless, the properties of graphene nanosystems with different edges, shapes, and sizes can be distinctly different, requiring comprehensive theoretical and experimental studies.
Due to the improvement in graphene fabrication, GQRs have attracted increasing attention in recent years. Note that GQRs are closely related to GQDs, and can be regarded as GQDs with additional inner edges. It is expected that the inter-edge coupling of states and enhanced finite-size effects in GQRs could yield novel properties. Indeed, some interesting properties of GQRs, such as the Aharonov–Bohm (AB) effects [42], have been theoretically predicted [39] and experimentally observed [34]. Further theoretical studies have focused on the AB effects on the persistent currents in GQRs [44] as well as the interplay of the AB effects and Klein tunneling in GQRs [45]. In addition, other theoretical studies have revealed the importance of edge and geometry effects on the energy spectra of GQRs in a magnetic field [36,43].
Among various GQRs, in this study, we focus on the hexagonal GQRs, consisting of n aromatic rings fused together at each side, which are denoted as n-HGQRs. As shown in Figure 1, n-HGQR (with the chemical formula C 24 ( n 1 ) H 12 ( n 1 ) ), which contains N = 156 ( n 1 ) electrons, has six-fold rotational symmetry. To date, there have been scarce experimental studies on n-HGQR, which may be due to the fact that it remains very difficult to realize the precise n-HGQRs with smooth edges. On the theoretical side, there have been only few studies on the properties of energy spectra associated with n-HGQRs in the presence of a magnetic field [46,47].
Owing to the presence of strong static correlation in the larger n-HGQRs (which will be shown and discussed later), conventional electronic structure methods may not be appropriate. For example, Kohn–Sham (KS) density functional theory (DFT) [48,49] using the common exchange-correlation (XC) energy functionals (e.g., the LDA (local density approximation), GGA (generalized-gradient approximation), hybrid as well as double-hybrid XC functionals) have been commonly applied to predict the properties of electronic systems without strong static correlation (i.e., electronic systems with a single-reference (SR) character or SR systems), while KS-DFT with the common XC functionals can provide unreliable predictions of the properties of electronic systems with strong static correlation (i.e., electronic systems with a multi-reference (MR) character or MR systems). On the other hand, although high-level ab-initio MR electronic structure methods [50,51,52] are expected to take care of strong static correlation, accurate MR methods are known to be computationally expensive, and, thus, inapplicable to nanosystems, such as the larger n-HGQRs.
Recently, thermally assisted occupation (TAO) DFT, which is a DFT using fractional TAO-orbital occupation numbers (obeying the Fermi–Dirac (FD) distribution with a fictitious temperature θ ) has been developed to describe the properties of nanosystems with strong static correlation [53]. Unlike finite-temperature DFT (FT-DFT) [54] (i.e., a DFT for studying the thermal equilibrium properties associated with systems at finite electronic temperatures; FT-DFT has been powered by the Mermin–Kohn–Sham (MKS) equations (see Section III of Ref. [49]), and the resulting MKS theory [49,54] is reduced to KS-DFT at zero electronic temperature (absolute zero)), TAO-DFT is a DFT for studying the ground-state properties associated with systems at absolute zero (even with a nonzero fictitious temperature θ , i.e., a parameter closely dependent on the strength of the static correlation of an electronic system at absolute zero). In the framework of TAO-DFT, the static correlation energy associated with the ground state of an electronic system is treated explicitly (though approximately) by the entropy contribution, and the corresponding natural orbital occupation numbers (NOONs) are approximately given by the TAO-orbital occupation numbers (TOONs) [53]. For the limit case of θ = 0 , TAO-DFT is reduced to KS-DFT. It is worth mentioning that TAO-DFT maintains a similar computational complexity as KS-DFT, while TAO-DFT can outperform KS-DFT for a diverse range of MR systems (even with the simplest LDA functional).
Moreover, a number of important extensions of TAO-DFT, such as the GGA [55] and hybrid [56,57] functionals in TAO-DFT, a simple model for defining the optimal system-independent θ in TAO-DFT [58], a self-consistent scheme for determining the optimal system-dependent θ in TAO-DFT [59], ab initio molecular dynamics based on TAO-DFT [60], and an approach for extracting excited-state properties within the TAO-DFT framework [61], have been recently developed. In addition, TAO-DFT has been successfully applied to predict the electronic properties [62,63,64,65,66,67,68,69,70,71,72,73,74,75], hydrogen storage properties [64,66,67], and spectroscopic properties [60,76,77] of various MR nanosystems in recent years.
As a result, in this work, we use TAO-DFT to explore the electronic properties (e.g., the singlet–triplet gaps, fundamental gaps and vertical electron affinities/ionization potentials, symmetrized von Neumann entropy as well as the occupation numbers and visualization of active TAO-orbitals) of n-HGQRs (n = 3–15). One of the objectives of this study is to address whether there is a transition from the nonradical to polyradical nature of ground-state n-HGQR with increasing system size, which has never been reported, although such a transition has been previously found in hexagonal GQDs [68] (i.e., closely related electronic systems). The significance of this work is to address whether unconventional electronic structure methods, such as TAO-DFT, are necessary for a reliable prediction of electronic properties of n-HGQRs.

2. Computational Details

We carried out all numerical calculations with the 6-31G(d) basis set on the software of Q-Chem 4.4 [78]. Geometry optimizations (see Supplementary Information (SI)) as well as single-point energy calculations were carried out using TAO-LDA, which is TAO-DFT [53] with the LDA XC and θ -dependent functionals, wherein θ = 7 mhartree (i.e., an optimal system-independent θ value for TAO-LDA calculations on both SR and MR systems) [53] was adopted. For comparison purposes, we also present the results of KS-LDA (i.e., KS-DFT using the LDA XC functional or TAO-LDA with θ = 0 ).

3. Results and Discussion

3.1. Singlet–Triplet Gap

To obtain the ground state of n-HGQR, we resorted to the singlet–triplet (ST) gap. First, we carried out KS-LDA and TAO-LDA calculations for the lowest spin-unrestricted triplet energy ( E UT ) and lowest spin-unrestricted singlet energy ( E US ) of n-HGQR at the respective optimized geometries. Subsequently, we computed the ST gap ( E ST ) of n-HGQR using
E ST = E UT E US .
As shown in Figure 2, TAO-LDA predicts that n-HGQRs (n = 3–15) (i.e., all the cases studied) should possess singlet ground states (see also Table S1 and Figure S1 (for additional information on the singlet–quintet gap of n-HGQR, calculated by spin-unrestricted TAO-LDA) in SI). Similarly, KS-LDA also predicts that n-HGQRs (n = 3–14) should have singlet ground states. However, the ST gap of TAO-LDA monotonically decreases with the n-HGQR size, while the ST gap of KS-LDA displays an unexpected increase beyond 11-HGQR.
To explore the possible reasons for discrepancies, we reveal the expectation values of the total spin-squared operator S ^ 2 associated with the lowest singlet and lowest triplet states of n-HGQR, computed using spin-unrestricted KS-LDA (see Table 1). If the KS-LDA wavefunction has spin contamination (which is the artificial mixing of different spin-states) [68,71,79], the corresponding S ^ 2 value will be larger than the exact S ^ 2 value (i.e., 0 for the lowest singlet state and 2 for the lowest triplet state; e.q., refer to Section I in the SI of Ref. [79]). Consequently, the difference between the S ^ 2 value of KS-LDA and the exact S ^ 2 value is commonly used to assess the degree of spin contamination in the KS-LDA wavefunction. As shown, for the smaller n-HGQRs, the S ^ 2 values of KS-LDA are rather close to the exact S ^ 2 values, indicating that the lowest singlet and lowest triplet states of n-HGQR, obtained with spin-unrestricted KS-LDA, essentially have no spin contamination. Nevertheless, for the larger n-HGQRs (e.g., n 11 ), the S ^ 2 values of KS-LDA are much larger than the exact S ^ 2 values, indicating that the lowest singlet and lowest triplet states of n-HGQR, obtained with spin-unrestricted KS-LDA, are highly spin-contaminated. Therefore, the larger n-HGQRs (e.g., n 11 ) are expected to possess a pronounced MR character in the lowest singlet and lowest triplet states, and the unexpected increase in the ST gap of KS-LDA beyond 11-HGQR can be an artifact intimately correlated with spin contamination.
In addition, we also examine the consequences of spin contamination in spin-unrestricted KS-LDA/TAO-LDA calculations for the lowest singlet states of n-HGQRs. Owing to the constraint of spin symmetry in the lowest singlet state of n-HGQR, for the exact theory, the lowest singlet energy of n-HGQR computed using the spin-restricted formalism must be the same as that computed using the spin-unrestricted formalism [53,55,56]. Accordingly, we carried out KS-LDA and TAO-LDA calculations for the lowest spin-restricted singlet energy ( E RS ) and lowest spin-unrestricted singlet energy ( E US ) of n-HGQR at the corresponding optimized geometries. Subsequently, we computed the E UR value of n-HGQR using
E UR = E RS E US .
As presented in Table 2, the E UR values of the larger n-HGQRs (e.g., n 11 ), calculated by KS-LDA, are very large (e.g., more than 1.6 kcal/mol), leading to unphysical spin-symmetry breaking effects in the respective spin-unrestricted KS-LDA calculations. By contrast, the E UR values of n-HGQRs (n = 3–15) (i.e., all the cases examined), calculated by TAO-LDA, remain essentially zero, yielding essentially no unphysical spin-symmetry breaking effects in the respective spin-unrestricted TAO-LDA calculations.
In short, the ST gaps associated with the larger n-HGQRs (e.g., n 11 ), calculated by KS-LDA, can be seriously affected by spin contamination, greatly degrading the accuracy of KS-LDA in predicting the ST gaps and possibly other electronic properties of the larger n-HGQRs (e.g., n 11 ). As will be shown later, the larger n-HGQRs possess an increasing polyradical nature in the ground states (i.e., the lowest singlet states), wherein KS-LDA can provide unreliable electronic properties. Therefore, we only report the results of TAO-LDA hereafter.

3.2. Fundamental Gap and Vertical Electron Affinity/Ionization Potential

Another important electronic property is the fundamental gap, which is known to vanish for metals, but is positive for semiconductors and insulators [80]. Here, we employ spin-unrestricted TAO-LDA to calculate the fundamental gap ( E g ) of ground-state n-HGQR (containing N electrons):
E g = IP v EA v ,
which is the difference between the IP v (vertical ionization potential) and EA v (vertical electron affinity) of ground-state n-HGQR. To obtain the IP v / EA v of ground-state n-HGQR, we perform spin-unrestricted TAO-LDA calculations with one electron removed from/added to neutral n-HGQR at the ground-state geometry, and compute the following energy difference [55,56]:
IP v = E N 1 E N ,
and
EA v = E N E N + 1 ,
where E N refers to the total energy associated with the N-electron system.
As shown in Figure 3, with the increase in system size n, the E g of ground-state n-HGQR is monotonically decreasing, which is a consequence of the monotonically decreasing IP v and monotonically increasing EA v of ground-state n-HGQR (see, also, Table S2 in SI). Moreover, the fundamental gaps of ground-state n-HGQRs (n = 5–10) are in the range of 1–3 eV (i.e., for the ideal energy gaps), which shows great promise for the applications of ground-state n-HGQRs (n = 5–10) in nanophotonics.

3.3. Symmetrized von Neumann Entropy

To assess the strength of MR character associated with ground-state n-HGQR, we adopted spin-unrestricted TAO-LDA to calculate the S vN (symmetrized von Neumann entropy) associated with ground-state n-HGQR [55,56]:
S vN = 1 2 σ = , i = 1 f i , σ ln ( f i , σ ) + ( 1 f i , σ ) ln ( 1 f i , σ ) .
Here, f i , σ (a value in the range of 0–1) is the occupation number associated with the i th TAO-orbital of σ -spin (i.e., up- or down-spin), which is intimately correlated with the respective natural spin-orbital occupation number [53,55,56]. For an SR system, the f i , σ values associated with all TAO-spin-orbitals must be very close to either 0 or 1, yielding a very small value of S vN . However, for an electronic system with pronounced MR character, the f i , σ values associated with the active TAO-spin-orbitals (i.e., the TAO-spin-orbitals with significant fractional occupations (e.g., 0.1–0.9)) can strongly deviate from the values of 0 and 1; hence, the corresponding value of S vN can noticeably increase when the f i , σ values associated with the active TAO-spin-orbitals are closer to 1 2 , and/or the number of active TAO-spin-orbitals increases. As a consequence, the value of S vN associated with ground-state n-HGQR can be adopted to assess the strength of MR character associated with ground-state n-HGQR. It is worth mentioning that the strength of MR character associated with ground-state n-HGQR may also be explored by other measures [81,82], due to the availability of TOONs (i.e., approximate NOONs) in TAO-DFT.
As shown in Figure 4, the S vN of ground-state n-HGQR, which remains very small for a small value of n, is monotonically increasing with the increase in n-HGQR size (see Table S2 in SI as well). This implies that the smaller ground-state n-HGQR should exhibit an SR character (i.e., nonradical nature), and the larger ground-state n-HGQR should possess an MR nature whose strength generally increases with increasing system size n.
Finally, we briefly comment on the fractional occupation number weighted density (FOD) method [83], which has been recently proposed for the real-space measure of strong static correlation effects at absolute zero. In the FOD method, the special density ρ FOD is actually computed using an approximate TAO-DFT method [53] (i.e., TAO-DFT without the θ -dependent functional) at some fictitious temperature θ , instead of using an approximate MKS method [49,54] (i.e., the MKS theory with the XC free energy functional being approximated with the XC functional) at finite electronic temperature. This can be clearly seen from the distinctly different physical meanings of TAO-DFT [53] and the MKS theory [49,54] (which is reduced to KS-DFT at absolute zero). In other words, the fractional orbital occupation numbers adopted in the FOD method are approximate TOONs, and, hence, can be viewed as approximate NOONs. Accordingly, the N FOD value (obtained from the integration of ρ FOD over all space), which is similar to the aforementioned S vN value, can also be used to measure the strength of the MR character associated with the ground state of an electronic system at absolute zero (even with a very large fictitious temperature θ ), rather than that associated with an electronic system in thermal equilibrium at a finite electronic temperature.

3.4. Active TAO-Orbital Occupation Numbers

As mentioned previously, the NOONs associated with the ground state of an electronic system are approximately given by the TOONs in TAO-DFT [53,55,56]. To provide further insight into the increase in S vN with increasing n-HGQR size and to understand the radical nature of the larger ground-state n-HGQR, we disclose the active TOONs of ground-state n-HGQR, obtained from spin-restricted TAO-LDA. For ground-state n-HGQR (containing N electrons), we denote the HOMO (highest occupied molecular orbital)/LUMO (lowest unoccupied molecular orbital) as the ( N / 2 ) th / ( N / 2 + 1 ) th TAO-orbital, and so forth. For brevity, we denote HOMO/LUMO as H/L. In addition, the TAO-orbitals with an orbital occupation number in the range of 0.2–1.8 (i.e., the TAO-spin-orbitals with an orbital occupation number in the range of 0.1–0.9) are viewed as the active TAO-orbitals.
As shown in Figure 5, the active TOONs of ground-state n-HGQR display peculiar six-member patterns (i.e., the active TOONs are arranged to form groups of six members each), which can be related to the six-fold rotational symmetry of n-HGQR. When the n-HGQR size is small (e.g., n < 7 ), all the TOONs are close to either 0 or 2. Therefore, the smaller ground-state n-HGQRs (e.g., n < 7 ) should possess a nonradical nature, showing consistency with the investigation of the other properties, such as the larger E ST and E g values and the smaller S vN values, associated with the relatively stable n-HGQRs. Nevertheless, as the n-HGQR size increases, the active TOONs get closer to 1, and/or the number of active TAO-orbitals increases. Accordingly, the larger ground-state n-HGQRs (e.g., n 7 ) should possess an increasingly polyradical nature, also showing consistency with the investigation of the other properties, such as the smaller E ST and E g values and the larger S vN values, associated with the relatively unstable n-HGQRs. As the system size grows, there is a transition from the nonradical to polyradical nature of ground-state n-HGQR, leading to the aforementioned increase in S vN .

3.5. Real-Space TAO-Orbital Representation

To further illustrate the transition from the nonradical to polyradical nature of ground-state n-HGQR in real space, we plot the visualization of active TAO-orbitals (H − 5, H − 4, …, H − 1, H, L, L + 1, …, L + 4, and L + 5) of ground-state n-HGQRs (n = 3, 5, 7, and 9), calculated by spin-restricted TAO-LDA. As shown in Figure 6, Figure 7, Figure 8 and Figure 9, when the n-HGQR size is small (e.g., n < 7 ), the active TAO-orbitals are delocalized over the entire n-HGQR. Nonetheless, as the n-HGQR size increases, the active TAO-orbitals have an increasing tendency to localize at the inner and outer edges of n-HGQR. Consequently, the active TAO-orbitals of the larger ground-state n-HGQR (e.g., n 7 ) are mainly localized at the two edges of n-HGQR.
Similar to the previous findings for other graphene nanosystems (e.g., linear acenes [62], zigzag GNRs [62], cyclacenes [65], Möbius cyclacenes [69], and hexagonal GQDs [68]), the increasing polyradical character associated with the larger ground-state n-HGQR (e.g., n 7 ) should be intimately related to the localization of active TAO-orbitals at the inner and outer edges of n-HGQR, which increases as the system size n increases.

4. Conclusions

In conclusion, the MR character of graphene nanosystems has posed a great challenge to conventional electronic structure methods. The challenge has been greatly reduced by TAO-DFT. In this study, we explored the electronic properties of n-HGQRs (n = 3–15) obtained with TAO-LDA. Since the larger n-HGQRs were found to have pronounced MR character in the lowest singlet and lowest triplet states, KS-DFT using the common XC functionals can be unreliable for predicting the electronic properties of n-HGQRs. For example, the lowest singlet and triplet states of the larger n-HGQRs (e.g., n 11 ), obtained with spin-unrestricted KS-LDA, were found to be highly spin-contaminated, leading to unphysical spin-symmetry breaking effects. Although accurate MR electronic structure methods are able to describe strong static correlation effects, these methods can be computationally intractable for the larger n-HGQRs. Consequently, it is reasonably justified that we employ TAO-LDA to study the electronic properties of n-HGQRs due to a good compromise between efficiency and accuracy.
From the TAO-LDA results, n-HGQRs (n = 3–15) are all ground-state singlets. With increasing n-HGQR size, the E ST , IP v , and E g values monotonically decrease, while the EA v and S vN values monotonically increase. With increasing system size, there is a transition from the nonradical to polyradical nature of ground-state n-HGQR. The increasing polyradical character associated with the larger ground-state n-HGQR (e.g., n 7 ) should be intimately related to the localization of active TAO-orbitals at the inner and outer edges of n-HGQR, which increases with increasing system size.
As for possible future work, we plan to explore how the radical nature emerges from the evolution between n-HGQRs and hexagonal GQDs (i.e., the coronene series) [68] to see the trend forming hexagonal GQDs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12223943/s1, Cartesian coordinates (in Å) for the lowest singlet states of n-HGQRs (n = 3–15), calculated by spin-unrestricted TAO-LDA; Cartesian coordinates (in Å) for the lowest triplet states of n-HGQRs (n = 3–15), calculated by spin-unrestricted TAO-LDA; Figure S1: Singlet–quintet gap of n-HGQR, calculated by spin-unrestricted TAO-LDA; Table S1: Singlet–triplet gap E ST (in kcal/mol) of n-HGQR, calculated by spin-unrestricted KS-LDA and TAO-LDA; Table S2: Vertical ionization potential IP v (in eV), vertical electron affinity EA v (in eV), fundamental gap E g (in eV), and symmetrized von Neumann entropy S vN of ground-state n-HGQR, calculated by spin-unrestricted TAO-LDA.

Author Contributions

Conceptualization, C.-C.C. and J.-D.C.; Data curation, C.-C.C.; Formal analysis, C.-C.C.; Funding acquisition, J.-D.C.; Investigation, C.-C.C.; Methodology, C.-C.C. and J.-D.C.; Project administration, J.-D.C.; Resources, J.-D.C.; Software, J.-D.C.; Supervision, J.-D.C.; Validation, C.-C.C. and J.-D.C.; Visualization, C.-C.C.; Writing—original draft, C.-C.C.; Writing—review and editing, J.-D.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Ministry of Science and Technology of Taiwan (Grant No. MOST110-2112-M-002-045-MY3).

Data Availability Statement

The data that support the findings of this study are available from the authors upon reasonable request.

Acknowledgments

This work was supported by the Ministry of Science and Technology of Taiwan (Grant No. MOST110-2112-M-002-045-MY3), National Taiwan University, and the National Center for Theoretical Sciences of Taiwan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.E.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Katsnelson, M.I.; Grigorieva, I.; Dubonos, S.; Firsov, A. Two-dimensional gas of massless Dirac fermions in graphene. Nature 2005, 438, 197–200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Zhang, Y.; Tan, Y.W.; Stormer, H.L.; Kim, P. Experimental observation of the quantum Hall effect and Berry’s phase in graphene. Nature 2005, 438, 201–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Neto, A.C.; Guinea, F.; Peres, N.M.; Novoselov, K.S.; Geim, A.K. The electronic properties of graphene. Rev. Mod. Phys. 2009, 81, 109. [Google Scholar] [CrossRef] [Green Version]
  5. Geim, A.K. Graphene: Status and prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [Green Version]
  6. Geim, A.K.; Novoselov, K.S. The rise of graphene. In Nanoscience and Technology: A Collection of Reviews from Nature Journals; World Scientific: Singapore, 2010; pp. 11–19. [Google Scholar]
  7. Stankovich, S.; Dikin, D.A.; Dommett, G.H.B.; Kohlhaas, K.M.; Zimney, E.J.; Stach, E.A.; Piner, R.D.; Nguyen, S.T.; Ruoff, R.S. Graphene-based composite materials. Nature 2006, 442, 282–286. [Google Scholar] [CrossRef]
  8. Nigar, S.; Zhou, Z.; Wang, H.; Imtiaz, M. Modulating the electronic and magnetic properties of graphene. RSC Adv. 2017, 7, 51546–51580. [Google Scholar] [CrossRef] [Green Version]
  9. Papageorgiou, D.G.; Kinloch, I.A.; Young, R.J. Mechanical properties of graphene and graphene-based nanocomposites. Prog. Mater. Sci. 2017, 90, 75–127. [Google Scholar] [CrossRef]
  10. Das, A.; Pisana, S.; Chakraborty, B.; Piscanec, S.; Saha, S.K.; Waghmare, U.V.; Novoselov, K.S.; Krishnamurthy, H.R.; Geim, A.K.; Ferrari, A.C.; et al. Monitoring dopants by Raman scattering in an electrochemically top-gated graphene transistor. Nat. Nanotechnol. 2008, 3, 210–215. [Google Scholar] [CrossRef] [Green Version]
  11. Britnell, L.; Gorbachev, R.; Jalil, R.; Belle, B.; Schedin, F.; Mishchenko, A.; Georgiou, T.; Katsnelson, M.; Eaves, L.; Morozov, S.; et al. Field-effect tunneling transistor based on vertical graphene heterostructures. Science 2012, 335, 947–950. [Google Scholar] [CrossRef]
  12. Stoller, M.D.; Park, S.; Zhu, Y.; An, J.; Ruoff, R.S. Graphene-based ultracapacitors. Nano Lett. 2008, 8, 3498–3502. [Google Scholar] [CrossRef] [PubMed]
  13. Pumera, M. Graphene-based nanomaterials for energy storage. Energy Environ. Sci. 2011, 4, 668–674. [Google Scholar] [CrossRef]
  14. Raccichini, R.; Varzi, A.; Passerini, S.; Scrosati, B. The role of graphene for electrochemical energy storage. Nat. Mater. 2015, 14, 271–279. [Google Scholar] [CrossRef]
  15. Steane, A. Quantum computing. Rep. Prog. Phys. 1998, 61, 117. [Google Scholar] [CrossRef] [Green Version]
  16. Calafell, I.A.; Cox, J.; Radonjić, M.; Saavedra, J.; de Abajo, F.G.; Rozema, L.; Walther, P. Quantum computing with graphene plasmons. NPJ Quantum Inf. 2019, 5, 1–7. [Google Scholar]
  17. Patel, S.K.S.; Choi, S.H.; Kang, Y.C.; Lee, J.K. Eco-Friendly Composite of Fe3O4-Reduced Graphene Oxide Particles for Efficient Enzyme Immobilization. ACS Appl. Mater. Interfaces 2017, 9, 2213–2222. [Google Scholar] [CrossRef] [PubMed]
  18. Otari, S.V.; Kumar, M.; Anwar, M.Z.; Thorat, N.D.; Patel, S.K.S.; Lee, D.; Lee, J.H.; Lee, J.K.; Kang, Y.C.; Zhang, L. Rapid synthesis and decoration of reduced graphene oxide with gold nanoparticles by thermostable peptides for memory device and photothermal applications. Sci. Rep. 2017, 7, 10980. [Google Scholar] [CrossRef]
  19. Williams, J.; DiCarlo, L.; Marcus, C. Quantum Hall effect in a gate-controlled pn junction of graphene. Science 2007, 317, 638–641. [Google Scholar] [CrossRef] [Green Version]
  20. Abanin, D.; Levitov, L. Quantized transport in graphene pn junctions in a magnetic field. Science 2007, 317, 641–643. [Google Scholar] [CrossRef] [Green Version]
  21. Özyilmaz, B.; Jarillo-Herrero, P.; Efetov, D.; Abanin, D.A.; Levitov, L.S.; Kim, P. Electronic transport and quantum Hall effect in bipolar graphene p-n-p junctions. Phys. Rev. Lett. 2007, 99, 166804. [Google Scholar] [CrossRef] [Green Version]
  22. Chen, H.; Müller, M.B.; Gilmore, K.J.; Wallace, G.G.; Li, D. Mechanically strong, electrically conductive, and biocompatible graphene paper. Adv. Mater. 2008, 20, 3557–3561. [Google Scholar] [CrossRef]
  23. Mak, K.F.; Sfeir, M.Y.; Wu, Y.; Lui, C.H.; Misewich, J.A.; Heinz, T.F. Measurement of the optical conductivity of graphene. Phys. Rev. Lett. 2008, 101, 196405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Balandin, A.A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C.N. Superior thermal conductivity of single-layer graphene. Nano Lett. 2008, 8, 902–907. [Google Scholar] [CrossRef] [PubMed]
  25. Nakada, K.; Fujita, M.; Dresselhaus, G.; Dresselhaus, M.S. Edge state in graphene ribbons: Nanometer size effect and edge shape dependence. Phys. Rev. B 1996, 54, 17954. [Google Scholar] [CrossRef] [Green Version]
  26. Wakabayashi, K.; Fujita, M.; Ajiki, H.; Sigrist, M. Electronic and magnetic properties of nanographite ribbons. Phys. Rev. B 1999, 59, 8271. [Google Scholar] [CrossRef] [Green Version]
  27. Ezawa, M. Peculiar width dependence of the electronic properties of carbon nanoribbons. Phys. Rev. B 2006, 73, 045432. [Google Scholar] [CrossRef] [Green Version]
  28. Brey, L.; Fertig, H. Electronic states of graphene nanoribbons studied with the Dirac equation. Phys. Rev. B 2006, 73, 235411. [Google Scholar] [CrossRef] [Green Version]
  29. Sasaki, K.i.; Murakami, S.; Saito, R. Gauge field for edge state in graphene. J. Phys. Soc. Jpn. 2006, 75, 074713. [Google Scholar] [CrossRef] [Green Version]
  30. Son, Y.W.; Cohen, M.L.; Louie, S.G. Energy gaps in graphene nanoribbons. Phys. Rev. Lett. 2006, 97, 216803. [Google Scholar] [CrossRef] [Green Version]
  31. Fernández-Rossier, J.; Palacios, J.J. Magnetism in graphene nanoislands. Phys. Rev. Lett. 2007, 99, 177204. [Google Scholar] [CrossRef] [Green Version]
  32. Ponomarenko, L.A.; Schedin, F.; Katsnelson, M.I.; Yang, R.; Hill, E.W.; Novoselov, K.S.; Geim, A.K. Chaotic Dirac billiard in graphene quantum dots. Science 2008, 320, 356–358. [Google Scholar] [CrossRef] [PubMed]
  33. Hod, O.; Barone, V.; Scuseria, G.E. Half-metallic graphene nanodots: A comprehensive first-principles theoretical study. Phys. Rev. B 2008, 77, 035411. [Google Scholar] [CrossRef] [Green Version]
  34. Russo, S.; Oostinga, J.B.; Wehenkel, D.; Heersche, H.B.; Sobhani, S.S.; Vandersypen, L.M.; Morpurgo, A.F. Observation of Aharonov-Bohm conductance oscillations in a graphene ring. Phys. Rev. B 2008, 77, 085413. [Google Scholar] [CrossRef] [Green Version]
  35. Hewageegana, P.; Apalkov, V. Electron localization in graphene quantum dots. Phys. Rev. B 2008, 77, 245426. [Google Scholar] [CrossRef]
  36. Bahamon, D.; Pereira, A.; Schulz, P. Inner and outer edge states in graphene rings: A numerical investigation. Phys. Rev. B 2009, 79, 125414. [Google Scholar] [CrossRef] [Green Version]
  37. Downing, C.; Stone, D.; Portnoi, M. Zero-energy states in graphene quantum dots and rings. Phys. Rev. B 2011, 84, 155437. [Google Scholar] [CrossRef] [Green Version]
  38. Bacon, M.; Bradley, S.J.; Nann, T. Graphene quantum dots. Part. Part. Syst. Charact. 2014, 31, 415–428. [Google Scholar] [CrossRef]
  39. Recher, P.; Trauzettel, B.; Rycerz, A.; Blanter, Y.M.; Beenakker, C.; Morpurgo, A. Aharonov-Bohm effect and broken valley degeneracy in graphene rings. Phys. Rev. B 2007, 76, 235404. [Google Scholar] [CrossRef] [Green Version]
  40. Potasz, P.; Güçlü, A.; Voznyy, O.; Folk, J.; Hawrylak, P. Electronic and magnetic properties of triangular graphene quantum rings. Phys. Rev. B 2011, 83, 174441. [Google Scholar] [CrossRef] [Green Version]
  41. El Khatib, M.; Evangelisti, S.; Leininger, T.; Bendazzoli, G.L. A theoretical study of closed polyacene structures. Phys. Chem. Chem. Phys. 2012, 14, 15666–15676. [Google Scholar] [CrossRef]
  42. Schelter, J.; Recher, P.; Trauzettel, B. The Aharonov–Bohm effect in graphene rings. Solid State Commun. 2012, 152, 1411–1419. [Google Scholar] [CrossRef]
  43. Da Costa, D.; Chaves, A.; Zarenia, M.; Pereira, J., Jr.; Farias, G.; Peeters, F. Geometry and edge effects on the energy levels of graphene quantum rings: A comparison between tight-binding and simplified Dirac models. Phys. Rev. B 2014, 89, 075418. [Google Scholar] [CrossRef]
  44. Yan, C.H.; Wei, L.F. Size effects in Aharonov–Bohm graphene rings. J. Phys. Condens. Matter 2010, 22, 295503. [Google Scholar] [CrossRef]
  45. Schelter, J.; Bohr, D.; Trauzettel, B. Interplay of the Aharonov-Bohm effect and Klein tunneling in graphene. Phys. Rev. B 2010, 81, 195441. [Google Scholar] [CrossRef] [Green Version]
  46. Munárriz, J.; Domínguez-Adame, F.; Malyshev, A. Toward graphene-based quantum interference devices. Nanotechnology 2011, 22, 365201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Wang, D.; Zhang, Z.; Zhang, J.; Deng, X.; Fan, Z.; Tang, G. Electronic and magnetic properties of zigzag-edged hexagonal graphene ring nanojunctions. Carbon 2015, 94, 996–1002. [Google Scholar] [CrossRef]
  48. Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864. [Google Scholar] [CrossRef] [Green Version]
  49. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133. [Google Scholar] [CrossRef] [Green Version]
  50. Hachmann, J.; Dorando, J.J.; Avilés, M.; Chan, G.K.L. The radical character of the acenes: A density matrix renormalization group study. J. Chem. Phys. 2007, 127, 134309. [Google Scholar] [CrossRef] [Green Version]
  51. Gidofalvi, G.; Mazziotti, D.A. Active-space two-electron reduced-density-matrix method: Complete active-space calculations without diagonalization of the N-electron Hamiltonian. J. Chem. Phys. 2008, 129, 134108. [Google Scholar] [CrossRef]
  52. Mizukami, W.; Kurashige, Y.; Yanai, T. More π electrons make a difference: Emergence of many radicals on graphene nanoribbons studied by ab initio DMRG theory. J. Chem. Theory Comput. 2013, 9, 401–407. [Google Scholar] [CrossRef] [PubMed]
  53. Chai, J.D. Density functional theory with fractional orbital occupations. J. Chem. Phys. 2012, 136, 154104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Mermin, N. Thermal properties of the inhomogeneous electron gas. Phys. Rev. 1965, 137, A1441. [Google Scholar] [CrossRef]
  55. Chai, J.D. Thermally-assisted-occupation density functional theory with generalized-gradient approximations. J. Chem. Phys. 2014, 140, 18A521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Chai, J.D. Role of exact exchange in thermally-assisted-occupation density functional theory: A proposal of new hybrid schemes. J. Chem. Phys. 2017, 146, 044102. [Google Scholar] [CrossRef] [Green Version]
  57. Xuan, F.; Chai, J.D.; Su, H. Local density approximation for the short-range exchange free energy functional. ACS Omega 2019, 4, 7675–7683. [Google Scholar] [CrossRef] [Green Version]
  58. Chen, B.J.; Chai, J.D. TAO-DFT fictitious temperature made simple. RSC Adv. 2022, 12, 12193–12210. [Google Scholar] [CrossRef]
  59. Lin, C.Y.; Hui, K.; Chung, J.H.; Chai, J.D. Self-consistent determination of the fictitious temperature in thermally-assisted-occupation density functional theory. RSC Adv. 2017, 7, 50496–50507. [Google Scholar] [CrossRef] [Green Version]
  60. Li, S.; Chai, J.D. TAO-DFT-based ab initio molecular dynamics. Front. Chem. 2020, 8, 589432. [Google Scholar] [CrossRef]
  61. Yeh, S.H.; Manjanath, A.; Cheng, Y.C.; Chai, J.D.; Hsu, C.P. Excitation energies from thermally assisted-occupation density functional theory: Theory and computational implementation. J. Chem. Phys. 2020, 153, 084120. [Google Scholar] [CrossRef]
  62. Wu, C.S.; Chai, J.D. Electronic Properties of Zigzag Graphene Nanoribbons Studied by TAO-DFT. J. Chem. Theory Comput. 2015, 11, 2003–2011. [Google Scholar] [CrossRef] [PubMed]
  63. Yeh, C.N.; Chai, J.D. Role of Kekulé and Non-Kekulé Structures in the Radical Character of Alternant Polycyclic Aromatic Hydrocarbons: A TAO-DFT Study. Sci. Rep. 2016, 6, 30562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Seenithurai, S.; Chai, J.D. Effect of Li adsorption on the electronic and hydrogen storage properties of acenes: A dispersion-corrected TAO-DFT study. Sci. Rep. 2016, 6, 33081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Wu, C.S.; Lee, P.Y.; Chai, J.D. Electronic properties of cyclacenes from TAO-DFT. Sci. Rep. 2016, 6, 37249. [Google Scholar] [CrossRef] [Green Version]
  66. Seenithurai, S.; Chai, J.D. Effect of Li Termination on the Electronic and Hydrogen Storage Properties of Linear Carbon Chains: A TAO-DFT Study. Sci. Rep. 2017, 7, 4966. [Google Scholar] [CrossRef] [Green Version]
  67. Seenithurai, S.; Chai, J.D. Electronic and Hydrogen Storage Properties of Li-Terminated Linear Boron Chains Studied by TAO-DFT. Sci. Rep. 2018, 8, 13538. [Google Scholar] [CrossRef]
  68. Yeh, C.N.; Wu, C.; Su, H.; Chai, J.D. Electronic properties of the coronene series from thermally-assisted-occupation density functional theory. RSC Adv. 2018, 8, 34350–34358. [Google Scholar] [CrossRef] [Green Version]
  69. Chung, J.H.; Chai, J.D. Electronic properties of Möbius cyclacenes studied by thermally-assisted-occupation density functional theory. Sci. Rep. 2019, 9, 2907. [Google Scholar] [CrossRef] [Green Version]
  70. Seenithurai, S.; Chai, J.D. Electronic properties of linear and cyclic boron nanoribbons from thermally-assisted-occupation density functional theory. Sci. Rep. 2019, 9, 12139. [Google Scholar] [CrossRef] [Green Version]
  71. Deng, Q.; Chai, J.D. Electronic properties of triangle-shaped graphene nanoflakes from TAO-DFT. ACS Omega 2019, 4, 14202–14210. [Google Scholar] [CrossRef] [Green Version]
  72. Huang, H.J.; Seenithurai, S.; Chai, J.D. TAO-DFT study on the electronic properties of diamond-shaped graphene nanoflakes. Nanomaterials 2020, 10, 1236. [Google Scholar] [CrossRef]
  73. Seenithurai, S.; Chai, J.D. TAO-DFT investigation of electronic properties of linear and cyclic carbon chains. Sci. Rep. 2020, 10, 13133. [Google Scholar] [CrossRef] [PubMed]
  74. Seenithurai, S.; Chai, J.D. Electronic properties of carbon nanobelts predicted by thermally-assisted-occupation DFT. Nanomaterials 2021, 11, 2224. [Google Scholar] [CrossRef] [PubMed]
  75. Hanson-Heine, M.W. Metal sandwich and ion complexes in cyclacene nanobelts. Mol. Phys. 2022, e2118187. [Google Scholar] [CrossRef]
  76. Hanson-Heine, M.W. Static correlation in vibrational frequencies studied using thermally-assisted-occupation density functional theory. Chem. Phys. Lett. 2020, 739, 137012. [Google Scholar] [CrossRef]
  77. Hanson-Heine, M.W. Static Electron Correlation in Anharmonic Molecular Vibrations: A Hybrid TAO-DFT Study. J. Phys. Chem. A 2022, 126, 7273–7282. [Google Scholar] [CrossRef] [PubMed]
  78. Shao, Y.; Gan, Z.; Epifanovsky, E.; Gilbert, A.T.; Wormit, M.; Kussmann, J.; Lange, A.W.; Behn, A.; Deng, J.; Feng, X.; et al. Advances in molecular quantum chemistry contained in the Q-Chem 4 program package. Mol. Phys. 2015, 113, 184–215. [Google Scholar] [CrossRef] [Green Version]
  79. Yeh, C.N.; Lee, P.Y.; Chai, J.D. Electronic and Optical Properties of the Narrowest Armchair Graphene Nanoribbons Studied by Density Functional Methods. Aust. J. Chem. 2016, 69, 960. [Google Scholar] [CrossRef] [Green Version]
  80. Perdew, J.P.; Yang, W.; Burke, K.; Yang, Z.; Gross, E.K.; Scheffler, M.; Scuseria, G.E.; Henderson, T.M.; Zhang, I.Y.; Ruzsinszky, A.; et al. Understanding band gaps of solids in generalized Kohn–Sham theory. Proc. Natl. Acad. Sci. USA 2017, 114, 2801–2806. [Google Scholar] [CrossRef] [Green Version]
  81. Takatsuka, K.; Fueno, T.; Yamaguchi, K. Distribution of odd electrons in ground-state molecules. Theor. Chim. Acta 1978, 48, 175–183. [Google Scholar] [CrossRef]
  82. Head-Gordon, M. Characterizing unpaired electrons from the one-particle density matrix. Chem. Phys. Lett. 2003, 372, 508–511. [Google Scholar] [CrossRef]
  83. Grimme, S.; Hansen, A. A Practicable Real-Space Measure and Visualization of Static Electron-Correlation Effects. Angew. Chem. Int. Ed. 2015, 54, 12308–12313. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Geometry of n-HGQR, consisting of n aromatic rings fused together at each side.
Figure 1. Geometry of n-HGQR, consisting of n aromatic rings fused together at each side.
Nanomaterials 12 03943 g001
Figure 2. Singlet–triplet gap of n-HGQR calculated by spin-unrestricted KS-LDA and TAO-LDA.
Figure 2. Singlet–triplet gap of n-HGQR calculated by spin-unrestricted KS-LDA and TAO-LDA.
Nanomaterials 12 03943 g002
Figure 3. (a) Vertical ionization potential, (b) vertical electron affinity as well as (c) fundamental gap associated with ground-state n-HGQR, calculated by spin-unrestricted TAO-LDA.
Figure 3. (a) Vertical ionization potential, (b) vertical electron affinity as well as (c) fundamental gap associated with ground-state n-HGQR, calculated by spin-unrestricted TAO-LDA.
Nanomaterials 12 03943 g003
Figure 4. Symmetrized von Neumann entropy associated with ground-state n-HGQR, calculated by spin-unrestricted TAO-LDA.
Figure 4. Symmetrized von Neumann entropy associated with ground-state n-HGQR, calculated by spin-unrestricted TAO-LDA.
Nanomaterials 12 03943 g004
Figure 5. Active TAO-orbital occupation numbers (H − 12, H − 11, …, H − 1, H, L, L + 1, …, L + 11, and L + 12) of ground-state n-HGQR, calculated by spin-restricted TAO-LDA.
Figure 5. Active TAO-orbital occupation numbers (H − 12, H − 11, …, H − 1, H, L, L + 1, …, L + 11, and L + 12) of ground-state n-HGQR, calculated by spin-restricted TAO-LDA.
Nanomaterials 12 03943 g005
Figure 6. Real-space representation of active TAO-orbitals, such as H − 5 (2.000), H − 4 (2.000), H − 3 (1.998), H − 2 (1.998), H − 1 (1.996), H (1.996), L (0.003), L + 1 (0.003), L + 2 (0.002), L + 3 (0.002), L + 4 (0.000), and L + 5 (0.000), of ground-state 3-HGQR, obtained from spin-restricted TAO-LDA, at isovalue = 0.02 e/Å 3 . The TOONs are given in parentheses.
Figure 6. Real-space representation of active TAO-orbitals, such as H − 5 (2.000), H − 4 (2.000), H − 3 (1.998), H − 2 (1.998), H − 1 (1.996), H (1.996), L (0.003), L + 1 (0.003), L + 2 (0.002), L + 3 (0.002), L + 4 (0.000), and L + 5 (0.000), of ground-state 3-HGQR, obtained from spin-restricted TAO-LDA, at isovalue = 0.02 e/Å 3 . The TOONs are given in parentheses.
Nanomaterials 12 03943 g006
Figure 7. Real-space representation of active TAO-orbitals, such as H − 5 (1.987), H − 4 (1.983), H − 3 (1.964), H − 2 (1.964), H − 1 (1.958), H (1.958), L (0.043), L + 1 (0.043), L + 2 (0.035), L + 3 (0.034), L + 4 (0.020), and L + 5 (0.011), of ground-state 5-HGQR, obtained from spin-restricted TAO-LDA, at isovalue = 0.02 e/Å 3 . The TOONs are given in parentheses.
Figure 7. Real-space representation of active TAO-orbitals, such as H − 5 (1.987), H − 4 (1.983), H − 3 (1.964), H − 2 (1.964), H − 1 (1.958), H (1.958), L (0.043), L + 1 (0.043), L + 2 (0.035), L + 3 (0.034), L + 4 (0.020), and L + 5 (0.011), of ground-state 5-HGQR, obtained from spin-restricted TAO-LDA, at isovalue = 0.02 e/Å 3 . The TOONs are given in parentheses.
Nanomaterials 12 03943 g007
Figure 8. Real-space representation of active TAO-orbitals, such as H − 5 (1.876), H − 4 (1.826), H − 3 (1.810), H − 2 (1.810), H − 1 (1.777), H (1.777), L (0.225), L + 1 (0.225), L + 2 (0.190), L + 3 (0.189), L + 4 (0.179), and L + 5 (0.121), of ground-state 7-HGQR, obtained from spin-restricted TAO-LDA, at isovalue = 0.02 e/Å 3 . The TOONs are given in parentheses.
Figure 8. Real-space representation of active TAO-orbitals, such as H − 5 (1.876), H − 4 (1.826), H − 3 (1.810), H − 2 (1.810), H − 1 (1.777), H (1.777), L (0.225), L + 1 (0.225), L + 2 (0.190), L + 3 (0.189), L + 4 (0.179), and L + 5 (0.121), of ground-state 7-HGQR, obtained from spin-restricted TAO-LDA, at isovalue = 0.02 e/Å 3 . The TOONs are given in parentheses.
Nanomaterials 12 03943 g008
Figure 9. Real-space representation of active TAO-orbitals, such as H − 5 (1.613), H − 4 (1.510), H − 3 (1.504), H − 2 (1.504), H − 1 (1.446), H (1.445), L (0.551), L + 1 (0.551), L + 2 (0.509), L + 3 (0.508), L + 4 (0.486), and L + 5 (0.400) of ground-state 9-HGQR, obtained from spin-restricted TAO-LDA, at isovalue = 0.02 e/Å 3 . The TOONs are given in parentheses.
Figure 9. Real-space representation of active TAO-orbitals, such as H − 5 (1.613), H − 4 (1.510), H − 3 (1.504), H − 2 (1.504), H − 1 (1.446), H (1.445), L (0.551), L + 1 (0.551), L + 2 (0.509), L + 3 (0.508), L + 4 (0.486), and L + 5 (0.400) of ground-state 9-HGQR, obtained from spin-restricted TAO-LDA, at isovalue = 0.02 e/Å 3 . The TOONs are given in parentheses.
Nanomaterials 12 03943 g009
Table 1. Expectation values of the total spin-squared operator S ^ 2 associated with the lowest singlet and lowest triplet states of n-HGQR, calculated by spin-unrestricted KS-LDA. For the exact theory, the value of S ^ 2 should be 0 for the lowest singlet state and two for the lowest triplet state.
Table 1. Expectation values of the total spin-squared operator S ^ 2 associated with the lowest singlet and lowest triplet states of n-HGQR, calculated by spin-unrestricted KS-LDA. For the exact theory, the value of S ^ 2 should be 0 for the lowest singlet state and two for the lowest triplet state.
nSingletTriplet
30.00002.0047
40.00002.0040
50.00002.0028
60.00002.0054
70.00002.0068
80.00002.0101
90.00012.0271
100.00042.8843
114.14325.1379
125.18046.2166
134.84295.9705
143.83385.0410
Table 2. Difference E UR (in kcal/mol) between the lowest spin-restricted singlet energy and lowest spin-unrestricted singlet energy of n-HGQR, calculated by KS-LDA and TAO-LDA.
Table 2. Difference E UR (in kcal/mol) between the lowest spin-restricted singlet energy and lowest spin-unrestricted singlet energy of n-HGQR, calculated by KS-LDA and TAO-LDA.
nKS-LDATAO-LDA
30.0000.000
40.0000.000
50.0000.000
60.0000.000
70.0000.000
80.0000.000
90.0000.000
100.0000.000
111.6090.000
127.6330.000
135.2610.000
142.4890.000
15 0.000
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, C.-C.; Chai, J.-D. Electronic Properties of Hexagonal Graphene Quantum Rings from TAO-DFT. Nanomaterials 2022, 12, 3943. https://doi.org/10.3390/nano12223943

AMA Style

Chen C-C, Chai J-D. Electronic Properties of Hexagonal Graphene Quantum Rings from TAO-DFT. Nanomaterials. 2022; 12(22):3943. https://doi.org/10.3390/nano12223943

Chicago/Turabian Style

Chen, Chi-Chun, and Jeng-Da Chai. 2022. "Electronic Properties of Hexagonal Graphene Quantum Rings from TAO-DFT" Nanomaterials 12, no. 22: 3943. https://doi.org/10.3390/nano12223943

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop