Next Article in Journal
Simple Synthesis and Characterization of Shell-Thickness-Controlled Ni/Ni3C Core-Shell Nanoparticles
Previous Article in Journal
Sodium-Based Cylindrical Plasmonic Waveguides in the Near-Infrared
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quantum Revivals in Curved Graphene Nanoflakes

by
Sergio de-la-Huerta-Sainz
1,
Angel Ballesteros
1 and
Nicolás A. Cordero
1,2,3,*
1
Physics Department, Universidad de Burgos, E-09001 Burgos, Spain
2
International Research Center in Critical Raw Materials for Advanced Industrial Technologies (ICCRAM), Universidad de Burgos, E-09001 Burgos, Spain
3
Institute Carlos I for Theoretical and Computational Physics (IC1), E-18016 Granada, Spain
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(12), 1953; https://doi.org/10.3390/nano12121953
Submission received: 23 April 2022 / Revised: 1 June 2022 / Accepted: 3 June 2022 / Published: 7 June 2022

Abstract

:
Graphene nanostructures have attracted a lot of attention in recent years due to their unconventional properties. We have employed Density Functional Theory to study the mechanical and electronic properties of curved graphene nanoflakes. We explore hexagonal flakes relaxed with different boundary conditions: (i) all atoms on a perfect spherical sector, (ii) only border atoms forced to be on the spherical sector, and (iii) only vertex atoms forced to be on the spherical sector. For each case, we have analysed the behaviour of curvature energy and of quantum regeneration times (classical and revival) as the spherical sector radius changes. Revival time presents in one case a divergence usually associated with a phase transition, probably caused by the pseudomagnetic field created by the curvature. This could be the first case of a phase transition in graphene nanostructures without the presence of external electric or magnetic fields.

1. Introduction

The experimental isolation of a single graphitic layer (now known as graphene) by means of the so-called “Scotch tape method” by Geim and Novoselov [1] has undoubtedly opened a new field in science. The proof is that in recent years, over 1% of all the scientific publications included in the Web of Science™ global citation database [2] are related to this one-atom-thick system.
The outstanding properties of this 2D material have led to many application proposals. To cite just a few of them: nanocomposites for bone tissue engineering [3], composites for multifunctional applications [4], lubrication [5,6], solar cells [7], ultracapacitors [8,9], batteries [10], sensors [11], catalysis [12], nanomedicine [13], fuel cells [14,15] or even energy harvesting [16].
Quantum revivals consist in the temporal periodic reconstruction of a wave packet in systems with a commensurable discrete spectrum [17]. They have recently attracted attention due to their possible interest in quantum devices [18,19,20,21]. Quantum revivals have been studied in infinite graphene under magnetic fields [22,23,24] or circularly polarized radiation [25], as well as in graphene rings [26] and graphene quantum dots [27].
One of the common misconceptions about graphene is that it is flat. In fact, Peierls instability [28,29] creates ripples in free-standing graphene [30,31,32,33,34,35]. Atomistic Monte Carlo simulations based on a very accurate many-body interatomic potential for carbon [36] gave ripples with a size distribution that peaked around 80 Å, in agreement with experiments that yield results in the 50–100 Å range [31]. Besides, when graphene is grown or deposited on a substrate, nanobubbles with diameters between 80 Å and 1000 Å appear [37,38,39,40,41].
Rectangular graphene flakes either free [42] or subjected to planar bending [43] have been analysed in the literature. Conical graphene rings in a magnetic field have also been studied, but only in the continuum limit approximation in the vicinity of the Dirac points and very far from the vertex [44,45]. Both in-plane and out-of-plane bending result in the appearance of a pseudo-magnetic field.
We present in this article the first (to the best of our knowledge) calculation of quantum revivals using Density Functional Theory (DFT) and use it to study the interplay between curvature and regeneration times in a non-planar graphene flake.

2. Materials and Methods

Since Wallace did the first calculation of the electronic properties of graphene using the tight-binding method [46], many theoretical models have been used to study this system: molecular mechanics, molecular dynamics, semiempirical methods, Density Functional Theory, Hartree–Fock (HF), post-HF including correlation, Monte Carlo simulations, hybrid methods, continuum models, etc. We have used the Density Functional Theory (DFT) formalism [47] within the Local Density Approximation (LDA) [48,49] as implemented in the Gaussian 09 [50] and Gaussian 16 [51] suites of programs. We have selected DFT for its balance between accuracy and computational effort, and have chosen LDA because it gives better results than gradient corrected approximations (GGAs) for graphitic systems [52,53,54] and because it has been previously used to successfully study the interaction between carbon nanostructures and several small molecules and atoms [55,56,57,58,59,60] as well as—very recently—graphene nanoribbons [61]. We have selected the 6-31G** basis set [62,63,64,65,66,67] that adds to the 6-31G set d-type and p-type Cartesian–Gaussian polarization functions and is commonly used for carbon nanostructures calculations.
In order to avoid non-trivial magnetic ground states induced by asymmetric edge extensions [68] we have used the hexagonal flake with zig-zag borders passivated with hydrogen atoms shown in Figure 1. The size of the flake (10 concentric hexagonal rings comprising 600 C atoms and 60 H atoms with a separation of approximately 47 Å between opposite vertices) has been selected as a reasonable compromise between the size of experimentally measured ripples and computational cost.
We have studied (quasi-)spherical nanoflakes of different curvature radii with the three different bending possibilities depicted in Figure 2 for a radius of 40 Å: forcing all 600 carbon atoms to lie on a spherical surface (upper panel), fixing only the 60 border atoms on the surface and allowing the rest to relax (middle panel) and forcing only the 12 vertex atoms to belong to the sphere and not imposing any condition on the rest (lower panel). Since carbon atoms tend to be in a sp 2 hybridization state, the nanoflake tends to flatten, as the restrictions are less demanding.

3. Results and Discussion

We have studied both curvature energy and quantum revivals in this nanostructure. The first one for checking if the macroscopic continuum limit model that predicts that curvature energy is proportional to the Gaussian curvature (i.e., the inverse of the radius squared) [70] holds at this scale. The second one for trying to find trends in regeneration times.

3.1. Curvature Energy

We present in Figure 3 the dependence of the energy of the nanoflake E with the inverse squared radius of the sphere 1 R 2 . Since we have taken the flat configuration as the energy origin, this plot represents the curvature energy. For the case of a perfect spherical surface that corresponds to our fixed surface calculations (shown in red in the figure), 1 R 2 is precisely the Gaussian curvature. For the fixed borders (depicted in blue) and fixed vertices (painted in green) cases, curvature is no longer constant, but we can take the value of 1 R 2 for the fixed atoms as an estimate of the curvature for comparison purposes. Logically, for a fixed value of R the geometry with only the vertices fixed is energetically more favourable than that with the borders fixed and this in turn is more stable than the one with all the atoms forced to lie on a spherical surface. As expected, for the three geometries the curvature energy increases with curvature but the trend for the fixed surface case is somewhat unexpected. The macroscopic continuum result E 1 R 2 would lead to a straight red line, but that does not seem to be the case.
To check if the macroscopic result holds at least for low curvatures, we present in Figure 4 a log–log plot of the curvature energy versus radius.
E = k 1 R 2 log E = log k 2 log R ,
and the red graph should present a constant slope equal to −2. This is clearly not the case and the continuum approximation is not valid for so small a flake.
This deviation from the continuum behaviour is undoubtedly due to the difference between the classical physics laws governing the macroscopic world and the quantum ones ruling at the nanoscale. This difference translates into two distinct aspects. The equilibrium geometry dictated by both sets of laws is different and, even for the same configuration, they lead to different energies. To check which one is more important in this case, we have made a series of non-self-consistent calculations with a hybrid Molecular Mechanics (MM)/DFT model. We have first used LDA calculations to determine a “semiclassical” Hooke potential for the C–C bond. To this end, we have used a graphene flake similar to the one in Figure 1, but with only 7 concentric hexagonal rings. We have changed the length of the central C–C bond and calculated the energy of the system for both stretched and compressed bond lengths using LDA. We have fitted the energies to a parabolic curve and determined a C–C pair potential for carbon nanoflakes. We have written an in-house code to determine the semiclassical equilibrium geometry for the spherical nanoflakes. This geometry is then used in a single-point calculation to determine the LDA energy of the nanoflake. We will label the results obtained by this procedure as “fixed surface (MM)”.
We have included in Figure 3 the energies resulting from this hybrid approach in gray. The energies are bigger that those corresponding to the self-consistent LDA calculation (in red) since, according to the variational principle, using a wave function different to that of the fundamental state leads to a higher energy. The greater the curvature, the bigger the energy increase.
In order to check if the continuum model is valid for this approach, we have also included the corresponding results in Figure 4. The graph is now nearly a straight line, but there is still some non-linearity. Equation (1) seems to be approximately valid, but the factor in front of log R depends slightly on R, adopting the form
E = k 1 R n log E = log k n log R .
We present in Figure 5 the value of n as R changes calculated using a 3-point finite differences method.
It is possible to fit the results in this figure using the simplest Padé approximant
n = a 0 + a 1 R 1 + b 1 R .
The result of this fitting is presented in Table 1.
We can use this Padé approximant to calculate the asymptotic behaviour of n.
lim R n = a 1 b 1 = 2.02 .
This result is very close to 2, which is the value predicted in the continuum model. Therefore, the semiclassical MM/DFT approach tends to the classical continuum macroscopic limit, proving that the main quantum contribution to the deviation from this model is due to the small change in the equilibrium geometry.

3.2. Quantum Revivals

We can write the initial state of a time-independent Hamiltonian H ^ as a linear combination of its eigenfunctions:
Ψ ( 0 ) = n = 0 a n u n ,
where a n are constants and u n are the eigenfunctions with energies E n ,
H ^ u n = E n u n .
The temporal evolution of this state can be written as
Ψ ( t ) = n = 0 a n u n e i E n t .
Since we have calculated the energy of curved graphene nanoflakes using DFT (i.e., solving the Kohn–Sham equations for the system) we know their energy spectrum and can therefore study the time evolution of a wave packet in these systems. Let us consider a superposition of eigenstates of the Hamiltonian concentrated around a central energy level n 0 characterised by an energy E n 0 . We can perform a Taylor expansion of the energy spectrum around E n 0 :
E n = E n 0 + E n 0 ( n n 0 ) + 1 2 ! E n 0 ( n n 0 ) 2 + 1 3 ! E n 0 ( n n 0 ) 3 + .
Taking into account Equations (7) and (8),
Ψ ( t ) = n = 0 a n u n e i E n 0 + E n 0 ( n n 0 ) + 1 2 ! E n 0 ( n n 0 ) 2 + 1 3 ! E n 0 ( n n 0 ) 3 + t .
Each term in the exponential (except the first one that is just a global phase) defines a characteristic time scale:
T Cl 2 π | E n 0 | is called c l a s s i c a l   t i m e ,
T Re 2 π | E n 0 | / 2 is called r e v i v a l   t i m e , and
T Sup 2 π | E n 0 | / 6 is called s u p e r - r e v i v a l   t i m e
(see [17] for further details).
We are going to consider a Gaussian initial wave packet,
a n = 1 σ π e ( n n 0 ) 2 2 σ 2 ,
where, we have selected the central level as the fourth level above the Highest Occupied Molecular Orbital (HOMO) n 0 = HOMO + 4 and, in order to get a sharply concentrated packet σ = 0.7 so that only five levels around n 0 have a significative contribution ( a n > 0.001 ).
The easiest way of visualising wave packet regeneration is making use of the squared modulus of the so called autocorrelation function that measures the overlap of the wave packet at times 0 and t:
A ( t ) 2 = Ψ ( 0 ) | Ψ ( t ) 2 .
A typical case is presented in Figure 6. A ( t ) 2 oscillates very fast, reaching a maximum every T Cl inside an envelope with T Re periodicity. T Sup is usually much larger that T Re and we will not consider it in this work.
In principle, calculating T Cl is straightforward. One only has to search for the first maximum of A ( t ) 2 . Determining T Re is not so easy. In this case, it is necessary to calculate the envelope of the function and calculate its first maximum. However, for some curvatures, the situation is not as clear as the one depicted in Figure 6. If T Re is not much bigger than T Cl , both times interfere, the pattern is more complicated and it is difficult to determine both of them, especially T Cl . It is then desirable to have another way of calculating these times. The solution is employing Equations (10) and (11). To do that, we have calculated an interpolating function by adjusting a parabolic curve to every three consecutive levels in the spectrum and used it to calculate the first two derivatives of the energy with respect to the level that appears in those equations. Therefore, we have two different ways of calculating these times. We will call the first one numerical (num.) and the second one analytical (an.).
Experiments for measuring quantum regeneration times are based on the use of two consecutive laser pulses [17]. The first one, called pump, creates the initial wave packet, while the second one, called probe, measures its time evolution. By changing the delay between the two pulses, different time scales can be explored. This pump–probe scheme was proposed by Alber, Ritsch, and Zoller [71], and was initially used to study atoms [72,73]. In this case, the second laser pulse was used to ionize the system and the photoionization signal was measured. The method was modified by Zewail (who was awarded the 1999 Nobel Prize in Chemistry for his studies in this field) and his group to study few-atom molecules, using the probe pulse to get the system to an upper fluorescent state and measure the fluorescence [74,75]. More recently, the scheme has been adapted so that the second pulse photoexcites the sample and the differential transmission spectra is analysed. This approach has made possible to study the wave packet evolution of CdSe quantum dots with a mean diameter of 6.4 nm (slightly larger than the diameter of the carbon nanoflake we have selected) [76].

3.2.1. Classical Time

We present in Figure 7 classical times determined both numerically (points) and analytically (lines) for the four geometries considered: fixed surface (red), fixed borders (blue), fixed vertices (green) and fixed surface (MM) (gray). Analytical estimations are presented as lines for clarity purposes, but they can only be calculated at the same points as numerical ones. We have simply linked two consecutive points with a straight line.
The agreement between numerical and analytical estimations is good for all geometries for low curvatures (below approximately 10 4 Å 2 ), but in two of the three self-consistent calculations, this agreement is lost for high curvatures. The reason is that, as we will see in the following subsection, T Re decreases as the curvature increases and it becomes less that one order of magnitude bigger than T Cl for the fixed surface and fixed border cases.
Recalling Equation (10), T Cl is proportional to the inverse of the first derivative of the energy with respect to the level index. This means classical time is related to the spreading of the energy spectrum. T Cl decreases as the separation among energy levels increases.
If we look at the analytical estimation for the classical times corresponding to the fixed surface and fixed border cases, they grow with curvature, while in the fixed vertices case this initial tendency breaks beyond 1.5 × 10 4 Å 2 and the graph becomes nearly flat. We have to remember that this third case corresponds to the geometry with lower restrictions and the system can adapt itself better to the deformation of its fixed points. This means that both total energy (see Figure 3) and energy level separation are less sensitive to curvature.
If we consider the numerical estimations for the three self-consistent geometries, T Cl remains essentially flat since the slight decrease in energy spectrum spreading is compensated by the interference from T Re . In the fixed surface case this interference overcomes the tendency of T Cl to grow and, in fact, it decreases for high curvatures.
Finally, if we concentrate on the non-self consistent fixed surface (MM) case, the behaviour is simpler. Numerical and analytical estimations agree perfectly except for the high curvature regime because, as we will see in the next subsection, T Re is much higher than T Cl . The spreading of the energy spectrum decreases monotonically with curvature and classical time increases in a linear way.

3.2.2. Revival Time

We show in Figure 8 revival times determined both numerically (points) and analytically (lines) for the four geometries considered: fixed surface (red), fixed borders (blue), fixed vertices (green) and fixed surface (MM) (gray). Analytical estimations are presented as lines for clarity purposes, but they can only be calculated at the same points as numerical ones. In this case, we have connected two consecutive points with a smoothed line.
In all cases, numerical and analytical estimates perfectly agree since superrevival times are much higher than revival ones and do not interfere with them.
According to Equation (11), T Re is proportional to the inverse of the second derivative of the energy with respect to the level index. This means revival time is related to the non-linearity of the energy spectrum. T Re increases as the spectrum gets closer to being linear (i.e., energy levels tend to be equally spaced).
Revival times for the fixed surface and fixed borders cases are very similar and decrease monotonically with curvature. T Re for the fixed vertices case coincides with them up to 1.5 × 10 4 Å 2 and from that point on it remains essentially constant. The situation is completely different for the fixed surface (MM) case. There is a pronounced peak around 8 × 10 5 Å 2 (in fact, T Re grows 12 orders of magnitude in this area and reaches several seconds). To check this feature is not an artifact, two of the authors wrote independently different codes (one in Scilab [77] and the other one in Mathematica [78]). The results obtained with both codes agreed.
Similar behaviours have been found in other 2D Dirac materials in external fields. For instance, planar graphene rings [26] and graphene flakes [27] in a perpendicular magnetic field or silicene in the presence of a perpendicular electric field [79]. In the latter case, the cause is a topological phase transition from a topological insulator to a band insulator at the charge neutrality point. However, to the best of our knowledge, this is the first time it appears in the absence of an external field. It is known that strain affects planar graphene revivals [24] and, as we have pointed out in the Introduction, graphene bending results in the appearance of a pseudo-magnetic field. It is possible that this pseudomagnetic field is responsible for this divergence in revival time for a particular radius. Taking into account that silicene is not a planar structure but a buckled one and that it exhibits this phenomenon when an electric field is applied, we plan to study in the future curved graphene flakes under an external electric field. This could provide a deeper understanding of the interplay between curvature effects and revival times in 2D nanostructures.

Author Contributions

Conceptualization, A.B. and N.A.C.; methodology, N.A.C.; software, S.d.-l.-H.-S. and N.A.C.; validation, S.d.-l.-H.-S. and N.A.C.; formal analysis, A.B. and N.A.C.; investigation, S.d.-l.-H.-S.; writing—original draft preparation, N.A.C.; writing—review and editing, A.B. and N.A.C.; visualization, S.d-l-H-S. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been partially supported by Agencia Estatal de Investigación (Spain) under grant PID2019-106802GB-I00/AEI/10.13039/501100011033 and by Spanish MICINN through the project PGC2018-097831-B-I00. S.d.-l.-H.-S. acknowledges support from Junta de Castilla y León and the European Social Fund through a predoctoral grant.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
DFTDensity Functional Theory
GGAGeneralized Gradient Approximation
HFHartree-Fock
HOMOHighest Occupied Molecular Orbital
LDALocal Density Approximation
MMMolecular Mechanics

References

  1. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; 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] [Green Version]
  2. Web of Science™. Available online: https://www.webofscience.com (accessed on 22 April 2022).
  3. Lalwani, G.; Henslee, A.M.; Farshid, B.; Lin, L.; Kasper, F.K.; Qin, Y.X.; Mikos, A.G.; Sitharaman, B. Two-Dimensional Nanostructure-Reinforced Biodegradable Polymeric Nanocomposites for Bone Tissue Engineering. Biomacromolecules 2013, 14, 900–909. [Google Scholar] [CrossRef] [Green Version]
  4. Dimov, D.; Amit, I.; Gorrie, O.; Barnes, M.D.; Townsend, N.J.; Neves, A.I.S.; Withers, F.; Russo, S.; Craciun, M.F. Ultrahigh Performance Nanoengineered Graphene–Concrete Composites for Multifunctional Applications. Adv. Funct. Mater. 2018, 28, 1705183. [Google Scholar] [CrossRef]
  5. Berman, D.; Erdemir, A.; Sumant, A.V. Reduced wear and friction enabled by graphene layers on sliding steel surfaces in dry nitrogen. Carbon 2013, 59, 167–175. [Google Scholar] [CrossRef]
  6. Berman, D.; Deshmukh, S.A.; Sankaranarayanan, S.K.R.S.; Erdemir, A.; Sumant, A.V. Extraordinary Macroscale Wear Resistance of One Atom Thick Graphene Layer. Adv. Funct. Mater. 2014, 24, 6640–6646. [Google Scholar] [CrossRef]
  7. Li, X.; Chen, W.; Zhang, S.; Wu, Z.; Wang, P.; Xu, Z.; Chen, H.; Yin, W.; Zhong, H.; Lin, S. 18.5% efficient graphene/GaAs van der Waals heterostructure solar cell. Nano Energy 2015, 16, 310–319. [Google Scholar] [CrossRef] [Green Version]
  8. 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]
  9. Zhao, J.; Zhang, J.; Yin, H.; Zhao, Y.; Xu, G.; Yuan, J.; Mo, X.; Tang, J.; Wang, F. Ultra-Fine Ruthenium Oxide Quantum Dots/Reduced Graphene Oxide Composite as Electrodes for High-Performance Supercapacitors. Nanomaterials 2022, 12, 1210. [Google Scholar] [CrossRef]
  10. Umar, M.; Nnadiekwe, C.C.; Haroon, M.; Abdulazeez, I.; Alhooshani, K.; Al-Saadi, A.A.; Peng, Q. A First-Principles Study on the Multilayer Graphene Nanosheets Anode Performance for Boron-Ion Battery. Nanomaterials 2022, 12, 1280. [Google Scholar] [CrossRef] [PubMed]
  11. Meng, S.; Wang, L.; Ji, X.; Yu, J.; Ma, X.; Zhang, J.; Zhao, W.; Ji, H.; Li, M.; Feng, H. Facile Gold-Nanoparticle Boosted Graphene Sensor Fabrication Enhanced Biochemical Signal Detection. Nanomaterials 2022, 12, 1327. [Google Scholar] [CrossRef]
  12. Askari, M.B.; Rozati, S.M.; Di Bartolomeo, A. Fabrication of Mn3O4-CeO2-rGO as Nanocatalyst for Electro-Oxidation of Methanol. Nanomaterials 2022, 12, 1187. [Google Scholar] [CrossRef] [PubMed]
  13. Oliveira, A.M.L.; Machado, M.; Silva, G.A.; Bitoque, D.B.; Tavares Ferreira, J.; Pinto, L.A.; Ferreira, Q. Graphene Oxide Thin Films with Drug Delivery Function. Nanomaterials 2022, 12, 1149. [Google Scholar] [CrossRef]
  14. Hu, S.; Lozada-Hidalgo, M.; Wang, F.; Mishchenko, A.; Schedin, F.; Nair, R.; Hill, E.; Boukhvalov, D.; Katsnelson, M.; Dryfe, R.; et al. Proton transport through one-atom-thick crystals. Nature 2014, 516, 227–230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Wu, X.; Wang, D.; Kang, X.; Zhang, D.; Yan, Y.; Guo, G.; Sun, Z.; Wang, X. Displacement Reaction-Assisted Synthesis of Sub-Nanometer Pt/Bi Boost Methanol-Tolerant Fuel Cells. Nanomaterials 2022, 12, 1301. [Google Scholar] [CrossRef] [PubMed]
  16. Mangum, J.M.; Harerimana, F.; Gikunda, M.N.; Thibado, P.M. Mechanisms of Spontaneous Curvature Inversion in Compressed Graphene Ripples for Energy Harvesting Applications via Molecular Dynamics Simulations. Membranes 2021, 11, 516. [Google Scholar] [CrossRef] [PubMed]
  17. Robinett, R. Quantum wave packet revivals. Phys. Rep. 2004, 392, 1–119. [Google Scholar] [CrossRef] [Green Version]
  18. Dooley, S.; Spiller, T.P. Fractional revivals, multiple-Schrödinger-cat states, and quantum carpets in the interaction of a qubit with N qubits. Phys. Rev. A 2014, 90, 012320. [Google Scholar] [CrossRef] [Green Version]
  19. Li, K.; Xia, F.; Wang, M.; Sun, P.; Liu, T.; Hu, W.; Kong, W.; Yun, M.; Dong, L. Discrete Talbot effect in dielectric graphene plasmonic waveguide arrays. Carbon 2017, 118, 192–199. [Google Scholar] [CrossRef]
  20. Kolaric, B.; Maes, B.; Clays, K.; Durt, T.; Caudano, Y. Strong Light–Matter Coupling as a New Tool for Molecular and Material Engineering: Quantum Approach. Adv. Quantum Technol. 2018, 1, 1800001. [Google Scholar] [CrossRef]
  21. Tsarev, M.V.; Ryabov, A.; Baum, P. Free-electron qubits and maximum-contrast attosecond pulses via temporal Talbot revivals. Phys. Rev. Res. 2021, 3, 043033. [Google Scholar] [CrossRef]
  22. Krueckl, V.; Kramer, T. Revivals of quantum wave packets in graphene. New J. Phys. 2009, 11, 093010. [Google Scholar] [CrossRef]
  23. Ali, R.; Saif, F. Chiral phase transition and quantum revivals in graphene. Mater. Res. Express 2015, 2, 095602. [Google Scholar] [CrossRef]
  24. Díaz-Bautista, E.; Betancur-Ocampo, Y. Phase-space representation of Landau and electron coherent states for uniaxially strained graphene. Phys. Rev. B 2020, 101, 125402. [Google Scholar] [CrossRef] [Green Version]
  25. López, A.; Di Teodoro, A.; Schliemann, J.; Berche, B.; Santos, B. Laser-induced modulation of the Landau level structure in single-layer graphene. Phys. Rev. B 2015, 92, 235411. [Google Scholar] [CrossRef] [Green Version]
  26. García, T.; Rodríguez-Bolívar, S.; Cordero, N.A.; Romera, E. Wavepacket revivals in monolayer and bilayer graphene rings. J. Phy. Condens. Matter 2013, 25, 235301. [Google Scholar] [CrossRef]
  27. García, T.; Cordero, N.A.; Romera, E. Zitterbewegung and quantum revivals in monolayer graphene quantum dots in magnetic fields. Phys. Rev. B 2014, 89, 075416. [Google Scholar] [CrossRef]
  28. Peierls, R. Quelques propriétés typiques des corps solides. Ann. L’inst. Henri Poincare 1935, 5, 177–222. [Google Scholar]
  29. Landau, L.D. On the theory of phase transitions. Zh. Eksp. Teor. Fiz. 1937, 7, 19–32, English Translation in Ukr. J. Phys. 2008, 53, 25–35. [Google Scholar] [CrossRef]
  30. Meyer, J.; Geim, A.; Katsnelson, M.; Novoselov, K.; Obergfell, D.; Roth, S.; Girit, C.; Zettl, A. On the roughness of single- and bi-layer graphene membranes. Solid State Commun. 2007, 143, 101–109. [Google Scholar] [CrossRef] [Green Version]
  31. Meyer, J.C.; Geim, A.K.; Katsnelson, M.I.; Novoselov, K.S.; Booth, T.J.; Roth, S. The structure of suspended graphene sheets. Nature 2007, 446, 60–63. [Google Scholar] [CrossRef] [PubMed]
  32. Deng, S.; Berry, V. Wrinkled, rippled and crumpled graphene: An overview of formation mechanism, electronic properties, and applications. Mater. Today 2016, 19, 197–212. [Google Scholar] [CrossRef]
  33. Fasolino, A.; Los, J.; Katsnelson, M. Intrinsic ripples in graphene. Nat. Mater. 2007, 6, 858–861. [Google Scholar] [CrossRef] [Green Version]
  34. Carlsson, J. Buckle or break. Nat. Mater. 2007, 6, 801–802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Lundeberg, M.B.; Folk, J.A. Rippled Graphene in an In-Plane Magnetic Field: Effects of a Random Vector Potential. Phys. Rev. Lett. 2010, 105, 146804. [Google Scholar] [CrossRef] [Green Version]
  36. Los, J.H.; Ghiringhelli, L.M.; Meijer, E.J.; Fasolino, A. Improved long-range reactive bond-order potential for carbon. I. Construction. Phys. Rev. B 2005, 72, 214102. [Google Scholar] [CrossRef] [Green Version]
  37. Ishigami, M.; Chen, J.H.; Cullen, W.G.; Fuhrer, M.S.; Williams, E.D. Atomic Structure of Graphene on SiO2. Nano Lett. 2007, 7, 1643–1648. [Google Scholar] [CrossRef] [Green Version]
  38. Geringer, V.; Liebmann, M.; Echtermeyer, T.; Runte, S.; Schmidt, M.; Rückamp, R.; Lemme, M.C.; Morgenstern, M. Intrinsic and extrinsic corrugation of monolayer graphene deposited on SiO2. Phys. Rev. Lett. 2009, 102, 076102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Stolyarova, E.; Stolyarov, D.; Bolotin, K.; Ryu, S.; Liu, L.; Rim, K.T.; Klima, M.; Hybertsen, M.; Pogorelsky, I.; Pavlishin, I.; et al. Observation of Graphene Bubbles and Effective Mass Transport under Graphene Films. Nano Lett. 2009, 9, 332–337. [Google Scholar] [CrossRef]
  40. Leconte, N.; Kim, H.; Kim, H.J.; Ha, D.H.; Watanabe, K.; Taniguchi, T.; Jung, J.; Jung, S. Graphene bubbles and their role in graphene quantum transport. Nanoscale 2017, 9, 6041–6047. [Google Scholar] [CrossRef]
  41. Georgiou, T.; Britnell, L.; Blake, P.; Gorbachev, R.V.; Gholinia, A.; Geim, A.K.; Casiraghi, C.; Novoselov, K.S. Graphene bubbles with controllable curvature. Appl. Phys. Lett. 2011, 99, 093103. [Google Scholar] [CrossRef] [Green Version]
  42. Huamán, A.; Foa Torres, L.E.F.; Balseiro, C.A.; Usaj, G. Quantum Hall edge states under periodic driving: A Floquet induced chirality switch. Phys. Rev. Res. 2021, 3, 013201. [Google Scholar] [CrossRef]
  43. Ornigotti, M.; Ornigotti, L.; Biancalana, F. Generation of half-integer harmonics and efficient THz-to-visible frequency conversion in strained graphene. APL Photon. 2021, 6, 060801. [Google Scholar] [CrossRef]
  44. Sinha, D.; Berche, B. Quantum oscillations and wave packet revival in conical graphene. Eur. Phys. J. B 2016, 89, 57. [Google Scholar] [CrossRef] [Green Version]
  45. Belouad, A.; Jellal, A.; Bahlouli, H. Electronic properties of graphene quantum ring with wedge disclination. Eur. Phys. J. B 2021, 94, 1–10. [Google Scholar] [CrossRef]
  46. Wallace, P.R. The Band Theory of Graphite. Phys. Rev. 1947, 71, 622–634. [Google Scholar] [CrossRef]
  47. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef] [Green Version]
  48. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef] [Green Version]
  49. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: A critical analysis. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar] [CrossRef] [Green Version]
  50. 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, 2009; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  51. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, 2016; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  52. Girifalco, L.A.; Hodak, M. Van der Waals binding energies in graphitic structures. Phys. Rev. B 2002, 65, 125404. [Google Scholar] [CrossRef]
  53. Hasegawa, M.; Nishidate, K. Semiempirical approach to the energetics of interlayer binding in graphite. Phys. Rev. B 2004, 70, 205431. [Google Scholar] [CrossRef] [Green Version]
  54. Torres-Rojas, R.M.; Contreras-Solorio, D.A.; Hernández, L.; Enciso, A. Band gap variation in bi, tri and few-layered 2D graphene/hBN heterostructures. Solid State Commun. 2022, 341, 114553. [Google Scholar] [CrossRef]
  55. Khantha, M.; Cordero, N.A.; Molina, L.M.; Alonso, J.A.; Girifalco, L.A. Interaction of lithium with graphene: An ab initio study. Phys. Rev. B 2004, 70, 125422. [Google Scholar] [CrossRef]
  56. Cordero, N.A.; Alonso, J.A. The interaction of sulfuric acid with graphene and formation of adsorbed crystals. Nanotechnology 2007, 18, 485705. [Google Scholar] [CrossRef]
  57. Khantha, M.; Cordero, N.A.; Alonso, J.A.; Cawkwell, M.; Girifalco, L.A. Interaction and concerted diffusion of lithium in a (5,5) carbon nanotube. Phys. Rev. B 2008, 78, 115430. [Google Scholar] [CrossRef] [Green Version]
  58. Cordero, N.A.; Alonso, J.A. Interaction of Surfactants Containing a Sulfuric Group with a (5,5) Carbon Nanotube. J. Phys. Chem. C 2010, 114, 17249–17256. [Google Scholar] [CrossRef]
  59. Ayala, I.G.; Cordero, N.A.; Alonso, J.A. Surfactant effect of sulfuric acid on the exfoliation of bilayer graphene. Phys. Rev. B 2011, 84, 165424. [Google Scholar] [CrossRef]
  60. Ayala, I.G.; Cordero, N.A. Interaction of sodium bisulfate with mono- and bi-layer graphene. J. Nanopart. Res. 2012, 14, 1071. [Google Scholar] [CrossRef]
  61. Zhao, F.; Cao, T.; Louie, S.G. Topological Phases in Graphene Nanoribbons Tuned by Electric Fields. Phys. Rev. Lett. 2021, 127, 166401. [Google Scholar] [CrossRef]
  62. Ditchfield, R.; Hehre, W.J.; Pople, J.A. Self-Consistent Molecular-Orbital Methods. IX. An Extended Gaussian-Type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724–728. [Google Scholar] [CrossRef]
  63. Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self—Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian—Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1972, 56, 2257–2261. [Google Scholar] [CrossRef]
  64. Hariharan, P.; Pople, J.A. The influence of polarization functions on molecular orbital hydrogenation energies. Theor. Chim. Acta 1973, 28, 213–222. [Google Scholar] [CrossRef]
  65. Hariharan, P.; Pople, J. Accuracy of AHn equilibrium geometries by single determinant molecular orbital theory. Mol. Phys. 1974, 27, 209–214. [Google Scholar] [CrossRef]
  66. Gordon, M.S. The isomers of silacyclopropane. Chem. Phys. Lett. 1980, 76, 163–168. [Google Scholar] [CrossRef]
  67. Francl, M.M.; Pietro, W.J.; Hehre, W.J.; Binkley, J.S.; Gordon, M.S.; DeFrees, D.J.; Pople, J.A. Self-consistent molecular orbital methods. XXIII. A polarization-type basis set for second-row elements. J. Chem. Phys. 1982, 77, 3654–3665. [Google Scholar] [CrossRef] [Green Version]
  68. Sun, Q.; Yao, X.; Gröning, O.; Eimre, K.; Pignedoli, C.A.; Müllen, K.; Narita, A.; Fasel, R.; Ruffieux, P. Coupled Spin States in Armchair Graphene Nanoribbons with Asymmetric Zigzag Edge Extensions. Nano Lett. 2020, 20, 6429–6436. [Google Scholar] [CrossRef]
  69. Dennington, R.; Keith, T.A.; Millam, J.M. GaussView Version 6, 2019; Semichem Inc.: Shawnee Mission, KS, USA, 2019. [Google Scholar]
  70. Zhou, Y.; Chen, Y.; Liu, B.; Wang, S.; Yang, Z.; Hu, M. Mechanics of nanoscale wrinkling of graphene on a non-developable surface. Carbon 2015, 84, 263–271. [Google Scholar] [CrossRef]
  71. Alber, G.; Ritsch, H.; Zoller, P. Generation and detection of Rydberg wave packets by short laser pulses. Phys. Rev. A 1986, 34, 1058–1064. [Google Scholar] [CrossRef]
  72. Yeazell, J.A.; Mallalieu, M.; Stroud, C.R. Observation of the collapse and revival of a Rydberg electronic wave packet. Phys. Rev. Lett. 1990, 64, 2007–2010. [Google Scholar] [CrossRef]
  73. Wals, J.; Fielding, H.H.; Christian, J.F.; Snoek, L.C.; van der Zande, W.J.; van Linden van den Heuvell, H.B. Observation of Rydberg wave packet dynamics in a Coulombic and magnetic field. Phys. Rev. Lett. 1994, 72, 3783–3786. [Google Scholar] [CrossRef] [Green Version]
  74. Dantus, M.; Rosker, M.J.; Zewail, A.H. Real-time femtosecond probing of “transition states” in chemical reactions. J. Chem. Phys. 1987, 87, 2395–2397. [Google Scholar] [CrossRef] [Green Version]
  75. Bowman, R.; Dantus, M.; Zewail, A. Femtosecond transition-state spectroscopy of iodine: From strongly bound to repulsive surface dynamics. Chem. Phys. Lett. 1989, 161, 297–302. [Google Scholar] [CrossRef]
  76. Dong, S.; Trivedi, D.; Chakrabortty, S.; Kobayashi, T.; Chan, Y.; Prezhdo, O.V.; Loh, Z.H. Observation of an Excitonic Quantum Coherence in CdSe Nanocrystals. Nano Lett. 2015, 15, 6875–6882. [Google Scholar] [CrossRef] [PubMed]
  77. Scilab. Available online: https://www.scilab.org/ (accessed on 22 April 2022).
  78. Mathematica. 2021. Available online: https://www.wolfram.com/mathematica (accessed on 22 April 2022).
  79. Bolívar, J.C.; Cordero, N.A.; Nagy, A.; Romera, E. Fidelity as a marker of topological phase transitions in 2D Dirac materials. Int. J. Quantum Chem. 2018, 118, e25674. [Google Scholar] [CrossRef]
Figure 1. Graphene flake used in the calculations (image generated using GausView 6 [69]).
Figure 1. Graphene flake used in the calculations (image generated using GausView 6 [69]).
Nanomaterials 12 01953 g001
Figure 2. Optimized geometries for a (quasi-)spherical graphene flake of radius 40 Å (image generated using GausView 6 [69]).
Figure 2. Optimized geometries for a (quasi-)spherical graphene flake of radius 40 Å (image generated using GausView 6 [69]).
Nanomaterials 12 01953 g002
Figure 3. Energy of a (quasi-)spherical graphene flake as a function of its curvature. Flat configuration is taken as energy origin. Lines are merely guides for the eye.
Figure 3. Energy of a (quasi-)spherical graphene flake as a function of its curvature. Flat configuration is taken as energy origin. Lines are merely guides for the eye.
Nanomaterials 12 01953 g003
Figure 4. Log–log plot of the energy of a spherical graphene flake as a function of its curvature. Flat configuration is taken as energy origin. Lines are merely guides for the eye.
Figure 4. Log–log plot of the energy of a spherical graphene flake as a function of its curvature. Flat configuration is taken as energy origin. Lines are merely guides for the eye.
Nanomaterials 12 01953 g004
Figure 5. Value of n in Equation (2) as a function of the radius of the spherical carbon nanoflake calculated with the hybrid MM/DFT approximation.
Figure 5. Value of n in Equation (2) as a function of the radius of the spherical carbon nanoflake calculated with the hybrid MM/DFT approximation.
Nanomaterials 12 01953 g005
Figure 6. A simple example of the time evolution of the squared modulus of the autocorrelation function in blue with its upper envelope in orange. Classical time is marked in red and revival time in green.
Figure 6. A simple example of the time evolution of the squared modulus of the autocorrelation function in blue with its upper envelope in orange. Classical time is marked in red and revival time in green.
Nanomaterials 12 01953 g006
Figure 7. Classical times for a curved graphene nanoflake. Points correspond to numerical values while dotted lines represent analytical ones.
Figure 7. Classical times for a curved graphene nanoflake. Points correspond to numerical values while dotted lines represent analytical ones.
Nanomaterials 12 01953 g007
Figure 8. Revival times for a curved graphene nanoflake. Points correspond to numerical values while dotted lines represent analytical ones.
Figure 8. Revival times for a curved graphene nanoflake. Points correspond to numerical values while dotted lines represent analytical ones.
Nanomaterials 12 01953 g008
Table 1. Results of the fitting of the data in Figure 5 to the Padé approximant in Equation (3).
Table 1. Results of the fitting of the data in Figure 5 to the Padé approximant in Equation (3).
a 0 a 1 / Å 1 b 1 / Å 1
−0.18180.04430.0219
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

de-la-Huerta-Sainz, S.; Ballesteros, A.; Cordero, N.A. Quantum Revivals in Curved Graphene Nanoflakes. Nanomaterials 2022, 12, 1953. https://doi.org/10.3390/nano12121953

AMA Style

de-la-Huerta-Sainz S, Ballesteros A, Cordero NA. Quantum Revivals in Curved Graphene Nanoflakes. Nanomaterials. 2022; 12(12):1953. https://doi.org/10.3390/nano12121953

Chicago/Turabian Style

de-la-Huerta-Sainz, Sergio, Angel Ballesteros, and Nicolás A. Cordero. 2022. "Quantum Revivals in Curved Graphene Nanoflakes" Nanomaterials 12, no. 12: 1953. https://doi.org/10.3390/nano12121953

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