Quantum Dynamics in a 1D Dot/Antidot Lattice: Landau Minibands and Graphene Wave Packet Motion in a Magnetic Field

  • Chapter
  • First Online:
Progress in Nanoscale and Low-Dimensional Materials and Devices

Part of the book series: Topics in Applied Physics ((TAP,volume 144))

  • 710 Accesses

Abstract

This work is focused on the analysis of the quantum dynamics of a model one-dimensional lattice array of quantum dots on a two dimensional electron sheet/layer in a normal magnetic field. Our analysis is carried out with the derivation of the Green’s function for the quantum dot lattice subject to Landau quantization using the corresponding “no-lattice” Green’s function, for propagation along the axis of the 1D quantum dot lattice. The frequency/energy poles of this Green’s function provide the dispersion relation for the Landau quantized energy spectrum, which exhibits Landau minibands, rather than discrete Landau levels. In the case of nonrelativistic carriers, the dispersion relation is explicitly exhibited in a closed form in terms of the Jacobi Theta Function of the third kind, and an approximate solution displaying the Landau miniband formation is obtained in terms of Laguerre polynomials. We also examine the case of “relativistic” graphene carriers, employing the appropriate Landau quantized graphene Green’s function in a study of the associated wave packet dynamics in a graphene antidot lattice in a normal magnetic field. In this, we analyze the effects of pseudospin on the wave packet dynamics in Landau quantized graphene, including the role of Zitterbewegung in circular orbit motion.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free ship** worldwide - see info
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free ship** worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

References

  1. L.L. Sohn, L.P. Kouwenhoven, G. Schoen, Mesoscopic Electron Transport (Springer and Kluwer, Netherlands, 1997)

    Book  Google Scholar 

  2. C.R. Kagan, C.B. Murray, Charge transport in strongly coupled quantum dot solids. Nat. Nanotechnol. 10(12), 1013–1026 (2015)

    Article  CAS  Google Scholar 

  3. V.P. Kunets, M. Rebello Sousa Dias, T. Rembert, M.E. Ware, Y.I. Mazur, V. Lopez-Richard, H.A. Mantooth, G.E. Marques, G.J. Salamo, Electron transport in quantum dot chains: dimensionality effects and hop** conductance. J. Appl. Phys. 113(18), 183709 (2013)

    Google Scholar 

  4. T. Jamieson, R. Bakhshi, D. Petrova, R. Pocock, M. Imani, A.M. Seifalian, Biological applications of quantum dots. Biomaterials 28(31), 4717–4732 (2007)

    Article  CAS  Google Scholar 

  5. L. Qi, X. Gao, Emerging application of quantum dots for drug delivery and therapy. Expert Opin. Drug Deliv. 5(3), 263–267 (2008). (PMID: 18318649)

    Google Scholar 

  6. National Research Council, High Magnetic Field Science and Its Application in the United States: Current Status and Future Directions (The National Academies Press, Washington, D.C., 2013)

    Google Scholar 

  7. D. Lai, Matter in strong magnetic fields. Rev. Mod. Phys. 73, 629–662 (2001)

    Article  CAS  Google Scholar 

  8. N.J.M. Horing, S. Bahrami, Landau minibands in an antidot lattice. AdvNanoBio M & D 1(1), 24 (2017)

    Google Scholar 

  9. N.J. Morgenstern Horing, S.Y. Liu, Green’s functions for a graphene sheet and quantum dot in a normal magnetic field. J. Phys. A: Math. Theor. 42(22), 225301 (2009)

    Google Scholar 

  10. ** Chen, V.I. Fal’ko, Hierarchy of gaps and magnetic minibands in graphene in the presence of the Abrikosov vortex lattice. Phys. Rev. B 93, 035427 (2016)

    Google Scholar 

  11. C. Kittel. Introduction to Solid State Physics, 7th edn. (Wiley, 1991)

    Google Scholar 

  12. N.J. Morgenstern Horing, M.M. Yildiz, Quantum theory of longitudinal dielectric response properties of a two-dimensional plasma in a magnetic field. Ann. Phys. 97(1), 216–241 (1976)

    Google Scholar 

  13. A. Erdelyi, H. Bateman, Higher Transcendental Functions, vol. 2 (McGraw-Hill, 1953)

    Google Scholar 

  14. T. Ando, Theory of electronic states and transport in carbon nanotubes. J. Phys. Soc. Jpn. 74(3), 777–817 (2005)

    Article  CAS  Google Scholar 

  15. Y. Zheng, T. Ando, Hall conductivity of a two-dimensional graphite system. Phys. Rev. B 65, 245420 (2002)

    Article  Google Scholar 

  16. N.J.M. Horing, Landau quantized dynamics and spectra for group-vi dichalcogenides, including a model quantum wire. AIP Adv. 7(6), 065316 (2017)

    Article  Google Scholar 

  17. N.J.M. Horing, Addendum: “Landau quantized dynamics and spectra for group-vi dichalcogenides, including a model quantum wire” [AIP Advances 7, 065316 (2017)]. AIP Adv. 8(4), 049901 (2018)

    Google Scholar 

  18. Q. Wang, R. Shen, L. Sheng, B.G. Wang, D.Y. **ng, Transient zitterbewegung of graphene superlattices. Phys. Rev. A 89, 022121 (2014)

    Article  Google Scholar 

  19. G.M. Maksimova, V.Y. Demikhovskii, E.V. Frolova, Wave packet dynamics in a monolayer graphene. Phys. Rev. B 78, 235321 (2008)

    Article  Google Scholar 

  20. V.Y. Demikhovskii, G.M. Maksimova, E.V. Frolova, Wave packet dynamics in a two-dimensional electron gas with spin orbit coupling: Splitting and zitterbewegung. Phys. Rev. B 78, 115401 (2008)

    Article  Google Scholar 

  21. R.A.W. Ayyubi, N.J.M. Horing, K. Sabeeh, Effect of pseudospin polarization on wave packet dynamics in graphene antidot lattices (gals) in the presence of a normal magnetic field. J. Appl. Phys. 129(7), 074301 (2021)

    Article  CAS  Google Scholar 

  22. D. Song, V. Paltoglou, S. Liu, Y. Zhu, D. Gallardo, L. Tang, J. Xu, M. Ablowitz, N.K. Efremidis, Z. Chen, Unveiling pseudospin and angular momentum in photonic graphene. Nat. Commun. 6, 6272 (2015)

    Article  CAS  Google Scholar 

  23. Tomasz M. Rusin, Wlodek Zawadzki, Zitterbewegung of electrons in graphene in a magnetic field. Phys. Rev. B 78, 125419 (2008)

    Google Scholar 

  24. E. Serna, I. Rodríguez Vargas, R. Pérez-Álvarez, L. Diago-Cisneros, Pseudospin-dependent zitterbewegung in monolayer graphene. J. Appl. Phys. 125(20), 203902 (2019)

    Google Scholar 

  25. L. Esaki, R. Tsu, Superlattice and negative differential conductivity in semiconductors. IBM J. Res. Dev. 14(1), 61–65 (1970)

    Article  CAS  Google Scholar 

  26. C. Heide, T. Higuchi, H.B. Weber, P. Hommelhoff, Coherent electron trajectory control in graphene. Phys. Rev. Lett. 121, 207401 (2018)

    Article  CAS  Google Scholar 

  27. A.H. Castro Neto, F. Guinea, N.M.R. Peres, K.S. Novoselov, A.K. Geim, The electronic properties of graphene. Rev. Mod. Phys. 81, 109–162 (2009)

    Google Scholar 

  28. C.H. Park, Y.W. Son, L. Yang, M.L. Cohen, S.G. Louie, Electron beam supercollimation in graphene superlattices. Nano Lett. 8(9), 2920 (2008)

    Article  CAS  Google Scholar 

  29. S. Choi, C.-H. Park, S.G. Louie, Electron supercollimation in graphene and Dirac fermion materials using one-dimensional disorder potentials. Phys. Rev. Lett. 113, 026802 (2014)

    Google Scholar 

  30. C.-H. Park, L. Yang, Y.-W. Son, M.L. Cohen, S.G. Louie, Anisotropic behaviours of massless Dirac fermions in graphene under periodic potentials (2008)

    Google Scholar 

  31. T.G. Pedersen, C. Flindt, J. Pedersen, N.A. Mortensen, A.-P. Jauho, K. Pedersen, Graphene antidot lattices: designed defects and spin qubits. Phys. Rev. Lett. 100, 136804 (2008)

    Article  Google Scholar 

  32. A.J.M. Giesbers, E.C. Peters, M. Burghard, K. Kern, Charge transport gap in graphene antidot lattices. Phys. Rev. B 86, 045445 (2012)

    Article  Google Scholar 

  33. N.J. Morgenstern Horing, Dichalcogenide Landau miniband dynamics and spectrum in an antidot superlattice. AIP Adv. 10(3), 035203 (2020)

    Google Scholar 

  34. T. Danz, A. Neff, J.H. Gaida, R. Bormann, C. Ropers, S. Schäfer, Ultrafast sublattice pseudospin relaxation in graphene probed by polarization-resolved photoluminescence. Phys. Rev. B 95, 241412 (2017)

    Article  Google Scholar 

  35. S. Aeschlimann, R. Krause, M. Chávez-Cervantes, H. Bromberger, R. Jago, E. Malić, A. Al-Temimy, C. Coletti, A. Cavalleri, I. Gierz, Ultrafast momentum imaging of pseudospin-flip excitations in graphene. Phys. Rev. B 96, 020301 (2017)

    Article  Google Scholar 

  36. M. Trushin, A. Grupp, G. Soavi, A. Budweg, D. De Fazio, U. Sassi, A. Lombardo, A.C. Ferrari, W. Belzig, A. Leitenstorfer, D. Brida, Ultrafast pseudospin dynamics in graphene. Phys. Rev. B 92, 165429 (2015)

    Article  Google Scholar 

  37. Chen-Di Han, Hong-Ya Xu, Ying-Cheng Lai, Pseudospin modulation in coupled graphene systems. Phys. Rev. Research 2, 033406 (2020)

    Google Scholar 

  38. M. Polini, F. Guinea, M. Lewenstein, H. Manoharan, V. Pellegrini, Artificial honeycomb lattices for electrons, atoms and photons. Nat. Nanotechnol. 8, 625–633 (2013)

    Article  CAS  Google Scholar 

  39. K. Gomes, W. Mar, W. Ko, F. Guinea, H. Manoharan, Designer Dirac fermions and topological phases in molecular graphene. Nature 483, 306–310 (2012)

    Google Scholar 

  40. O. Bahat-Treidel, O. Peleg, M. Grobman, N. Shapira, M. Segev, T. Pereg-Barnea, Klein tunneling in deformed honeycomb lattices. Phys. Rev. Lett. 104, 063901 (2010)

    Article  Google Scholar 

  41. P. Soltan-Panahi, J. Struck, P. Hauke, A. Bick, W. Plenkers, G. Meineke, C. Becker, P. Windpassinger, M. Lewenstein, K. Sengstock, Multi-component quantum gases in spin-dependent hexagonal lattices. Nat. Phys. 7, 05 (2010)

    Google Scholar 

  42. Tomasz M. Rusin, Wlodek Zawadzki, Transient zitterbewegung of charge carriers in mono- and bilayer graphene, and carbon nanotubes. Phys. Rev. B 76, 195439 (2007)

    Google Scholar 

  43. A.O. Barut, A.J. Bracken, Zitterbewegung and the internal geometry of the electron. Phys. Rev. D 23, 2454–2463 (1981)

    Article  CAS  Google Scholar 

  44. F. Schwabl, Advanced Quantum Mechanics, 4th edn. (Springer, Berlin, 2008)

    Google Scholar 

Download references

Acknowledgement

This chapter replicates some work presented in our earlier publication “Landau Minibands in an Antidot Lattice” in AdvNanoBio M & D 1 (1), 24 (2017) [8]. The appearance of the work in this chapter is facilitated by the gracious consent of the journal editor, Prof. Dr. Anton Ficai of SciEdTech.eu.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Norman J. M. Horing .

Editor information

Editors and Affiliations

Appendices

Appendix A

16.11 Zitterbewegung Phenomenon

Zitterbewegung is the jittery motion of the Dirac electron. It occurs when one tries to confine the Dirac electrons. According to the Heisenberg uncertainty principle, localization of the electron wave packet leads to uncertainty in momentum. For particles with zero rest mass (massless Dirac particles), uncertainty in momentum translates into uncertainty in energy of the particle (This should be contrasted with the nonrelativistic case, where the position-momentum uncertainty relation is independent of the energy-time uncertainty relation) [27].

16.1.1 16.11.1 Prediction and Interpretation of Zitterbewegung by Schrödinger

Schrödinger discovered a highly oscillatory motion of the electron with velocity c during his work on the time evolution of the position operator, which he named Zitterbewegung [43]. Schrödinger attempted to explain this phenomenon in terms of microscopic dynamical variables i.e coordinate and momentum. To determine the time evolution of the position operator, Schrödinger used Dirac’s Hamiltonian for the free electron-positron system, which is

$$\begin{aligned} H=c \mathbf {\alpha } . \textbf{q} + m c^2 \beta . \,\,\left[ \beta =\begin{pmatrix} I &{} 0\\ 0 &{} -I \end{pmatrix}\right] \end{aligned}$$
(16.77)

In the above equation \(\mathbf {\alpha }\) and \(\beta \) satisfy the following anti-commutation relations:

$$\begin{aligned} \{ \mathbf {\alpha }_i,\mathbf {\alpha }_j\}= 2 \delta _{ij} I; \qquad \{ \mathbf {\alpha }_i,\beta \}= 0; \qquad \beta ^2 =I, \end{aligned}$$

where \(i,j=1,2,3\) and I is a unit matrix. Also the momentum operator \(\textbf{q}\) and coordinate operator \(\textbf{x}\) commute with \(\mathbf {\alpha }\) and \(\beta \), and satisfy the canonical anti-commutation relations for Fermions

$$\begin{aligned} \{ \textbf{x}_i,\textbf{x}_j\}=\{ \textbf{q}_i,\textbf{q}_j\} = 0 \qquad and \qquad \{ \textbf{x}_i,\textbf{q}_j\}= i\hbar \delta _{ij} I . \end{aligned}$$

In the Heisenberg picture, the time derivative of any one of these operators, say A, which does not have explicit time dependence is given by

$$\begin{aligned} \frac{dA}{dt} =\frac{ i}{\hbar } [H,A]. \end{aligned}$$
(16.78)

As a result

$$\begin{aligned} \frac{d\textbf{q}}{dt} = 0 , \qquad \frac{d H}{dt} = 0, \qquad \frac{d\textbf{x}}{dt} = c\mathbf {\alpha }, \end{aligned}$$
(16.79)

and (16.77) and (16.78) for the operator \(\mathbf {\alpha }\) give

$$\begin{aligned} -i\hbar \frac{d\mathbf {\alpha }}{dt} = [H,\mathbf {\alpha }]= 2H\alpha -\{H,\mathbf {\alpha }\} = 2H \mathbf {\alpha }- 2c\textbf{q}. \end{aligned}$$
(16.80)

Above equation can be written as

$$\begin{aligned} -i \hbar \frac{d\mathbf {\alpha }}{dt} = 2H \mathbf {\eta }, \end{aligned}$$

where we have defined

$$\begin{aligned} \mathbf {\eta } =\mathbf {\alpha } - c H^{-1} \textbf{q}. \end{aligned}$$
(16.81)

Schrödinger noted that

$$\begin{aligned} -i \hbar \frac{d \mathbf {\eta }}{dt} = -i \hbar \frac{d \mathbf {\alpha }}{dt} = 2H \mathbf {\eta }, \end{aligned}$$

which yields

$$\begin{aligned} \mathbf {\eta } (t) = e^{\frac{2iHt}{\hbar }}\mathbf {\eta }_0 \qquad with \qquad \mathbf {\eta }_0(0)= \mathbf {\alpha }(0) - c H^{-1} \textbf{q}. \end{aligned}$$
(16.82)

Using \(\{H,\mathbf {\eta }\}=0=\{H,\mathbf {\eta }_0\}\), (16.82) can also be written as

$$\begin{aligned} \mathbf {\eta }(t) = \mathbf {\eta }_0 e^{\frac{-2iHt}{\hbar }}. \end{aligned}$$
(16.83)

Combining (16.80), (16.81) and (16.83), Schrödinger obtained

$$\begin{aligned} \frac{d \textbf{x}}{dt} = c \mathbf {\alpha } = c^2 H^{-1} \textbf{q} + c \mathbf {\eta }_0 e^{\frac{-2iHt}{\hbar }}, \end{aligned}$$

which on integration yields

$$\begin{aligned} \textbf{x}(t) = \textbf{a} + c^2 H^{-1} \textbf{q} t + \frac{i}{\hbar } c \mathbf {\eta }_0 H^{-1} e^{\frac{-2iHt}{\hbar }}, \end{aligned}$$
(16.84)

where \(\textbf{a}\) is an operator which is a constant of integration with the definition

$$\begin{aligned} \textbf{a} = \textbf{x}(0) - \frac{i}{2} \hbar c^2 H^{-2} \textbf{q}. \end{aligned}$$
(16.85)

Defining

$$\begin{aligned} \textbf{x}_A(t) = \textbf{a} + c^2 H^{-1} \textbf{q} t, \end{aligned}$$

equation (16.84) will become

$$\begin{aligned} \textbf{x}(t) = \textbf{x}_A(t) + \mathbf {\xi } (t), \end{aligned}$$

where we have defined

$$\begin{aligned} \mathbf {\xi } (t) = \frac{i}{2} \hbar c \mathbf {\eta }_0 H^{-1} e^{\frac{-2iHt}{\hbar }} = \frac{i}{2} \hbar c \mathbf {\eta } H^{-1}. \end{aligned}$$
(16.86)

Here, the term \(\mathbf {\xi } (t)\) corresponds to a microscopic coordinate which oscillates at high frequency known as Zitterbewegung (ZB). This motion is superimposed on a macroscopic type of motion associated with the coordinate \(\textbf{x}_A\). Characteristic amplitude associated with Zitterbewegung is \(\frac{\hbar }{2mc}\) and this amplitude is equal to half of the Compton wavelength of an electron, while the characteristic angular frequency of the Zitterbewegung is \(\frac{2mc^2}{\hbar }\).

Schrödinger went on to note, that if the orbital-angular momentum \(\textbf{L}\) and spin vector \(\textbf{S}\) are introduced as

$$\begin{aligned} \textbf{S} = -\frac{i}{4}\hbar \mathbf {\alpha } \times \mathbf {\alpha } \qquad and \qquad \textbf{L} =\textbf{x} \times \textbf{q} \end{aligned}$$

then both \(\textbf{L}\) and \(\textbf{S}\) alone are not constant but their sum \(\textbf{L} + \textbf{S}\) is a constant of the motion. Schrödinger found

$$\begin{aligned} \textbf{S}(t) = \textbf{S}_A - \mathbf {\eta } (t) \times \textbf{q} \end{aligned}$$

and

$$\begin{aligned} \textbf{L}(t) = \textbf{L}_A + \mathbf {\eta } (t) \times \textbf{q}, \end{aligned}$$

where \(\textbf{S}_A\) and \(\textbf{L}_A\) are constants and the term \(\mathbf {\eta } (t) \times \textbf{q}\) is oscillatory. Finally he found that

$$\begin{aligned} \textbf{L} + \textbf{S} = \textbf{L}_A + \textbf{S}_A . \end{aligned}$$
(16.87)

This constant of the motion allowed Schrödinger to explain Zitterbewegung in terms of microscopic dynamical variables spin and orbital angular momentum [43].

16.1.2 16.11.2 Zitterbewegung: Interpretation in Terms of Interference Between Positive and Negative Energy States

It can be easily verified that (16.84) along with (16.82) and (16.85) can be written as

$$\begin{aligned} \textbf{x} (t) = \textbf{x} (0)+ \frac{c^2 \textbf{q}}{H}t + \frac{\hbar c}{2 i H} \left( e^{\frac{2iHt}{\hbar }} -1 \right) \left( \mathbf {\alpha }(0) - \frac{c \textbf{q}}{H} \right) . \end{aligned}$$
(16.88)

Also (16.80) can be brought to the form

$$\begin{aligned} H\mathbf {\alpha } +\mathbf {\alpha } H = 2 c \textbf{q}, \end{aligned}$$

and hence

$$\begin{aligned} H \left( \mathbf {\alpha } - \frac{c \textbf{q}}{H} \right) - \left( \mathbf {\alpha } - \frac{c \textbf{q}}{H} \right) H = 0. \end{aligned}$$
(16.89)

Along with the initial term \(\textbf{x}(0)\) at \(t = 0\), (16.88) also carries two more terms, one of them is linear in time t and corresponds to group velocity, while the second term is oscillatory in nature and is known as Zitterbewegung. \(\mathbf {\alpha } - \frac{c \textbf{q}}{H}\) is an operator whose matrix elements are required to evaluate the average value \(\int \psi ^\dagger (0,\textbf{x}) \textbf{x}(t) \psi (0,\textbf{x}) d^3x\). Nonvanishing matrix elements of this operator only lie between states of the same momentum. Also from (16.89), the anticommutator only vanishes when the energies are of opposite sign. Hence it can be concluded that ZB is a consequence of interference between positive and negative energy states as predicted by relativistic quantum mechanics [44].

Appendix B

16.12 Contour Integration

To solve the second integral in (16.69), let us denote the integral by I [21]

$$\begin{aligned} I = \frac{\alpha d \omega _g}{2\pi } \sum _{n^\prime =-\infty }^\infty \int _{-\infty }^\infty d\eta e^{- i \omega _{g}\eta t}&\mathcal {G}^0(x_1;n^\prime d;\eta ) \int _{-\frac{\pi }{d}}^{\frac{\pi }{d}} dp e^{- ipn^\prime d} \nonumber \\&\times \left[ I_2-\alpha \dot{\tilde{\mathcal {G}}}^0(p;0,0;\eta )\right] ^{-1} \tilde{\mathcal {G}}^0(p;x_2;\eta ). \end{aligned}$$
(16.90)

(\(I_2\) is unit matrix of order 2) where

$$\begin{aligned} \eta = \frac{\omega }{\omega _g} \qquad and \qquad d\omega = \omega _g d\eta . \end{aligned}$$

Note that in the above integral, each “no lattice” Green’s function has real poles at \(\eta \)=\(\pm \sqrt{n}\). These “no lattice” poles of the term

$$\begin{aligned} T_1 (\eta ) = \left[ I_2 -\alpha \dot{\tilde{\mathcal {G}}}^0(p;0,0;\eta )\right] ^{-1} \tilde{\mathcal {G}}^0(p;x_2;\eta ) \end{aligned}$$
(16.91)

can be seen to cancel by re-expressing the “no lattice” Green’s functions \(\dot{\tilde{\mathcal {G}}}^0(p;0,0;\eta )\) and \(\tilde{\mathcal {G}}^0(p;x_2;\eta )\). For this purpose, by using (16.58) and (16.59), one can easily write \(\alpha \dot{\tilde{\mathcal {G}}}^0(p;0,0;\eta )\) matrix as

$$\begin{aligned} \alpha \dot{\tilde{\mathcal {G}}}^0(p;0,0;\eta ) = \frac{1}{\prod _{n=0}^\infty (\eta ^2-n) } \begin{pmatrix} a_{11} &{} a_{12} \\ a_{12} &{} a_{11} \end{pmatrix} \end{aligned}$$
(16.92)

where we have defined (j corresponds to the index r in (16.58, 16.59))

$$\begin{aligned} a_{11}(\eta )= \alpha _1 \eta \sum _{m=0}^\infty L_m \left[ \frac{\omega ^2_g l^2}{4\gamma ^2}\left( j\frac{d}{l} \right) ^2\right] \prod _{n=0 \atop n\ne m}^\infty (\eta ^2-n), \end{aligned}$$

and

$$\begin{aligned} a_{12}(\eta )= \alpha _2\sum _{m=1}^\infty L_{m-1}^1 \left[ \frac{\omega ^2_g l^2}{4\gamma ^2}\left( j\frac{d}{l} \right) ^2\right] \prod _{n=0 \atop n\ne m+1}^\infty (\eta ^2-n), \end{aligned}$$

with

$$\begin{aligned} \alpha _1= \frac{\alpha \omega _g}{4\pi \hbar \gamma ^2} \sum _{j=-\infty }^\infty e^{i pjd} e^{-\frac{\omega ^2_g l^2}{8\gamma ^2}\left( j\frac{d}{l} \right) ^2}, \end{aligned}$$

and

$$\begin{aligned} \alpha _2= \frac{i \alpha \omega _ g^2 l}{8\pi \hbar \gamma ^3} \sum _{j=-\infty }^\infty \left( j \frac{d}{l} \right) e^{i pjd} e^{-\frac{\omega ^2_g l^2}{8\gamma ^2}\left( j\frac{d}{l}\right) ^2}. \end{aligned}$$

Similarly, the \(\tilde{\mathcal {G}}^0(p;x_2;\eta )\) matrix can be written as

$$\begin{aligned} \tilde{\mathcal {G}}^0(p;x_2;\eta )= \frac{1}{\prod _{n=0}^\infty (\eta ^2-n) } \begin{pmatrix} b_{11} &{} b_{12} \\ b_{12} &{} b_{11} \end{pmatrix}, \end{aligned}$$
(16.93)

where we have defined

$$\begin{aligned} b_{11}(\eta )=\beta _1\eta \sum _{m=0}^\infty L_m \left[ \frac{\omega ^2_g l^2}{4\gamma ^2}\left( j\frac{d}{l} \right) ^2\right] \prod _{n=0\atop n\ne m}^\infty (\eta ^2-n), \end{aligned}$$

and

$$\begin{aligned} b_{12}(\eta )= \beta _2 \sum _{m=1}^\infty L_{m-1}^1 \left[ \frac{\omega ^2_g}{4\gamma ^2 l^2}\left( j\frac{d}{l} \right) ^2\right] \prod _{n=0\atop n\ne m+1}^\infty (\eta ^2-n) , \end{aligned}$$

with

$$\begin{aligned} \beta _1= \frac{ \omega _g}{4\pi \hbar \gamma ^2} \sum _{j=-\infty }^\infty e^{i pjd} e^{-\frac{\omega ^2_g l^2}{8\gamma ^2}\left( \frac{jd-x_2}{l} \right) ^2}, \end{aligned}$$

and

$$\begin{aligned} \beta _2=\frac{ i \omega _g^2 l}{8\pi \hbar \gamma ^3} \sum _{j=-\infty }^\infty \left( \frac{jd-x_2}{l} \right) e^{i pjd} e^{-\frac{\omega ^2_g l^2}{8\gamma ^2}\left( \frac{jd-x_2}{l}\right) ^2}. \end{aligned}$$

Note that in the above equations, n is a Landau index and m is a dummy index for the Landau levels; the maximum value of n and m is the same.

Finally substituting (16.92) and (16.93) in (16.91), the matrix \(T_1\) becomes

$$\begin{aligned} T_1 (\eta ) = \Bigg [ \begin{pmatrix} 1 &{} 0 \\ 0 &{} 1 \end{pmatrix} - \frac{1}{\prod _{n=0}^\infty (\eta ^2-n)} \begin{pmatrix} a_{11}(\eta ) &{} a_{12}(\eta ) \\ a_{12}(\eta ) &{} a_{11}(\eta ) \end{pmatrix} \Bigg ]^{-1} \frac{1}{\prod _{n=0}^\infty (\eta ^2-n)} \begin{pmatrix} b_{11}(\eta ) &{} b_{12}(\eta ) \\ b_{12}(\eta ) &{} b_{11}(\eta ) \end{pmatrix} \end{aligned}$$
$$\begin{aligned} = \left( \frac{1}{\prod _{n=0}^\infty (\eta ^2-n)}\right) ^{-1} \Bigg [ \begin{pmatrix} \prod _{n=0}^\infty (\eta ^2-n) &{} 0 \\ 0 &{} \prod _{n=0}^\infty (\eta ^2-n) \end{pmatrix} - \begin{pmatrix} a_{11}(\eta ) &{} a_{12}(\eta ) \\ a_{12}(\eta ) &{} a_{11}(\eta ) \end{pmatrix} \Bigg ]^{-1} \\ \times \frac{1}{\prod _{n=0}^\infty (\eta ^2-n)} \begin{pmatrix} b_{11}(\eta ) &{} b_{12}(\eta ) \\ b_{12}(\eta ) &{} b_{11}(\eta ) \end{pmatrix} \end{aligned}$$
$$\begin{aligned} = \Bigg [ \begin{pmatrix} \prod _{n=0}^\infty (\eta ^2-n) &{} 0 \\ 0 &{} \prod _{n=0}^\infty (\eta ^2-n) \end{pmatrix} - \begin{pmatrix} a_{11}(\eta ) &{} a_{12}(\eta ) \\ a_{12}(\eta ) &{} a_{11}(\eta ) \end{pmatrix} \Bigg ]^{-1} \begin{pmatrix} b_{11}(\eta ) &{} b_{12}(\eta ) \\ b_{12}(\eta ) &{} b_{11}(\eta ) \end{pmatrix}. \end{aligned}$$

In the above expression, the real “no lattice” poles (\(\eta \)=\(\pm \sqrt{n}\)) cancel. The above expression can be solved numerically. In this computation, it is also necessary to address the poles of the actual lattice Green’s function, which are defined by the zeros of the denominator, \(det\left[ I_2-\alpha \dot{\tilde{\mathcal {G}}}^0(p;0,0;\eta )\right] =0\), and these roots depend on p (and are not spaced by integer multiples of \(\omega _g\)): These poles are treated using the line broadening discussion of Sect. 16.5.2. The resultant \(2 \times 2\) matrix with \(c_{ij}(\eta )\) (where \(i,j=1,2\)) as matrix elements can be written as (substituting \(q=pd\))

$$\begin{aligned} Q (\eta ) = \int _{-\pi }^{\pi } dq e^{{- iqn^\prime }}T_1 (\eta ) = \begin{pmatrix} c_{11}(\eta ) &{} c_{12}(\eta ) \\ c_{21}(\eta ) &{} c_{22}(\eta ) \end{pmatrix}. \end{aligned}$$
(16.94)

Note that the above integral can be numerically calculated by applying the trapezoidal rule in the limits \(-\pi \) to \(\pi \), while kee** the trapezoidal step equal to \(\frac{\pi }{10}\); this gives an accuracy up to five decimal points for each value of \(n^\prime \).

Putting (16.94) in (16.90), the integral I becomes

$$\begin{aligned} I = \frac{\alpha \omega _g}{2\pi } \sum _{n^\prime =-\infty }^\infty \int _{-\infty }^\infty d\eta e^{- i \omega _g\eta t} \mathcal {G}^0(x_1,n^\prime d;\eta ) . Q(\eta ). \end{aligned}$$
(16.95)

It can be seen that the “no lattice” poles of the second integral (p integral) in (16.90) have been removed; and the poles of the actual lattice Green’s function have been dealt with using the line broadening of Sect. 16.5.2, so the only remaining poles that we have to deal with now are the poles of the matrix \(\mathcal {G}^0(x_1,n^\prime d;\eta )\).

Using (16.67) and (16.68), the matrix \( \alpha \mathcal {G}^0(x_1,n^\prime d;\eta )\) can be written as

$$\begin{aligned} \alpha \mathcal {G}^0(x_1,n^\prime d;\eta )= \begin{pmatrix} \gamma _{1}\frac{1}{\eta } &{} 0 \\ 0 &{} \gamma _{1} \frac{1}{\eta } \end{pmatrix} + \sum _{n=1}^\infty \begin{pmatrix} \gamma _{1} \eta \frac{ L_n \left[ \Upsilon \right] }{\eta ^2 -n} &{} \gamma _{2} \frac{ L_{n-1}^1 \left[ \Upsilon \right] }{\eta ^2 -n} \\ \gamma _{2} \frac{ L_{n-1}^1 \left[ \Upsilon \right] }{\eta ^2-n} &{} \gamma _{1} \eta \frac{ L_n \left[ \Upsilon \right] }{\eta ^2 -n} \end{pmatrix}, \end{aligned}$$
(16.96)

where \(\Upsilon \), \(\gamma _1\) and \(\gamma _2\) are defined in (16.74) and (16.75). Using the matrices \(Q(\eta )\) and \( \alpha \mathcal {G}^0(x_1,n^\prime d;\eta )\) in (16.95), and breaking the matrix into two matrices, we get

$$\begin{aligned} I = \frac{ \omega _g}{2\pi } \sum _{n^\prime =-\infty }^\infty \int _{-\infty }^\infty d\eta e^{- i \omega _g\eta t} \begin{pmatrix} \gamma _{1}\frac{1}{\eta } c_{11}(\eta ) &{} \gamma _{1}\frac{1}{\eta } c_{12}(\eta ) \\ \gamma _{1}\frac{1}{\eta } c_{21}(\eta ) &{} \gamma _{1}\frac{1}{\eta } c_{22}(\eta ) \nonumber \end{pmatrix} \\ + \frac{ \omega _g}{2\pi } \sum _{n^\prime =-\infty }^\infty \int _{-\infty }^\infty d\eta e^{- i \omega _g\eta t} \begin{pmatrix} M_{11}(\eta ) &{} M_{12}(\eta ) \\ M_{21}(\eta ) &{} M_{22}(\eta ) \end{pmatrix}. \end{aligned}$$
(16.97)

(The first and second matrices correspond to n=0 and \(n>0\) Landau minibands, respectively.) In the above expression

$$\begin{aligned} M_{11 \atop 22}(\eta ) = \sum _{n=1}^\infty \Bigg ( \gamma _{1} \eta \frac{ L_n \left[ \Upsilon \right] }{\eta ^2 -n} c_{11 \atop 22 }(\eta ) + \gamma _{2} \frac{ L_{n-1}^1 \left( \Upsilon \right) }{\eta ^2 -n} c_{21 \atop 12}(\eta )\Bigg ) , \end{aligned}$$

and

$$\begin{aligned} M_{12 \atop 21}(\eta ) = \sum _{n=1}^\infty \Bigg ( \gamma _{1} \eta \frac{ L_n \left[ \Upsilon \right] }{\eta ^2 -n} c_{12 \atop 21}(\eta ) + \gamma _{2} \frac{ L_{n-1}^1 \left[ \Upsilon \right] }{\eta ^2 -n} c_{22 \atop 11}(\eta ) \Bigg ). \end{aligned}$$

In (16.97), the first matrix has a pole at \(\eta \)=0, while the second matrix has poles at \(\eta \)=\(\pm \sqrt{n}\). We now use contour integration with the Jordan lemma (closing the contour in the lower half plane for \(t>0\)) to evaluate the integrals. Results for the two terms in (16.97) are

$$\begin{aligned} \int _{-\infty }^\infty d\eta e^{- i \omega _g\eta t} \frac{c_{ij}(\eta )}{\eta } =- i\pi \eta _+(t) \left[ c_{ij}(\eta )\right] _{\eta =0} , \end{aligned}$$
(16.98)
$$\begin{aligned} \int _{-\infty }^\infty d\eta e^{- i \omega _g\eta t} M_{11 \atop 22 }(\eta )&= -i \pi \eta _+(t) \sum _{n=1}^\infty \cos \left( \frac{\omega _g l}{\gamma } \frac{t}{\tau _o} \sqrt{n} \right) \nonumber \\&\times \left[ \gamma _1 L_n \left[ \Upsilon \right] c_{11 \atop 22}(\eta ) + \frac{\gamma _2}{\sqrt{n}} L_{n-1}^1 \left[ \Upsilon \right] c_{21 \atop 12} (\eta ) \right] _{\eta =\sqrt{n}} \end{aligned}$$
(16.99)

and

$$\begin{aligned} \int _{-\infty }^\infty d\eta e^{- i \omega _g\eta t} M_{12 \atop 21}(\eta )&=- \pi \eta _+(t) \sum _{n=1}^\infty \sin \left( \frac{\omega _g l}{\gamma } \frac{t}{\tau _o} \sqrt{n} \right) \nonumber \\&\times \left[ \gamma _1 L_n \left[ \Upsilon \right] c_{12 \atop 21}(\eta ) + \frac{\gamma _2}{\sqrt{n}} L_{n-1}^1 \left[ \Upsilon \right] c_{22 \atop 11} (\eta ) \right] _{\eta =\sqrt{n}}. \end{aligned}$$
(16.100)

(\(\eta _+(t)\) is the Heaviside unit step function and ij=1, 2). In the calculation of the expressions given by (16.98), (16.99) and (16.100), we have also used

$$\begin{aligned} \left[ c_{11 \atop 22}(\eta )\right] _{\eta =-\sqrt{n}} = \left[ c_{11 \atop 22}(\eta )\right] _{\eta =\sqrt{n}}, \end{aligned}$$

and

$$\begin{aligned} \left[ c_{12 \atop 21}(\eta )\right] _{\eta =-\sqrt{n}} = - \left[ c_{12 \atop 21}(\eta )\right] _{\eta =\sqrt{n}}, \end{aligned}$$

which we found during the calculations [21].

Hence, (16.98) along with (16.99) and (16.100) provide the complete solution for the integral I, which is the time representation of the second term of the full Green’s function \(\mathcal {G}(x_1,x_2; t)_{K}\) given in (16.69).

Rights and permissions

Reprints and permissions

Copyright information

© 2022 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Horing, N.J.M., Ayyubi, R.A.W., Sabeeh, K., Bahrami, S. (2022). Quantum Dynamics in a 1D Dot/Antidot Lattice: Landau Minibands and Graphene Wave Packet Motion in a Magnetic Field. In: Ünlü, H., Horing, N.J.M. (eds) Progress in Nanoscale and Low-Dimensional Materials and Devices. Topics in Applied Physics, vol 144. Springer, Cham. https://doi.org/10.1007/978-3-030-93460-6_16

Download citation

Publish with us

Policies and ethics

Navigation