puntos cuanticos propiedades electronicas

25
 F ront. Phys., 2012, 7(3): 328–352 DOI 10.1007/s11467-011-0200-5 REVIEW ARTICLE Electronic and optical properties of semiconductor and graphene quantum dots Wei-dong Sheng 1,2,,  Marek Korkusinski 1 ,  Alev Devrim G¨ cl¨ u 1 ,  Michal Zielinski 1,3 , Pawel Potasz 1,4 ,  Eugene S. Kadantsev 1 ,  Oleksandr Voznyy 1 ,  Pawel Hawrylak 1,1 Institute for Micros tructur al Sciences, National Resea rch Council of Canada, Ottawa, Canada 2 Department of Physics, Fudan University, Shanghai 200433, China 3 Institute of Physics, Nicolaus Copernicus University, Torun, Poland 4 Institute of Physics, Wro claw University of Te chnolo gy, Wro claw, Poland E-mail:  [email protected],  [email protected] Received April 21, 2011; accepted July 27, 2011 Our recent work on the electronic and optical properties of semiconductor and graphene quan- tum dots is reviewed. For strained self-assembled InAs quantum dots on GaAs or InP substrate atomic positions and strain distribution are described using valence-force eld approach and con- tinuous elasticity theory. The strain is coupled with the eective mass,  k ·  p, eective bond-orbital and atomistic tight-binding models for the description of the conduction and valence band states. The single-particle states are used as input to the calculation of optical properties, with electron- electron interactions included via conguration interaction (CI) method. This methodology is used to describe multiexciton complexes in quantum dot lasers, and in particular the hidden symmetry as the underlying principle of multiexciton energy levels, manipulating emission from biexcitons for entangled photon pairs, and optical control and detection of electron spins using gates. The self-assembled quantum dots are compared with graphene quantum dots, one carbon atom-thick nanostructures. It is shown that the control of size, shape and character of the edge of graphene dots allows to manipul ate simultaneously the elect ronic, optical, and magnetic properties in a single material system. Keywords  quantum dots, electronic structure, multiexc iton, graphene, magnetism PACS numbers  78.67.Hc, 73.21.La, 73.63.Kv , 73.22.Pr Contents 1 Introduction 329 2 Self-assembled quantum dots 330 2.1 Strain distribution 330 2.2 Electronic structure 331 2 .2 .1 Ee ct ive -m as s a p pr oxi m at io n 33 1 2 .2 .2 Parabolic conne me n t m od el 3 32 2.2.3 Eight-band k ·  p approach 332 2.2.4 Eect ive bo nd-or bital model 333 2.2.5 Empirical tight-bindi ng method 334 2.3 Optical properties 336 2.3.1 Photoluminesce nce: Polar ization and anisotropy 336 2.3.2 Electro n–elec tron interactions and multiexciton complexes 337 2.3.3 Hidden symmetry 338 2.3.4 Fine struct ure: Electro n–hole exchange interaction 338 2.4 Quantum dots in magnetic elds 338 2.4.1 Multi exciton F ock–Da rwin spectrum 339 2.4 .2 Ele ctron g fac tors : Dis tri bution and anisotropy 339 2. 4.3 Ho le g  factors: Envelope orbital momentum 340 2.5 Quantum dots in electric elds 341 2. 5.1 Quan tum-conned Star k e ec t 341 2.5 .2 Electrical tu ning o f exc iton g  fact ors 341 2.6 Single InAs/In P se lf-asse mble d q uantu m dots on nanotemplates 342 3 Graphene quantum dots 343 3.1 Introduction 343 c  Higher Education Press and Springer-Verlag Berlin Heidelberg 2012

Upload: carlos-eduardo

Post on 05-Oct-2015

14 views

Category:

Documents


0 download

DESCRIPTION

esto me da las p´ropiedades electroncias de los qd

TRANSCRIPT

  • Front. Phys., 2012, 7(3): 328352

    DOI 10.1007/s11467-011-0200-5

    REVIEW ARTICLE

    Electronic and optical properties of semiconductor and

    graphene quantum dots

    Wei-dong Sheng1,2,, Marek Korkusinski1, Alev Devrim Guclu1, Michal Zielinski1,3,

    Pawel Potasz1,4, Eugene S. Kadantsev1, Oleksandr Voznyy1, Pawel Hawrylak1,

    1 Institute for Microstructural Sciences, National Research Council of Canada, Ottawa, Canada

    2 Department of Physics, Fudan University, Shanghai 200433, China

    3 Institute of Physics, Nicolaus Copernicus University, Torun, Poland

    4 Institute of Physics, Wroclaw University of Technology, Wroclaw, Poland

    E-mail: [email protected], [email protected]

    Received April 21, 2011; accepted July 27, 2011

    Our recent work on the electronic and optical properties of semiconductor and graphene quan-tum dots is reviewed. For strained self-assembled InAs quantum dots on GaAs or InP substrateatomic positions and strain distribution are described using valence-force eld approach and con-tinuous elasticity theory. The strain is coupled with the eective mass, k p, eective bond-orbitaland atomistic tight-binding models for the description of the conduction and valence band states.The single-particle states are used as input to the calculation of optical properties, with electron-electron interactions included via conguration interaction (CI) method. This methodology is usedto describe multiexciton complexes in quantum dot lasers, and in particular the hidden symmetryas the underlying principle of multiexciton energy levels, manipulating emission from biexcitonsfor entangled photon pairs, and optical control and detection of electron spins using gates. Theself-assembled quantum dots are compared with graphene quantum dots, one carbon atom-thicknanostructures. It is shown that the control of size, shape and character of the edge of graphenedots allows to manipulate simultaneously the electronic, optical, and magnetic properties in a singlematerial system.

    Keywords quantum dots, electronic structure, multiexciton, graphene, magnetism

    PACS numbers 78.67.Hc, 73.21.La, 73.63.Kv, 73.22.Pr

    Contents

    1 Introduction 3292 Self-assembled quantum dots 330

    2.1 Strain distribution 3302.2 Electronic structure 331

    2.2.1 Eective-mass approximation 3312.2.2 Parabolic connement model 3322.2.3 Eight-band k p approach 3322.2.4 Eective bond-orbital model 3332.2.5 Empirical tight-binding method 334

    2.3 Optical properties 3362.3.1 Photoluminescence: Polarization

    and anisotropy 3362.3.2 Electronelectron interactions and

    multiexciton complexes 337

    2.3.3 Hidden symmetry 3382.3.4 Fine structure: Electronhole

    exchange interaction 3382.4 Quantum dots in magnetic elds 338

    2.4.1 Multiexciton FockDarwin spectrum 3392.4.2 Electron g factors: Distribution and

    anisotropy 3392.4.3 Hole g factors: Envelope orbital

    momentum 3402.5 Quantum dots in electric elds 341

    2.5.1 Quantum-conned Stark eect 3412.5.2 Electrical tuning of exciton g factors 341

    2.6 Single InAs/InP self-assembled quantumdots on nanotemplates 342

    3 Graphene quantum dots 3433.1 Introduction 343

    c Higher Education Press and Springer-Verlag Berlin Heidelberg 2012

  • Wei-dong Sheng, et al., Front. Phys., 2012, 7(3) 329

    3.2 Electronic structure Tight-bindingapproach 344

    3.3 Dirac fermions 3443.4 Graphene quantum dots 3453.5 Shape and edge eects 3453.6 Gated quantum dots: Beyond tight-binding

    approach 3463.7 Magnetism in triangular quantum dots 3473.8 Excitons in triangular quantum dots 3483.9 Eect of imperfections 349

    4 Conclusions 349Acknowledgements 349References 349

    1 Introduction

    In this article we highlight some of our recent works to-wards the understanding of electronic and optical prop-erties of semiconductor [15] and graphene quantum dots[6]. The recent research on semiconductor quantum dotsfollows naturally the evolution of semiconductor tech-nology from transistors and lasers based on bulk sili-con and bulk gallium arsenide to eld eect transistorsand quantum-well lasers. In these systems, the control ofmaterial composition in one dimension, e.g., molecularbeam epitaxy has led to integrated circuits and revolu-tionary changes in information technology. Semiconduc-tor quantum dots are a natural step forward in allowingfor the control of material composition in three dimen-sions and at the nanoscale. Hence quantum dots are anexample of nanoscience and nanotechnology in semicon-ductors. There are four major classes of quantum dots:(i) lateral gated quantum dots, (ii) self-assembled quan-tum dots, (iii) colloidal nanocrystals, and the most recentaddition, (iv) graphene quantum dots.

    The lateral gated quantum dots are created at a semi-conductor GaAlAs/GaAs heterojunction containing atwo-dimensional electron gas as in the eld-eect tran-sistor (FET). On top of the GaAs surface a pattern ofmetallic gates with nanometer dimensions is deposited.When negative voltage is applied to the gates, electronsresiding underneath these gates at the GaAlAs/GaAsheterojunction feel repulsive potential and are pushedout from under the gates [1, 3]. By designing the gatesin such a way that the repulsive potential under the gatesseen by electrons resembles a volcano, a controlled num-ber of electrons, down to one, can be trapped in thevolcanos crater. The counting of electrons in the lateralquantum dot and subsequent isolation of a single electronhave been demonstrated at the Institute for Microstruc-tural Sciences [79]. It is now possible to construct dou-ble [1013] and triple quantum dot molecules [1417]where individual electrons are isolated, quantum me-chanically coupled, and manipulated in real time. Some

    of the quantum information aspects of lateral quantumdot molecules can be found in our recent review [4]. Withnanocrystals being a very active and well-covered eld, inthis review we will focus on self-assembled and graphenequantum dots.

    The appearance of self-assembled quantum dots dur-ing the growth of InAs layers on GaAs in molecular beamepitaxy was noted as early as 1985 by Marzin and co-workers [18] and much of the pioneering work has beencarried out by Petro and co-workers [19] in the early90s. From early theoretical and experimental researchself-assembled quantum dots are nding applications inquantum dot lasers and ampliers [2025] and solar cells[26]. Current research, some theoretical aspects of whichare reviewed in this paper, is focused on single quantumdots and quantum dot molecules with potential appli-cations as single photon sources, sources of entangledphoton pairs, and when charged, quantum bits. From atheoretical point of view these structures are challengingas they are neither few-atom molecules nor solids andinvolve collective behavior of millions of atoms. We willdescribe our attempts at providing an understanding ofelectronic and optical properties of million-atom nanos-tructures at dierent levels of sophistication. In partic-ular, we will discuss to what extent the single-particlestates in quantum dots can be viewed as states of twoquantum harmonic oscillators. When the dots are lledwith electrons, the generalized Hunds rule allows us topredict the spin of the ground state [27]. When the dotsare populated with electrons and holes, as in a quan-tum dot laser, the hidden symmetry replaces the Hundsrule as an underlying principle governing the propertiesof multiexciton complexes [2832]. The single exciton,controlling the absorption of photons by a quantum-dot-based solar cell, can be understood in terms of mixing ofbright and dark congurations by Coulomb interactions[33]. Manipulating electronic and optical properties ofsingle self-assembled quantum dots has become possiblewith the growth of InAs quantum dots on InP templates,pioneered by Williams and co-workers [34]. This enabledthe gating of individual dots [35] and embedding of thesedots in a photonic cavity [36]. These quantum dots areparticularly interesting because they emit at the telecomwavelength. We will describe some of the properties ofInAs/InP quantum dots using the eective bond-orbitalmodel.

    Finally, the recent isolation of a single, atomicallythick carbon graphene layer [37] opened a new eld ofresearch. Since graphene does not have a gap, size quan-tization opens an energy gap and turns graphene into asemiconductor. Unlike in semiconductor quantum dots,the gap in graphene quantum dots can be tuned fromzero to perhaps even the gap of the benzene ring. How-ever, graphene dots need to be terminated and the edgesplay a very important role, with zigzag edges leading

  • 330 Wei-dong Sheng, et al., Front. Phys., 2012, 7(3)

    to energy shells in the middle of the gap and a nitemagnetic moment. The magnetic moment can in turn berelated to another property of graphene the sublatticesymmetry. The research on graphene quantum dots is ata very early stage [38, 39] and we hope that this shortreview will stimulate its rapid progress.

    The research eld covering semiconductor andgraphene quantum dots has increased tremendously sinceearly 1990s with contributions from many outstandingscientists. It is not possible to cover the entire eld andgive credit to everyone in such a short article. We focushere primarily on the work carried out at the NRC In-stitute for Microstructural Sciences and refer the readerto existing reviews for additional coverage of the eld.

    2 Self-assembled quantum dots

    Self-assembled quantum dots consist of a low-bandgapsemiconductor A, typically InAs, embedded in a higher-bandgap semiconductor B, typically GaAs or InP [13]. The two materials have similar symmetry but dier-ent lattice constants. The dots are formed during theStranskyKrastanow process of, e.g., molecular beamepitaxy of InAs on GaAs. In this process, the strainbuilding up in the InAs layer is relieved by the formationof quasi-two-dimensional islands. The islands, typicallypyramidal or lens-shaped, are capped with GaAs. An ex-ample of a lens-shaped InAs quantum dot on a wettinglayer is shown in Fig. 1. When electrons and/or holes areinjected into the sample, they become conned in InAsquantum dots. The remainder of this section describesthe electronic properties of self-assembled quantum dots.Since the two materials are strained, we start with theeect of strain, followed by a description of electronicproperties of these systems.

    Fig. 1 Indium (red) and Arsenic (blue) atoms in a lens-shapedInAs quantum dot with diameter of 25 nm and height of 3.5 nmon a 0.6 nm high wetting layer. The GaAs barrier material atomsare not shown.

    2.1 Strain distribution

    As aforementioned, self-assembled quantum dots areformed to relax the strain due to the lattice mismatchbetween two materials grown on top of each other, likeInAs/GaAs, InAs/InP or CdTe/ZnTe. The strain dis-tribution in the vicinity of a quantum dot can be de-termined by either the continuum elasticity theory orthe atomistic valence-force-eld approach [40, 41]. The

    domain of strain calculation is typically a rectangularcomputational box which generally includes the entirestructure, with characteristic sizes on the micrometerscale. Depending on the device conguration, xed orfree-standing boundary conditions are implemented [42].

    In the framework of the continuum elasticity theory,the strain tensor is dened for each unit cell of thestructure as

    jk =12

    (ujrk

    +ukrj

    )(1)

    where rj and uk are the components of the position vec-tor r and displacement vector u, respectively. It can beobtained by minimizing the following elastic energy func-tional:

    E =12

    [C11

    (2xx +

    2yy +

    2zz

    )+C44

    (2xy +

    2yz +

    2zx

    )+2C12

    (xxyy + yyzz + zzxx

    )]d3r (2)

    where C11, C44, and C12 are position-dependent elasticconstants, assuming the value of the quantum dot or thebarrier matrix materials, respectively.

    In the atomistic approach, the elastic energy of eachatom i is a function of the positions of its nearest neigh-bors,

    Ei =12

    4j=1

    ijd2ij

    (R2ij d2ij)2

    +3

    j=1

    4k=j+1

    ijik

    dijdik

    (Rij Rik + 13dijdik

    )2(3)

    with dij being the bond length, and ij and ij the forceconstants between atoms i and j. By minimizing thetotal elastic energy which is a sum over all atoms, theequilibrium positions of all the atoms are computed andthe corresponding strain tensor can therefore be obtained[41].

    The minimization procedure can proceed either viathe standard conjugate gradient method, or by a moreintuitive force-eld approach. As the strain energy is afunction of the positions of all the unit cells or atoms,i.e., E = E(r1, r2, ), the force exerted on each unit istherefore given by Fi = E/ri. In one iteration, eachunit is displaced proportionally to Fi. The iterative pro-cedure ends when the amplitudes of all the forces becomesuciently small. In zinc-blende semiconductors such asInAs or GaAs, shear strain would induce piezoelectriccharge along the interface between neighboring unit cells[42], as given by

    P (r) = 2e14[yz(r)

    x+

    xz(r)y

    +xy(r)

    x

    ](4)

    The piezoelectric potential can be obtained by solvingthe corresponding Poisson equation.

  • Wei-dong Sheng, et al., Front. Phys., 2012, 7(3) 331

    In Table 1 we list the elastic constants (in unit of 1011

    dyne/cm1), piezoelectric modulus (Cm2), ideal bondlengths (A) and force constants (103 dyne) of InAs, GaAsand InP. Figure 2 shows the shear strain component xyand piezoelectric potential calculated for a lens-shapedInAs/GaAs self-assembled quantum dot with a base di-ameter of 19.8 nm and a height of 2.8 nm placed on atwo-monolayer wetting layer (ML). Red/blue area corre-sponds to positive/negative part of the strain and poten-tial.

    Table 1 List of material parameters used in the strain calcula-tion by the continuum or atomistic elasticity theory.

    InAs GaAs InP

    C11 8.329 11.879 10.22

    C12 4.526 5.376 5.76

    C44 3.96 5.94 4.6

    e14 0.045 0.16 0.035

    d 2.622 2.448 2.537

    35.18 41.19 43.04

    5.50 8.95 6.24

    Fig. 2 Isosurface plots of the shear strain component xy andpiezoelectric potential in a lens-shaped InAs/GaAs self-assembledquantum dot.

    2.2 Electronic structure

    Within the framework of continuum elasticity, the eectof strain on the electronic structure of a self-assembledquantum dot is described by the BirPikus deformationpotential theory [43]. There are two major componentsof the strain which modies the band-edge energies, thehydrostatic (Hs) and the biaxial (Bs), as dened in thefollowing:

    Hs = xx + yy + zz

    B2s = (xx yy)2 + (yy zz)2 + (zz xx)2 (5)The band-edge energies of the conduction bands aremainly aected by the hydrostatic component of thestrain while the heavy- and light-hole bands are furtherinuenced by the biaxial strain components, i.e.,

    Uc = acHs + Ecbo

    Uhh = avHs bvBs + EvboUlh = avHs + bvBs + Evbo (6)

    where Ecbo and Evbo are the band osets between the dot

    and matrix materials before strain, and ac, av, and bvare the deformation potential parameters [40].

    Figure 3 plots the conning potentials for electrons,heavy- and light-holes along the growth direction (upperpanel) and along the diameter of the bottom base of thedot. The height of the dot is chosen to be the same asthat in Fig. 2 while the diameter is twice as large for bet-ter visualization. It is seen that the eective band-gap inthe presence of the strain is increased to about 695 meVfrom 475 meV in the strain free bulk InAs. The strain isalso seen to enhance the depth of the conning potentialfor the heavy-holes to about 300 meV from 85 meV. Theheavy- and light-hole bands which are degenerate in thebulk are now lifted by the biaxial strain component. Thesplitting in the middle of the dot is about 177 meV.

    Fig. 3 Conning potentials for electrons, heavy- and light-holes(from top to bottom) deformed by the strain InAs/GaAs self-assembled quantum dot.

    Besides the eect on the band-edge energies, the strainalso modies the eective masses of electrons and holesin the dot.

    2.2.1 Eective-mass approximation

    The states conned in the quantum dot are often com-posed of components from both the conduction and va-lence bands. With the biaxial strain present in the self-assembled quantum dots, the degeneracy between theheavy and light holes is lifted and the low-lying statesin the valence band have mostly a heavy hole character.

  • 332 Wei-dong Sheng, et al., Front. Phys., 2012, 7(3)

    If the eect of band mixing between the conduction andvalence bands is neglected, the problem could be reducedto two separate single-band eective-mass equations, onefor electron and the other for hole, as given by [44]

    He = 2

    2m//e

    (2

    x2+

    2

    y2

    )

    2

    2me

    2

    z2+ Uc + Vp

    Hh =2

    2m//h

    (2

    x2+

    2

    y2

    )+

    2

    2mh

    2

    z2+ UhhVp (7)

    where Vp is the piezoelectric potential. Although the elec-tron eective mass is almost isotropic in most IIIV semi-conductors, two independent components, m//e and me ,are used to reect the anisotropic geometries of the quan-tum dots, and so are the analogous masses for holes.

    Since there is no a priori rule on how to choose theeective masses within the model itself, m//e , me , m

    //h ,

    and mh are treated as adjustable parameters which aredetermined by tting the calculated energy spectrum tothat obtained by a more sophisticated approach like themultiband k p or the empirical tight-binding methoddescribed below. A way to avoid the tting procedureis to regard the eective masses as position-dependentvariables [45].

    It is found that the electron eective mass in quantumdots is generally larger than the bulk value and becomesanisotropic in the dots of large aspect ratio between thevertical and lateral dimensions. Unlike the bulk mate-rial, the hole eective mass is seen to be almost isotropicin the dots of small aspect ratio. For an example of atInAs/GaAs quantum dots, the most appropriate valuefor the electron and hole eective mass is believed to bethe electron eective mass in bulk GaAs (0.067 m0) andthe vertical heavy-hole eective mass in bulk InAs (0.34m0), respectively.

    2.2.2 Parabolic connement model

    The eective mass model has been applied to lens-shapedquantum dots [46]. It was shown that, in the adiabaticapproximation, the radial conning potentials for elec-trons and heavy holes can be well approximated byparabolas with a given depth and radius of the dot.The single-electron Hamiltonian is reduced to a Hamilto-nian of two bosons, two 1D harmonic oscillators (HOs)with creation (annihilation) operators a+m(am) and en-ergy level spacing e [7],

    Hho = e

    (m +

    12

    )a+mam + e

    (n +

    12

    )a+n an (8)

    The corresponding wave functions are those of thetwo-dimensional harmonic oscillator. A similar modelis applied to heavy holes, and the analytical solutionfor HO states in a perpendicular magnetic eld can befound, e.g., in Ref. [7]. It was recently demonstrated that

    InAs/InP and disk-like InAs/GaAs dots grown using theindium-ush technique developed by Wasilewski and co-workers [47] can be well described by the HO model.If the HO states are populated by photoexcited elec-trons and holes with increasing excitation power, as il-lustrated in Fig. 4(c), the resulting emission spectrumconsists of several peaks corresponding to electron andhole shells. The evolution of the emission spectrum withan increasing magnetic eld is shown in Figs. 4(a) and(b). The spectrum strikingly resembles that of the 2DHO in a magnetic eld. As already mentioned, the HOmodel works quite well, but the parameters of the model,masses, and connement energies e and h are ttingparameters that must be obtained by other methods likethe multiband k p or empirical tight-binding methoddescribed below.

    Fig. 4 Emission spectrum of an ensemble of InAs/GaAs self-assembled quantum dots as a function of applied magnetic eldshows harmonic oscillator states. Reproduced from Ref. [48], Copy-right c 2004 American Physical Society.

    2.2.3 Eight-band k p approach

    The eight-band k p method uses eight Bloch functionsat the point of the Brillouin zone as the basis func-tions to describe electron states with a nite wave vec-tor. As the lateral size of quantum dots is usually muchlarger than the lattice constant, the k p method hasbeen widely used in the calculation of conned electronstates. In general, the multiband k p Hamiltonian canbe written as

    Hkp = Ebo +Axkxkx +Aykyky +Azkz kz+Bxykxky +Byzkykz +Bxzkxkz+Cxkx +Cyky +Czkz + VP (9)

  • Wei-dong Sheng, et al., Front. Phys., 2012, 7(3) 333

    where Ebo is the matrix for the band osets, and A,B, and C are the matrices of coecients [49]. Withinthe deformation potential theory, an additional part Hs,which has a similar structure as Hkp, is added to takeinto account the eects due to strain [50]. The computa-tional box for the calculation of the electronic structureof quantum dots is not necessarily as large as that usedfor the strain since the conned states are mostly local-ized inside the quantum dots. The Hamiltonian is rstdiscretized on a reduced three-dimensional mesh, whichresults in a large, sparse matrix, and then is diagonalizedby utilizing the Lanczos algorithm [51].

    The wave function of a single-particle state, , is thesum of products of envelope functions k(r) and basisfunctions uk(r), i.e., (r) =

    n n(r)un(r). The basis

    functions are usually taken as |s, |x, |y, and |z, cou-pled with two eigenspinors | and | . These functionsare referred to as the uncoupled spinorbital basis. Be-cause of the spinorbit interaction, only the total angularmomentum is a good quantum number. Therefore, theeigenfunctions of the total angular momentum operatorare considered to be a more convenient choice of what isreferred to as the coupled spinorbital basis. These ba-sis functions are closer to the band-edge Bloch functionsthan the uncoupled set. Note that the coecient matri-ces in the k p Hamiltonian take dierent forms in theuncoupled and coupled spinorbital basis [52].

    Figure 5 shows the density of states for a lens-shapedInAs/GaAs self-assembled quantum dot grown on a 2ML wetting layer with a base diameter of 25.4 nm anda height of 2.8 nm. The conned states in the conduc-tion bands are seen to form almost equally spaced clus-ters, with increasing number of states in each one, inanalogy to the HO model. The density of states in thevalence band is found to increase steadily as the energyapproaches the continuum. Figure 5 also shows the com-position of the states, i.e., the proportion of each kind ofcomponents. It is seen that the states in the conductionband are dominated by the components from the sameband, however with decreasing proportion as the energyincreases. In the valence band, the low-lying states arefound to be dominated by the heavy-hole components,while the proportion of the light-hole components is non-negligible in high-lying states. Close to the continuum,the proportion of the heavy- and light-hole componentsis seen to be saturated at around 60% and 25%, respec-tively.

    Since the eight-band k p method keeps a good bal-ance between handling complicated band mixing ef-fects and computational complexity, it has becomewidely used in the calculation of the electronic struc-ture of self-assembled quantum dots, and successfullyexplained many interesting phenomena, such as invertedelectron-hole alignment in intermixed single quantumdots [53] and spontaneous localization of hole states in

    Fig. 5 Density of states (lower panel) and composition ofstates (upper panel) calculated for a lens-shaped InAs/GaAs self-assembled quantum dot. Small solid dots (blue) are for the compo-nents from the conduction bands, larger ones (red) for the heavy-hole components, and open dots (green) for the light-hole compo-nents.

    quantum-dot molecules [54]. The k p method predictedthe transition between bonding and anti-bonding statesas ground states of a valence hole in a vertical quantumdot molecule as a function of the separation between theconstituent quantum dots [55, 56]. This unusual behav-ior related to strong spinorbit coupling was recently ob-served experimentally by Doty et al. [57].

    2.2.4 Eective bond-orbital model

    The multiband k p method accounts for the properstructure of the valence band, including heavy, light, andspin split-o hole subbands. It is, however, limited tothe vicinity of the point, and therefore is expected tobreak down as the size of the nanostructure decreases.The eective bond-orbital model (EBOM) [58] is an em-pirical sp3 tight-binding method in which indium-arsenicdimers are replaced by an eective atom. Hence the fullsymmetry of the zinc-blende lattice is reduced to that ofa fcc lattice. Although EBOM misses the lack of inversionsymmetry of zinc-blende structures and the microscopicatomic structure of the unit cell, it can, however, repro-duce the eective masses of electrons and holes at the point, as well as conduction and valence band edgesat both the and X points with the second nearest-neighbor interactions included [59]. The tight-bindingHamiltonian matrix describing hopping between eectives and p orbitals of eective atoms on sites R is given by

    H(Rs,Rs) = E000ss R,R + E110ss R, + E

    200ss R,

    H(Rp,Rp) = R,[E110xx

    2p + E

    011xx (1 2p )

    ]+R,

    [E200xx

    2p + E

    002xx (1 2p)

    ]

  • 334 Wei-dong Sheng, et al., Front. Phys., 2012, 7(3)

    +E000xx R,R

    H(Rs,Rp) = E110sx pR,

    H(Rp,Rp) = E110xy ppR, (10)

    with R = RR, and =

    a

    2[(1,1, 0), (1, 0,1), (0,1,1)]

    = a[(1, 0, 0), (0,1, 0), (0, 0,1)] (11)

    being the positions of the nearest and next-nearest neigh-bors, respectively. The eect of strain is incorporatedthrough the deformation potential theory and throughthe piezoelectric eect [60]. Table 2 shows the next-nearest-neighbor parameters for GaAs and InAs.

    Table 2 Next-nearest-neighbor parameters (in eV).

    GaAs InAs

    E000ss 3.581011 2.698585

    E110ss 0.039499 0.124187E200ss 0.264835 0.118389E110sx 0.387077 0.368502

    E000xx 3.301012 3.068847E110xx 0.424999 0.412500

    E011xx 0.070000 0.112499E200xx 0.186297 0.156340E002xx 0.138401 0.154132

    E110xy 0.495000 0.525000

    The matrix elements for spin-orbit interaction aregiven by H(Rx,Ry) = i/3, H(Rx,Ry) = i/3,H(Rx,Rz) = /3, H(Ry,Rz) = i/3, and theirconjugate terms, with being the spin-orbit splitting inthe valence band. When the spin-orbit interaction is notvery strong, it is possible to separate the components ofan envelope function into two groups. One group consistsof components for spin up basis functions, |s, |x, |y,and |z, and the other one consists of components forspin down basis functions. A pseudospin can be assignedto a state if its polarization, dened as

    p =|s(r)|2 + |x(r)|2 + |y(r)|2

    +|z(r)|2 dr (12)is either p 1 (a spin up state) or p 0 (a spindown state). The assignment of pseudospin can be car-ried out by introducing a small magnetic eld ( 1 mT)along the growth direction of the quantum dots to liftthe degeneracy induced by the time-reversal symmetry.

    2.2.5 Empirical tight-binding method

    In the empirical tight-binding method we rst expandthe wave function in the basis of atomic orbitals

    =R,

    cR|R (13)

    and next form the Hamiltonian matrix in this atomic ba-sis [61]. The matrix elements are treated as parametersdetermined by comparison with ab-initio and/or experi-mental results for bulk materials. Signicant amount ofwork has been devoted to ab-initio determination of va-lence and conduction bands, eective masses and bandosets for strained materials of quantum dot and barriermaterials [62, 63].

    By comparison, in the pseudopotential approach [64,67], the self-consistent potential seen by an electron ina bulk material is replaced by a sum of eective atomicpotentials. These atomic potentials are next used to gen-erate the one-electron potential of the nanostructure.

    In our tight-binding approach [61] the wave functionon each atom is described by ten valence orbitals foreach spin: one of type s, three of type p, ve of type d,and an additional s orbital accounting for higher-lyingstates. Each orbital is doubly spin-degenerate, thus re-sulting in a total of 20 bands. The resulting Hamiltonianof a quasiparticle in an N -atom quantum dot, written inthe language of second quantization, is

    HTB =N

    i=1

    20=1

    ic+ici +

    Ni=1

    20=1,=1

    i,c+ici

    +N

    i=1

    4j=1

    20,=1

    ti,jc+icj (14)

    where c+i (ci) is the creation (annihilation) operatorof a carrier on the orbital localized on the site i, iis the corresponding on-site energy; and ti,j describesthe hopping of the particle between orbitals on neigh-boring sites. Coupling to farther neighbors is neglected.Finally, i, accounts for the spinorbit interaction byintroducing nite matrix elements connecting p or-bitals of opposite spin, residing on the same atom, fol-lowing the description given by Chadi [68]. For example,py, |H | , pz = i/3. Spinorbit-type coupling be-tween d orbitals is neglected. Here we assume that eachsite holds 20 orbitals and is surrounded by 4 neighbors.

    Hopping, i.e., o-diagonal matrix elements of ourHamiltonian are calculated according to the recipe givenby Slater and Koster [69]. In this approach, the hop-ping matrix elements ti,j are expressed as geometricfunctions of two-center integrals and depend only on therelative positions of the atoms i and j. Contributionsfrom three-center and higher integrals are neglected. Forexample, if the two atoms are connected by a bond alongthe x axis, then orbitals s and pz create a bond andthe matrix element ts,pz = Vs,pz = 0 vanishes becauseof the symmetry. On the other hand, if the direction ofthe bond is along y axis, then the bond is of a typeand ts,pz = Vs,pz is nite. In the general case the near-est neighbors are connected by bonds of any directiond = |d| (lx+ my + nz), with d being the bond length

  • Wei-dong Sheng, et al., Front. Phys., 2012, 7(3) 335

    and l, m, n the direction cosines. Then the tunnelingts,pz element can be expressed in terms of projecting thepz orbital onto the bond and in the direction perpendicu-lar to it. Since the perpendicular projections give -typebonds, their contribution is zero. The Hamiltonian ma-trix element is thus ti,s,j,pz = nVsp . Similar sets of rulesare dened for all other t matrix elements [69]. This ap-proach reduces the number of unknown matrix elementsas they can be related via SlaterKoster rules to a rel-atively small subset of two-center integrals V, . Thisis particularly useful within the framework of empiri-cal tight binding, where E, , and V, parametersare not directly calculated, but rather obtained by t-ting the TB bulk model results to experimentally knownband gaps and eective masses at high symmetry pointsof the Brillouin zone [70]. We stress here that we havebeen tting the TB model not only to bulk propertiesat point, but also at X and L points to account formultivalley couplings.

    The TB parametrization used so far are given, e.g., inRefs. [7173], where it was demonstrated that the inclu-sion of d orbitals in the basis allows to obtain much betterts of the masses and energy gaps to the target materialvalues. In particular, the treatment of the conductionband edge is signicantly improved, which is importantfor small nanostructures [74]. In this work we use ourown parametrization, analogous to work by Klimeck etal. [72], but giving a better agreement with target bulkproperties. More details will be presented in our futurework.

    In order to address the treatment of the interface be-tween InAs and GaAs we note that these two materialsshare the same anion (Arsenic). Thus during the ttingprocedure the diagonal matrix elements on arsenic arekept the same in both materials. This approach removesthe necessity of averaging on-site matrix elements for in-terface atoms. Additionally, to account for the band o-set (BO) between the materials forming the interface,the t for InAs is performed in such a way that the topof the valence band of InAs is set to be equal to the BOvalue relative to the GaAs. This removes the necessityof shifting values of diagonal matrix element for inter-face atoms, which would result in two dierent sets ofparameters for Arsenic: one for InAs and another forGaAs.

    Finally, there is the second type of interface that ariseson the edges of the computational box. Here, the ap-pearance of free surfaces leads to the existence of dan-gling bonds. Their presence results in spurious surfacestates, with energy inside the gap of the barrier ma-terial, making it dicult to distinguish spurious statesfrom the single-particle states of the quantum dot. Anenergy shift for dangling bonds that mimics the passiva-tion procedure, described in Ref. [75], is performed in or-der to move the energies of surface-localized states away

    from the energies corresponding to conned quantum-dot states.

    Then, a parallel Lanczos diagonalizer is used to resolvethe Kramers-degenerate doublets. Because the computa-tional domain necessary for the converged tight-bindingcalculation involves the order of 1 million atoms, theresulting tight-binding matrices are very large, i.e., 20million by 20 million. This presents a signicant numer-ical problem, but utilizing matrix sparsity, parallel com-putation paradigm, and the fact that we need only sev-eral lowest electron and hole states rather than the entireeigenspectrum of the Hamiltonian, we achieve linear scal-ing of the computational resources as a function of thenumber of atoms.

    To illustrate the application of the dierent methods,we show in Fig. 6 the results of calculation for the samedot obtained with the eective bond-orbital method (la-beled as EBOM), the tight-binding (TB) model, and theempirical pseudopotential method (EMP). The inputparameters in EBOM and TB calculations are chosen tobe similar to the ones used in empirical pseudopotentialcalculations (EMP1) from Ref. [64]. Another empiricalpseudopotential calculation (EMP2) is shown for com-parison [65]. As expected, the structure of electron statesis similar in all cases, however, the hole states dier sig-nicantly. EBOM maintains approximately a shell-likestructure of hole levels. This diers from both TB andEMP and can be attributed to the replacement of zinc-blende with cubic symmetry in EBOM. There is a goodagreement between TB and EMP1 calculation, with acharacteristic large/large/small level spacing betweensubsequent hole levels h1 h2 h2 h3 h3 h4.Surprisingly, two pseudopotential (EMP1/EMP2) cal-culations predict dierent details of hole levels, most

    Fig. 6 Electron (blue/upper) and hole (red/lower) single-particleenergies calculated for eective bond-orbital method (EBOM) andtight-binding (TB) model. The parameters of the TB model werechosen to be similar to empirical pseudopotential calculation from[64]; results of this calculation are shown as EMP1. Another em-pirical pseudopotential calculation (EMP2) is shown for compari-son [65]. Reproduced from Ref. [61], Copyright c 2010 AmericanPhysical Society.

  • 336 Wei-dong Sheng, et al., Front. Phys., 2012, 7(3)

    likely due to a slightly dierent choice of tting parame-ters or tted results in pseudopotential tting procedure.More specically, due to the insucient number of ttingparameters, the pseudopotentials used in the EMP cal-culations give eective masses of GaAs and InAs o fromthe experimental values by more than 30% [60]. Conse-quently, the separations among the energy levels in thevalence bands calculated by both EBOM and TB are dif-ferent from those by EMP. Overall there is a good agree-ment between EBOM, the tight-binding and empiricalpseudopotential models.

    2.3 Optical properties

    2.3.1 Photoluminescence: Polarization and anisotropy

    Let us begin our discussion with the EBOM model andrst temporarily neglect the electronelectron interac-tion. The momentum matrix element between an elec-tron state e =

    enun and a hole state h =

    hnun

    along the polarization direction e is given by

    h|e p|e =mn

    un|e p|umhn|em

    +m

    hm|e p|em (15)

    If we neglect the contribution from the envelope-functionpart of the wave function [76], this element can be furthersimplied as

    h|px|e = iP0 [hx|es+ hx|es

    hs|ex hs|ex]

    h|py|e = iP0 [hy|es+ hy|es

    hs|ey hs|ey]

    (16)

    where iP0 = s|px|x = s|py|y denotes the couplingbetween the conduction and valence bands. For circularpolarization + or , the momentum matrix element isthen given by phe = [h|px|e ih|py|e] /

    2. It is

    straightforward to show that phe = p+he in the absence of

    the magnetic eld.Figure 7 gives a schematic view of interband transition

    in an elongated quantum dot. The emission is found tobe partially polarized, which is regarded to be relatedto the structural anisotropy [77]. The degree of linearpolarization of interband transitions is then dened by

    Peh =|e|px|h|2 |e|py|h|2|e|px|h|2 + |e|py|h|2 (17)

    which is found to be closely related to the polarizationof envelope functions as given by [78]

    Pef =|hx|hx|2 |hy |hy|2|hx|hx|2 + |hy |hy|2

    (18)

    For the ground hole state in a quantum dot elongatedalong the x direction, the probability of the electron inthe px orbital is found to be larger than py [78]. Hence,it is this selective occupancy that leads to the opticalanisotropy.

    Fig. 7 Schematic view of the interband transition between theground electronic state e and hole state h and the intersubbandtransitions from e to the rst two excited states xe and

    ye in a

    quantum dot elongated along the x = [110] direction.

    Figure 7 also shows two intersubband transitions fromthe ground electronic state. These transitions are polar-ized along the long and short axes of the structure. In atraditional single-band picture, it is trivial to nd thatthe intersubband transitions from the ground state e toxe and

    ye are respectively polarized along the x and y

    directions due to the spatial symmetry of the states. Inthe presence of the mixing between the conduction andvalence bands, however, the polarization of intersubbandtransitions becomes a non-trivial issue.

    In the multiband formalism, the electronic states con-sist not only of components from the conduction band,but also those from the valence bands. Let us considerthe expansions, e =

    kS |k, xe =

    kX |k, and

    ye =

    kY |k where the summation is over the basis{s, x, y, z}. The momentum matrix elements for the tran-sitions e xe and e ye are then given by [79]xe |px|e = iP0

    [sX |xS xX |sS]xe |py|e = iP0

    [sX |yS yX |sS]ye |px|e = iP0

    [sY |xS xY |sS]ye |py|e = iP0

    [sY |yS yY |sS] (19)Here the momentum matrix elements among the enve-lope functions have been shown to be much smaller thanP0 and hence are neglected. It is therefore seen that thepolarization of the intersubband transitions now dependson those minor components of the electronic states, suchas yS and

    yX from the valence bands in the presence of

    the mixing between the conduction and valence bands.Further study has shown that the linearly polarized

    intersubband transitions are due to the simultaneous

  • Wei-dong Sheng, et al., Front. Phys., 2012, 7(3) 337

    vanishing of the four coupling terms, sX |yS, yX |sS,sY |xS, and xY |sS. The physics behind this rule canbe understood in terms of directional interactions be-tween the local atomistic orbitals [79].

    Figure 8 shows the linear polarization of the primaryinterband transition against the aspect ratio for the dotswith various lateral sizes and heights. It is seen that thelinear polarization of the primary interband transitionexhibits a quadratic dependence on the lateral aspectratio (). This dependence is also aected by other struc-tural parameters such as the lateral size and height. Moreimportantly, it is found that Peh() exhibits almost thesame behavior for dots of similar aspect ratio betweenthe lateral size and height [80]. This reveals the possibil-ity of optical characterization of structural properties ofself-assembled quantum dots.

    Fig. 8 Linear polarization of the primary interband transition,Peh in dotted lines, calculated as a function of the aspect ratioof the quantum dots with various lateral sizes and heights. Theshape and size of the symbols correspond to the denoted struc-tures. Quadratic t is shown in solid lines. Reproduced from Ref.[80], Copyright c 2008 American Physical Society.

    2.3.2 Electronelectron interactions and multiexcitoncomplexes

    The absorption and emission spectra of a quantum dotare determined by an exciton, an interacting electron-hole pair. There are four exciton states, two dark andtwo bright. The electronhole exchange interaction, tobe discussed later, leads to the splitting of the four ex-citon states into a dark doublet and a bright doublet[81], and modies the polarization of emitted photons.The emission spectra from a quantum dot with dierentnumber of excitons are dierent due to electron-electroninteractions and this allows us to identify the emissionfrom a single exciton, and hence the emission of a singlephoton. This is the principle behind the quantum dot asthe single photon emitter. An emission cascade from abiexciton through two potentially indistinguishable exci-ton states to the quantum dot ground state leads to theemission of a pair of entangled photons [82, 83].

    In a quantum-dot laser, the number of electronholepairs in a quantum dot increases with external excitationpower. The electrons and holes interact via Coulomb in-teraction, and these electronhole pairs form multiexci-

    ton complexes. A given N -exciton complex is equivalentto a specic articial excitonic atom. An understandingof the electronic properties such as the ground state en-ergy, total spin, total angular momentum, and emissionand absorption spectra requires an understanding of therole of electronelectron and electronhole interactionsin electronhole complexes occupying electronic shells ofa quantum dot. For a given number of electronhole pairsN the interacting Hamiltonian reads,

    Hex =

    i

    Eei c+i ci

    i

    Ehi h+i hi

    ijkl

    V heijklh+i c

    +j ckhl

    +12

    ijkl

    V eeijklc+i c

    +j ckcl +

    12

    ijkl

    V hhijklh+i h

    +j hkhl

    (20)

    where Eei and Ehi are the energy levels of the conned

    states in the conduction and valence bands, respectively.A general Coulomb matrix element is dened by

    Vij,kl =e2

    40

    i (r1)

    j (r2)

    1|r1 r2|k(r2)l(r1)

    (21)

    and we have V eeij,kl = Vij,kl. Note here that a connedstate in the valence band h obtained from any multi-band methods can be regarded as a hole state only aftera conjugate transformation, h h. Hence we ndthat V hhij,kl = Vkl,ij and V

    heij,kl = Vik,jl. Once the single-

    particle states are obtained, the Hamiltonian for multi-excitons can be solved by the congurationinteractionmethod [30, 46, 61, 84].

    A multiexciton complex, such as a biexciton, is com-posed of more than one electronhole pair. If CiN is thei-th eigenstate of an N -exciton complex, its wave func-tion can be generally written as iN =

    ciN |C iN . Re-

    combination of an electronhole pair from the N -excitoncomplex would reduce the number of excitons to N 1.By taking into account every possible recombination, onecan obtain the radiative lifetime, dened by

    1k

    =ne2

    20m0c3

    [f

    +k () + f

    k ()

    ]2d (22)

    with fk () being the oscillator strength,

    fk () =

    2m0

    i

    f(EiN )|ciN |2f

    |cfN1|2

    |CfN1|P |CiN |2(EiN EfN1 )(23)

    The radiative lifetime of a single exciton is typicallyaround 1 ns. Note that the nal state of the (N 1)-exciton complex, fN1, may not be its ground state afterthe recombination of an electronhole pair. The emissionintensity is given by

  • 338 Wei-dong Sheng, et al., Front. Phys., 2012, 7(3)

    IN () =

    i

    f(EiN )|ciN |2f

    |cfN1|2

    |CfN1|P |CiN |2 (EiN EfN1 )(24)

    where the probability function f(EiN ) is given byexp(EiN/T )/

    j exp(EjN/T ) and CiN is the i-th

    eigenstate of the N -exciton system. The operator Pdescribes all possible ways of electronhole recombina-tion, i.e., P =

    nm p

    nmhncm. In the absence of

    magnetic eld we have I+N (E) = IN (E). The emis-

    sion spectrum observed in experiments is a sum of thecontribution from individual multiple exciton complexesweighed by the corresponding occupation number nk,i.e., I() =

    nkIk(). Recent photon correlation ex-

    periments allow to untangle complicated emission spec-tra and extract spectra corresponding to xed excitonnumbers [85].

    2.3.3 Hidden symmetry

    The main diculty with determining the ground stateof the multiexciton complex exists for partially lled de-generate quantum dot states. All congurations havethe same energy and there is no single congurationwhich dominates the ground state. Fortunately, it wasshown that the fully interacting electronhole Hamil-tonian and exciton creation operator on any degener-ate shell satises the following commutation relation:[Hex, P+] = EXP+ where P+ creates an exciton andEX is a single exciton binding energy [28]. This commu-tation relation allows to construct exact eigenstates ofthe fully interacting Hamiltonian, called multiplicativestates. These states are the ground states of multiex-citon complexes and, as a consequence, emission from adegenerate shell takes place with energy EX independentof the population of this shell, i.e., the number of mul-tiexciton complexes N . This property of quantum dotsis called hidden symmetry. For more information on thehidden symmetry in quantum dots we refer the reader toRefs. [28, 30, 32, 61, 84].

    2.3.4 Fine structure: Electronhole exchangeinteraction

    Let us denote the total energy of an ideal ground state ofa semiconductor, i.e., a fully occupied valence band plusan empty conduction band, as E0. If one removes an elec-tron from a state i in the valence band to a state j inthe conduction band, the energy of the system changesto Eij . If one adds an electron back to the same statein the valence band, the dierence between Eij and E0would just be the interaction energy between two elec-trons, one in state i and the other one in state j, i.e,E0 +C = Eij +Vij,ij Vij,ji. Here C is the energy of in-

    teractions between the electron in the conduction band,and each electron in the valence band, and can be re-garded as a constant. Hence, the energy of an electronhole pair is given by Vij,ji + Vij,ij with the rst termbeing the electronhole direct attraction and the otherone the electronhole exchange repulsion. In general,the matrix element for the electronhole exchange inter-action is given by

    Vij,kl =e2

    40

    i(r1)j (r2)

    1|r1 r2|

    k(r2)l(r1)

    (25)

    In bulk, the electronhole exchange interaction arisesfrom the overlap between the s orbitals in cations, wheremost of the electron wave function is localized, and p or-bitals in anions, accounting for most of the wave functionof the hole. This also accounts for the mixing betweenconduction and valence bands.

    The electronhole exchange interaction is responsiblefor the ne structure in the optical spectra of an exciton.Without the exchange interaction, the exciton composedof an electron with spin s = 1/2 and a heavy holewith spin = 3/2 would be four-fold degenerate. WithL = s+ as the total angular momentum, the four statesform a bright doublet with L = 1 and a dark doubletwith L = 2. The electronhole exchange interactionsplits the dark and bright doublets, with dark excitons atlower energy. Even more importantly, the bright doubletis also split into two linearly polarized exciton states bythe long-range electronhole exchange interaction. Thesplitting of the two exciton doublets is a function ofthe anisotropy of the quantum dot. Since the splittingprevents the emission of entangled photon pairs in thebiexciton cascade, the theory of the ne structure of ex-citon has been extensively studied, e.g., by Takagahara[86, 87], Ivchenko and co-workers [88], and Kadantsevet al. [89] and analyzed in atomistic approaches [61, 90].Since the ne structure splitting is on the order of tensto hundreds of eV, cautions must be taken in extract-ing numbers from multi-million atom simulations. Sig-nicant eort is also devoted to controlling the excitonne structure [9193].

    2.4 Quantum dots in magnetic elds

    The eect of magnetic eld has been incorporated intothe EBOMHamiltonian by introducing Peierls phase fac-tors as follows:

    H(R,R) ei eRRR

    A(r)dr H(R,R) (26)where A(r) is the vector potential. The Zeeman eect isincluded by adding the spin terms to the diagonal matrixelements,

    H(R(),R()) H(R(),R()) 12g0BBz (27)

  • Wei-dong Sheng, et al., Front. Phys., 2012, 7(3) 339

    where g0 is the g factor of a bare electron. The mag-netic properties of quantum dots such as the eectiveelectron g factors depend sensitively on the parametersused in the calculation. In the EBOM Hamiltonian allthe tting parameters are established in a closed form asa function of band edges and eective masses, and thuscan be uniquely determined.

    2.4.1 Multiexciton FockDarwin spectrum

    Multiexciton complexes in quantum dots have been in-vestigated by us in Refs. [48, 94]. As mentioned previ-ously, the connement for the electrons in the conductionbands in a at quantum dot can be well approximated asa two-dimensional parabolic potential. The energy spec-trum of conned electronic states is therefore the same asthat of a harmonic oscillator, which is also known as theFockDarwin spectrum in the presence of applied mag-netic eld. In experiments, shown in Fig. 4, one observesthe emission spectrum from multiple exciton complexesinstead of the energy spectrum of single-particle states.Nevertheless, there is a correlation between the spectraof single-particle states and emission of multiple excitons.According to Koopmans theorem for a few-particle sys-tem, it is a good approximation to treat the additionenergy, EN EN1, i.e., emission energy of multipleexcitons as the corresponding single-particle excitationenergy, Een Ehn . Hence, it is not surprising to observeFockDarwin-like spectra in experiments on ensemble ofquantum dots [48] as well as in single dots [95].

    Fig. 9 Contour plot of the multiexciton emission spectrum of alens-shaped InAs/GaAs self-assembled quantum dot.

    Numerical simulation of multiexciton emission in-volves several steps from determination of strain dis-tribution, calculation of the single-particle energy spec-trum, computing multiexciton states by using theconguration-interaction method, solving rate equationsto determine occupation of multiple exciton states tonally obtaining the emission spectrum [94]. Figure 9shows the multiexciton emission spectrum calculated fora lens-shaped InAs/GaAs self-assembled quantum dotwhich is grown on a 2 ML wetting layer and has a base

    diameter of 19.8 nm and height of 3.4 nm. For a dot ofsuch small lateral dimensions, the diamagnetic shift ofthe single exciton line is barely noticeable. However, thesplitting of the two p-like states and crossing of the p andd orbitals is observed.

    2.4.2 Electron g factors: Distribution and anisotropy

    In bulk semiconductors such as GaAs and InAs, the elec-tron g factor can deviate substantially from the bare elec-tron g factor due to strong band mixing eects and thespinorbit interaction. In turn, in self-assembled quan-tum dots the complicated environment like the strongquantum connement and long-ranged strain eld makeselectron g factors greatly dierent from those in the bulk[96]. Interestingly, in InAs/GaAs self-assembled quan-tum dots the ensemble average value of the electron gfactor does not deviate too much from that in single dots[97]. It implies that the g factors for individual islandsare similar, and therefore the inhomogeneous broaden-ing of the quantum dot size and composition does notinuence the measured g factors signicantly.

    Fig. 10 Electron g factors as a function of the single excitonemission energy calculated for InAs/GaAs quantum dots with var-ious sizes, shapes and composition proles. Circles and diamondsstand for lens-shaped and pyramidal dots, respectively. The size ofthe symbols corresponds to the denoted structures.

    Figure 10 shows the electron g factors calculated asa function of the emission energy for an ensemble ofInAs/GaAs dots with various sizes, shapes and com-position proles [98]. For all the samples, the electrong factors are found to carry a negative sign and havemagnitudes smaller than 2.0. Except for some extremelylarge or small samples, the electron g factors fall between0.5 and 1.0. To understand this interesting behavior,we need to recall that an electronic state e in quantumdots, in the coupled spinorbital basis, is composed ofcomponents from the conduction bands (s), heavy (hh),light (lh), and split-o (sh) hole subbands,

    e = s|s+ hh|hh+ lh|lh+ sh|sh (28)The electronic state is dominated by the s componentfrom the conduction bands followed by the light and

  • 340 Wei-dong Sheng, et al., Front. Phys., 2012, 7(3)

    heavy hole components from the valence bands. In thebulk, the holes in dierent bands have distinctive g fac-tors, i.e., gshh = 6, gslh = gssh = 2 where is anempirical parameter. In quantum dots we propose thatthe longitudinal electron g factor can be expressed as

    ge = 2.0 |s|s|2 6 |hh|hh|22 |lh|lh|2 2 |sh|sh|2 (29)

    with now being a tting parameter. By tting this em-pirical formulation with the result from numerical calcu-lation, we nd a weak dependence of on the dimensionsof the dots. Typically, it falls between 9.4 and 10.1. Thetheory therefore well explains why the uctuation of size,shape, and even composition prole does not have muchinuence on the electron g factor [99].

    Electron g factors are almost isotropic in bulk semi-conductors such as InAs or GaAs. However, like the elec-tron eective mass [44], they become anisotropic in self-assembled quantum dots. The in-plane electron g factoris given by

    g//e = 2|s|s|2+g//lh|lh|lh|2+g//sh|sh|sh|2(30)and further we have an anisotropic electron g factor [100]

    ge g//e = 6 |hh|hh|2 + 2 |lh|lh|2 (31)As g//lh = 4 and g//sh = 2 are isotropic within thegrowth plane of the dots, the in-plane electron g factorremains isotropic within the lateral dimensions.

    2.4.3 Hole g factors: Envelope orbital momentum

    In contrast with electronic states, the valence-bandstates have more complicated composition structuredue to strong mixing between heavy- and light-holesubbands. As these two bands have distinctive g fac-tor tensors [see Eq. (29)], the hole states in quantumdots thus exhibit interesting behavior. Figure 11 showsthe hole g factors calculated for the same set of thequantum dots as in Fig. 10. Compared with electrong factors, the hole g factors are seen to be distributed

    Fig. 11 Hole g factors against the single exciton emission en-ergy calculated for the same set of InAs/GaAs quantum dots as inFig. 10.

    in more wider region, ranged from 2.5 to 1.5. Further-more, their dependence on the emission energy is foundto be more irregular.

    In bulk InAs and GaAs both electron and hole g fac-tors are negative. In InAs/GaAs quantum dots, electrong factors are found to always carry a negative sign. Thereason why some dots can have positive hole g factors liesin the nonzero envelope orbital momenta (NEOM) car-ried by the components of the hole states [101]. Keepingin mind the expansion of any single-particle wave func-tion as in Eq. (28), in general, one only needs to considerthe contribution from the Bloch functions to the overallg factor because envelope functions usually do not carryany NEOM, and therefore have no eect on the g factor.

    However, we nd that the light-hole component of theground hole states in a quantum dot does carry NEOMdespite the fact that the heavy-hole part does not. Thereason why the heavy hole component of a zero angularmomentum can mix with the light-hole parts of nonzeroangular momenta lies in the fact that the total angularmomentum, i.e., that of the envelope part plus that ofthe Bloch part, is a good quantum number in a systemwith cylindrical symmetry. Even when the symmetry isbroken by the shear strain, we nd that the conservationof the total angular momentum is still a good approx-imation. Including the contribution from the envelopefunctions, we have the overall hole g factor

    gh = gsh + goh (32)

    where goh denotes the contribution from NEOM,

    goh = 2

    nhh,lh,sh,sn|Lz|n+ n|Lz|n (33)

    Although the light-hole component carries NEOM, itscontribution to the overall g factor is small because ofits small projection in the hole state. It is found thatthe heavy-hole part gains more and more NEOM as theheight of the dot increases while the lateral dimensionis xed. If the contribution from NEOM carried by theheavy-hole component exceeds that from the Bloch partin absolute amplitude, the overall hole g factor is seen tochange its sign and becomes positive. If the positive holeg factor and the negative electron g factor are similar inamplitude, the overall exciton g factor can even vanish[101].

    Apart from the sign change, the hole g factor also ex-hibits an interesting anisotropy. As the biaxial strain inquantum dots splits the heavy and light hole apart, thelow-lying states in the valence bands are dominated bytheir heavy-hole components (see Fig. 5). Recall that thebasis functions for the heavy-hole band at the pointare given by |x + i|y and |x i|y , with orbitalangular momenta oriented along the growth (z) direc-tion. The spinorbit interaction couples the spin to theorbital angular momentum and results in the z direction

  • Wei-dong Sheng, et al., Front. Phys., 2012, 7(3) 341

    being the easy-axis of polarization of total angular mo-mentum (quasi-spin). If a hole state is composed of onlythe heavy-hole component, it would lead to the hole spinfrozen along the growth direction and thus a zero g factorin the presence of a magnetic eld applied in the growthplane [102].

    For a hole state in a realistic quantum dot, its mi-nor light-hole and split-hole components help to retain asmall g factor in the Voigt conguration. Unlike an al-most isotropic in-plane electron g factor [99], the in-planehole g factor can be highly anisotropic due to NEOMcarried by mainly the light-hole parts [103].

    2.5 Quantum dots in electric elds

    As in bulk materials and quantum wells, electronic andoptical properties of self-assembled quantum dots canbe probed and then tuned by applying external electricelds. The electric eld can be applied either along thegrowth direction or in the growth plane.

    2.5.1 Quantum-conned Stark eect

    In the presence of an external electric eld F appliedalong the z (growth) direction, the Hamiltonian for anelectron conned in quantum dots is given by H + qeFz.For at quantum dots with strong connement along thevertical direction, the additional term for the electriceld qeFz can usually be regarded as a perturbation.Hence the energy of the electron ground state 0 in theelectric eld becomes

    E0 0|H|0+ qeF 0|z|0

    +q2eF2n=1

    0|z|nn|z|0E0 En (34)

    with E0 = 0|H |0 being the energy in the ab-sence of the eld. The coecients of the linear andquadratic terms are given by = qez and =q2e

    n=1 |z0n|2/E0n, respectively. Combining the for-mulations for electrons and holes together, we have thetransition energy for an electronhole pair given by

    Ee Eh = (Ee Eh) + pF + F 2 (35)where p = qe(ze zh) is the built-in dipole moment,and = e h measures the polarization of the elec-tron and hole states. As the eect of electric eld on theexciton binding energy can be approximated by a similarexpansion, the above expression may also be used for theemission energy of single excitons. If there is a non-zerobuilt-in dipole moment in the dot, the Stark shift wouldbe asymmetric [104].

    Since in the Stark shift the linear term in electriceld is the electronhole dipole moment, the informa-tion about the relative positions of the electron and hole

    states in the quantum dots can be determined by experi-ments. For a quantum dot with a homogeneous composi-tion prole, the ground hole state is found to be localizedcloser to the bottom of the dot than the electron state[40]. This is due to the fact that the electron state is af-fected by only the hydrostatic strain while hole states areheavily inuenced by the biaxial strain component. Sur-prisingly, an inverted electronhole alignment was foundin an experiment on the Stark eect in self-assembledquantum dots [105], which was later attributed to theinter-diusion eect [53].

    The lateral electric eld has also been applied to exci-ton complexes in quantum dots. The two main reasonsare i) attempt to modify the anisotropy of the quantumdot and hence the exciton ne structure [86, 106, 107]and ii) modication of the biexciton binding energy [35,114]. It was shown that the electric eld alone cannotremove the exciton ne structure splitting. However, theremoval of the biexciton binding energy provides an al-ternative route to generation of entangled photon pairswithout the need for the modication of quantum dotstructural parameters such as anisotropy [35, 114].

    2.5.2 Electrical tuning of exciton g factors

    The eect of an applied electric eld on a quantum dotis not limited to the Stark shift observed in the opticaltransitions. Other properties such as polarization of thephotoluminescence emission and even exciton eectiveg factor may also be aected. Electrical tuning of exci-ton g factors in single [108] and stacked [109] InAs/GaAsquantum dots has already been demonstrated in recentexperiments.

    The spin splitting of an exciton state consists of thecontributions from both electron and hole states, i.e.,gx = ge+gh. Although it is still very dicult to measureelectron and hole g factors independently in experiment,ge and gh can be calculated separately. Our model sys-tem here is a pyramidal dot with a base length of b=20 nm. Figure 12 plots the hole g factor as a function ofthe applied eld for the dot of various height. For all the

    Fig. 12 Hole g factor as a function of the applied electric eldfor a pyramidal InAs/GaAs quantum dot of various height whichis illustrated schematically in the inset.

  • 342 Wei-dong Sheng, et al., Front. Phys., 2012, 7(3)

    dots, we nd that electron g factor is strongly resistive tothe applied electric eld and exhibits very little changeover a broad range of the eld strength, and thereforedo not show its result in the gure. On the contrary, thehole g factor is seen quite sensitive to the applied eldand even changes its sign in the dots of high aspect ratio.

    The g factors of electronic states are known to be in-sensitive to the size, geometry, and even the compositionprole of the quantum dots [99]. The independence ofge on the applied electric eld can be attributed to thesame physical mechanism. The hole g factor has beenshown to increase with the height of the dot [101]. Thisdimensional dependence of the hole g factor is explainedin terms of NEOM carried in the ground state of holes.As the eect of an applied electric eld is equivalent tothe change in the eective connement, the electric de-pendence of the hole g factor can be understood in thesimilar way [110].

    For eective tuning of exciton g factors in single quan-tum dots a relatively strong electric eld is usually re-quired. It has already been shown that, with the sameapplied electric eld, a double dot often exhibit a largerStark shift compared with the single dot of the samedimension. The model system adopted here involves twocoupled disk-like InAs quantum dots, each having a di-ameter of 15.3 nm, separated by a GaAs barrier witha thickness of 4.5 nm. Figure 13 plots the g factors ofthe ground and rst excited hole states calculated as afunction of the electric eld for the coupled dots with aseparation of d = 4.5 nm. A resonance structure can beseen in the spectra of hole g factors, which is very dier-ent from the monotonic dependence of their counterpartsin the single quantum dots. The probability density ofthe ground and rst hole excited states at F = 0, 10kV/cm (o resonance) and F = 5.4 kV/cm (on reso-nance) is plotted to highlight the origin of the g factor

    Fig. 13 g factors of the ground and rst excited hole states in alens-shaped double InAs/GaAs quantum dot, calculated as a func-tion of the applied electric eld. Probability density of the statesat the various electric elds are illustrated. Reproduced from Ref.[111], Copyright c 2009 Institute of Physics.

    resonances [111]. It is noted that recent theoretical work[112] by using the eight-band k p method reports onthe exciton g factors of the opposite sign to our resultand the experiment [109]. Finally, we would like to pointout that a magnetic impurity may lead to dramatic vari-ation in electron g factors [113] though electrical tuninghas been shown to be nearly ineective.

    2.6 Single InAs/InP self-assembled quantum dots onnanotemplates

    Formation of self-assembled quantum dots during theStranskiKrastanow growth of epitaxial layers is a veryuseful way of controlling matter in three dimensions inthe one-dimensional growth. The disadvantage of this ap-proach is the random in-plane nucleation of quantumdots and the variation of their sizes. Williams and co-workers [34, 115] proposed a dierent approach, involv-ing growth of single InAs quantum dots on InP nan-otemplates. The fabrication of nanotemplates starts withlithographically dened patterns for the initialization ofthe growth of InP pyramid. The starting area of the pyra-mid is dened with the accuracy of tens of nanometers.The growth of the pyramid is governed by the stability ofcrystallographic facets and, when interrupted, results ina formation of nanotemplates dened with atomic preci-sion. InAs dots are grown on such nanotemplates and arecovered with InP to complete the pyramid. Such struc-ture, shown in Fig. 14(a), can be functionalized by eithermetallic gates or by building photonic crystals aroundthem for a fully controlled light-matter system [36]. Therst steps toward the theory of InAs quantum dots onInP templates have been made by some of us in Ref. [42].

    As an example of a deterministically functionalizedsingle quantum dot we show the result of the applicationof gates to the single quantum dot, resulting in a fullytunable optical structure. Figure 14(b) shows the emis-sion spectrum of a single photo-excited quantum dot as afunction of the bottom gate voltage. We see a number ofemission lines, which correspond to charged exciton com-plexes X0, X, X2, or dierent number of electrons Nin the initial states (right-hand axis). The black circlesand lines create a phase diagram of the emission spec-trum as predicted in Ref. [27]. The emission for X2

    consists of two lines, corresponding to the singlet andtriplet two-electron complexes in the nal state, respec-tively. Hence, the electron spin can be detected optically.The eect of spin can also be seen in the emission fromother multiexciton complexes.

    Because the nanotemplate can be elongated, it al-lows for the control of the shape of the quantum dot.One would expect that the resulting splitting of the twop-shell states would be a measure of the elongation.Unfortunately this is not so straightforward. Figure15(a) shows the InAs quantum dot on an elongated

  • Wei-dong Sheng, et al., Front. Phys., 2012, 7(3) 343

    Fig. 14 (a) Single InAs quantum dot in an InP pyramid withmetallic gates on the sides of the pyramid. (b) Emission spectrumas a function of the bottom gate voltage. Dierent emission linescorrespond to dierent charged exciton states, X0,X, X2, ordierent number of electrons in the initial states (right-hand axis).The black circles and lines create a phase diagram of the emissionspectrum as predicted in Ref. [27]. Reproduced from Ref. [116],Copyright c 2009 Institute of Physics.

    template. Figure 15(b) shows the emission from the ve-exciton complex in which three electrons and three holespopulate the two p-shell states. We see that even fora fully symmetric quantum dot there are two emissionpeaks. These two peaks correspond to two dierent four-exciton nal states, singletsinglet and triplettriplet,which dier by exchange and correlation energy. As thestructure becomes highly asymmetric, we can attemptto extract the p-shell splitting from the emission spec-tra. Much work is needed for the full understanding ofInAs/InP quantum dots on patterned substrates to real-ize this promising technology.

    3 Graphene quantum dots

    3.1 Introduction

    Carbon atom, a basis of organic chemistry, gives rise toa rich variety of chemical structures, allotropes, due tothe exibility of its -bonds. One of the allotropes isgraphene, a two-dimensional sheet of carbon atoms ar-ranged in a hexagonal honeycomb lattice. The theory ofgraphene has been developed at the National ResearchCouncil of Canada by Wallace as early as 1947 [117].Since then graphene was considered as a starting pointin the understanding of other allotropes such as graphite

    Fig. 15 (a) Schematic view of a single InAs quantum dot ona InP pyramid template. (b) Emission spectrum from the ve-exciton to the four-exciton complex as a function of the elongationof the template. The two emission peaks correspond to two four-exciton nal states, determined by spin for symmetric structureand splitting of the p shell for asymmetric structure. Reproducedfrom Ref. [42], Copyright c 2005 American Physical Society.

    (stack of graphene layers), carbon nanotubes (rolled-upcylinders of graphene), and fullerenes (wrapped grapheneby the introduction of pentagones on the honeycomblattice). Graphene has been investigated extensively inGraphite Intercalation Compounds (GIC) [118]. Inter-calation of graphite with, e.g., Lithium or Potassiumleads to increased separation of graphene sheets and in-troduction of electrons or holes. GIC were equivalent todoped semiconductors. The theory of optical propertiesof graphene in GIC was developed by Blinowski and co-workers [119] and by one of us [120] and has been com-pared with reectivity experiments. The theory of elec-tronic screening and plasmons in GIC was also developedat that time [121]. Recently, following the rst experi-mental isolation of a single graphene sheet in 2004 byGeim, Novoselov and co-workers [37], both experimen-tal and theoretical research on graphene has increasedexponentially due to the unique physical properties andpromising potential for applications [122].

    One of the most interesting electronic properties ofgraphene is the zero-energy gap and relativistic natureof quasi-particle dispersion close to the Fermi level pre-dicted by Wallace [117]: low energy excitations are mass-less Dirac fermions, mimicking the physics of quantumelectrodynamics at much lower velocities than light. Onthe other hand, with ongoing improvements in nanofabri-cation techniques [123], the zero-energy gap of the Diracquasi-particles can be opened by engineering the size,

  • 344 Wei-dong Sheng, et al., Front. Phys., 2012, 7(3)

    shape, edge, and carrier density. This in turn oers pos-sibilities to control electronic [38, 124128, 130132],magnetic [39, 123, 130138] and optical [138142] prop-erties of a single-material nanostructure simultaneously.As a result, there is a growing interest in studying lower-dimensional structures such as graphene ribbons [144147], and more recently graphene quantum dots [126129, 130]. In the following we will concentrate on theelectronic structure, optical properties and magnetism ofgraphene quantum dots. For transport properties, pleaserefer to Ref. [143] for more detail review.

    3.2 Electronic structure Tight-binding approach

    The sp2 hybridization of carbon atom 2s and px, py or-bitals leads to trigonal -bonds responsible for a robusthexagonal lattice of carbon atoms. The remaining pz or-bitals from each carbon atom form the -band, whichis well separated from the lled -band. Hence, the elec-tronic states of graphene in the vicinity of the Fermi levelcan be understood in terms of pz electrons on a hexag-onal lattice. The hexagonal honeycomb lattice consistsof two triangular sublattices as shown in Fig. 16. Redatoms form the sublattice A, and blue atoms form thesublattice B. There are two carbon atoms, of type A andB, in a unit cell. Following Wallace [117], the wavefunc-tion of graphene can be written as a linear combinationof pz orbitals localized on sublattices A and B:

    k(r) =RA

    A(RA)z(r RA)

    +RB

    B(RB)z(r RB) (36)

    where z is a localized pz orbital. R represents the po-sition of an atom on sublattice = A or B, and is afunction of the lattice vectors a1,a2, and b shown in Fig.16. Within the tight-binding formalism, the Hamiltoniancan be written as:

    H = i,j,

    tijcicj (37)

    Fig. 16 Honeycomb lattice structure of graphene. The two inter-penetrating triangular sublattices are illustrated by blue and redcarbon atoms.

    where the electrons can hop between sites i and j via thetunneling matrix element ti,j . For graphene, the nearestneighbor hopping energy is about t 2.8 eV and thenext-nearest neighbor hopping energy t 0.1 eV [122].For simplicity, let us consider only the rst nearest neigh-bor interaction.

    For bulk graphene the wave function that satises theperiodicity of both sublattices can be classied by thewave vector k and written as:

    k(r) = AkRA

    eikRAz(r RA)

    +BkRB

    eikRB+bz(r RB) (38)

    Solving the Schrodinger equation for graphene involvesnding the solution for the coecients of the two sublat-tices for each wave vector k:(

    0 tf(k)

    tf(k) 0

    )(Ak

    Bk

    )= E

    (Ak

    Bk

    )(39)

    where

    f(k) = 1 + eika1 + eika2 (40)

    The energy spectrum Ek and the wave function are thengiven by

    E(k) = t|f(k)| (41)Ak = eikBk (42)

    with k being the phase of f(k). The energy spectrumE(k) is shown in Fig. 17. For the charge neutral system,each carbon atom gives one electron to the pz orbital. Asa result, the Fermi level is at E(k) = 0, and the signcorresponds to the electron and hole branches, respec-tively. Although the electron-hole symmetry is conservedhere, if one includes second neighbor interaction t, theelectron and hole branches become asymmetric.

    Fig. 17 Energy spectrum of graphene obtained from the nearest-neighbor tight-binding model.

    3.3 Dirac fermions

    One of the most striking properties of graphene is re-

  • Wei-dong Sheng, et al., Front. Phys., 2012, 7(3) 345

    vealed if we expand the energy spectrum close to Fermilevel Ek = 0, around the K and K points of the Bril-louin zone. In contrast with semiconductor quantumdots, where the energy spectrum E(q) = q2/(2m) isquadratic, in graphene we nd a linear dispersion

    E(q) vFq (43)where q is the momentum relative to one of the K orK points, called Dirac points, and vF = 3tb/2 106m/s is the velocity of Dirac fermions. We thus obtainrelativistic-like dispersion relation with a velocity 300times smaller than that of light. We can also expandEq. (39) around one of the K points which gives thetwo-dimensional Dirac equation:

    ivF (r) E(r) (44)The wave functions around K and K points are givenby

    (k) 12

    (eik

    1

    )around K (45)

    12

    (e+ik

    1

    )around K (46)

    As discussed above, the wave function for Dirac Fermionshas two components corresponding to the two sublat-tices. The two-component character is often referred to aspseudo-spin. In analogy with real spin, the Dirac Fermionwave function acquires Berrys phase if we adiabaticallychange the wave vector k along a closed loop enclosingthe Dirac point. Hence the sublattice structure adds anontrivial topological eect to the physics of graphene.

    3.4 Graphene quantum dots

    As we can see from Eq. (43) and Fig. 17, graphene is agapless material. As a result, in analogy with the phe-nomenon of Klein tunneling for massless particles in rel-ativistic quantum mechanics, it is not possible to conneDirac electrons in graphene electrostatically using metal-lic gates as in semiconductor quantum dots. Various indi-rect ways were proposed for opening a gap to conne theelectrons, such as size quantization in a graphene ribbon[144147] or using bilayer graphene [148155]. In this re-view we will focus on graphene islands with size quantiza-tion in the two dimensions, i.e., graphene quantum dotswith edges created by, e.g., etching [123]. Quantum con-nement in such systems was experimentally observed[38, 124, 125] and there is increasing interest in creat-ing quantum dots with well controlled shapes and edges[123, 129].

    3.5 Shape and edge eects

    Graphene sheet can be cut along dierent crystallo-

    graphic directions as seen in the top panel of Fig. 18,resulting in dierent types of edges. The most stableedge types are armchair and zigzag edges [156, 157]. To-gether with the shape of the quantum dot, the edge typeplays an important role in determining the electronic,magnetic, and optical properties. In this section, we willcompare the properties of quantum dots of various shapeand edge, and study their properties as a function of theirsize. We consider three dierent quantum dots as illus-trated in Fig. 18: (a) hexagonal dot with armchair edges,(b) hexagonal dot with zigzag edges, and (c) triangulardot with zigzag edges. Their energy spectra, shown onthe lower panel of Fig. 18, were calculated by numeri-cally solving the tight-binding Hamiltonian of Eq. (37)in the nearest neighbors approximation. Clearly the en-ergy spectrum around the Fermi level (E = 0) stronglydepends on the structure. While the hexagonal armchairand trigonal zigzag dots have a well dened gap of theorder of 0.2t (shown by a red arrow), for the hexagonalzigzag dot the gap is much smaller. Moreover, in additionto valence and conduction bands, the trigonal zigzag dotspectrum shows a shell of degenerate levels at the Fermilevel [39, 130, 132136, 138]. As we will see in the fol-lowing section, this degenerate band is responsible forinteresting electronic and magnetic properties.

    Fig. 18 Single-particle tight-binding spectrum of (a) arm-chair hexagonal, (b) zigzag hexagonal, and (c) zigzag triangulargraphene quantum dot structures consisting of similar number ofcarbon atoms. Top panel shows the atomic positions. Reproducedfrom Ref. [133], Copyright c 2010 American Physical Society.

    In order to understand further the electronic proper-ties near the Fermi level, in Fig. 19 we plot the energy gapas a function of the number of atoms. For the hexagonalarmchair dot (red dots), the gap decays as the inverseof the square root of number of atoms N , from hundredatom to million atom nanostructures. This is expectedfor conned Dirac fermions with photon-like linear en-ergy dispersion (Egap kmin 2/x 1/

    N), as

    pointed out in Refs. [130, 139, 158]. However, the re-placement of the edge from armchair to the zigzag has

  • 346 Wei-dong Sheng, et al., Front. Phys., 2012, 7(3)

    a signicant eect on the energy gap. The energy gap ofhexagonal structure with zigzag edges decreases rapidlyas the number of atoms increases. This is due to thezigzag edges leading to localized states at the edge ofthe quantum dot, similar to whispering gallery modesof photons localized at the edge of photonic microdisk[159]. Figure 19 also shows the eect on the energy gapof deforming the hexagonal structure into a triangle whilekeeping zigzag edges. Unlike the hexagonal zigzag struc-ture, all three edges of the triangle are composed of atomsof the same sublattice. As a result, they do not hybridizeand form a degenerate band at Fermi level. The energygap shown in Fig. 19 corresponds to transitions from thetopmost valence to the lowest conduction band state.The energy gap in the triangular zigzag structure followsthe power law Egap

    N since it is due to connement

    of bulk states, whereas the gap for the hexagonal arm-chair structure is due to the hybridization of the edgestates. We note that the energy gap changes from 2.5eV (green light) for a quantum dot with 100 atomsto 30 meV (8 THz) for a quantum dot with a millionatoms and a diameter of 0.4 micrometer. The pres-ence of a partially occupied band of degenerate statesin the middle of a well dened energy gap oers uniqueopportunity to control magnetic and optical propertiesof triangular graphene nanostructures simultaneously.

    Fig. 19 Energy band gap as a function of number of carbonatoms for triangular zigzag, hexagonal armchair and hexagonalzigzag structures. Reproduced from Ref. [133], Copyright c 2010American Physical Society.

    3.6 Gated quantum dots: Beyond tight-bindingapproach

    While the basic tight-binding model of Eq. (37) ac-curately describes the electronic properties of bulkgraphene, the situation can be more complicated forgated nite-size systems. The experimental values usedfor rst and second nearest neighbors hopping terms tand t are based on the assumption that electronic oc-cupation on every site is constant (equal to one), whichmay not be true at the edges of a quantum dot. More-

    over, for a gated system, the average occupation per sitecan be higher or lower than one. Thus, one must takeinto account the deviation of the electronic density fromthe bulk values due to edge eects and doping via theelectrostatic gate [39].

    Interaction eects due to additional charge density oneach site can be investigated within a mean-eld ap-proach using a combination of tight-binding and self-consistent HartreeFock methods (TBHF). For pz elec-trons, the interacting many-body Hamiltonian can bewritten as

    H = il

    tilcicl

    +12

    ijkl

    ij|V |klcicjckcl (47)

    and the corresponding mean-eld HartreeFock Hamil-tonian is

    HMF = il

    tilcicl +

    il

    jk

    jk (ij|V |kl

    ij|V |lk)cicl (48)where the operator ci creates a pz electron on site i withspin . Note that at this stage the unknown hoppingterms til do not include the eect of electronelectroninteractions. We want to express the Hamiltonian for oursystem as a function of experimentally measured bulktight-binding parameters il . For the graphene sheet,the mean-eld Hamiltonian is written as

    HbulkMF = il

    tilcicl +

    il

    jk

    bulkjk (ij|V |kl

    ij|V |lk)cicl (49)

    il

    ilcicl (50)

    We can now re-express our mean-eld Hamiltonian as

    HMF = HMF HbulkMF + HbulkMF = il

    ilcicl

    +il

    jk

    (jk bulkjk )(ij|V |kl

    ij|V |lk)cicl (51)which must be solved self-consistently to obtain Hartree-Fock quasi-particle states.

    In order to take into account the eect of induced gatecharge away from charge neutrality, we assume that elec-trons in the graphene island interact with the electronstransferred to the gate via the term vgii given by

    vgii(qind) =Nsitej=1

    qind/Nsite

    (xi xj)2 + (yi yj)2 + d2gate(52)

  • Wei-dong Sheng, et al., Front. Phys., 2012, 7(3) 347

    where (xi, yi) are the coordinates of the atoms. Thismodel assumes that the induced charge qind = N issmeared out at positions (xi, yi) at a distance dgate fromthe quantum dot. Our nal HartreeFock Hamiltonian isthen given by

    HMF = il

    ilcicl

    +il

    jk

    (jk bulkjk )(ij|V |kl

    ij|V |lk)cicl +i

    vgii(qind)cici (53)

    After self-consistent diagonalization of the Hamilto-nian as in Eq. (53) for a TBHF reference level dened byN = Nref , we obtain TB+HF quasi-particles denoted bythe creation operator bp, with eigenvalues p and eigen-functions |p. We can then start lling the conductionstates above Fermi level one by one to investigate cor-relation eects. The general Hamiltonian for the wholesystem can be written as

    H = H + HMF HMFfrom which, after extensive algebra, it can be shown thatin the rotated basis of bp quasi-particles and by neglect-ing scatterings from/to the degenerate shell, the Hamil-tonian for Nadd + Nref electrons reduces to the congu-ration interaction problem for the Nadd added electronsgiven by

    H =p

    pbpbp +

    12

    pqrs

    pq|V |rsbpbqbrbs

    +pq

    p|vg(Nadd)|qbpbq

    +2pp|vg(Nadd)|p (54)

    where the indices without the prime sign (p, q, r, s) runover states above the TBHF reference level, while theindex with the prime sign p runs over valence states(below the TBHF reference level). Here the rst termon the right-hand side represents the energies of quasi-particles; the second term describes the interaction be-tween the added quasiparticles, the third term describesthe interaction between the quasiparticles and the gate;and the last term is a constant giving the interactionenergy between the Nref electrons and the charge onthe gate. We then build all possible many-body con-gurations within the degenerate shell for a given elec-tron number Nadd, for which Hamiltonian matrices cor-responding to dierent Sz subspaces are constructed anddiagonalized.

    3.7 Magnetism in triangular quantum dots

    As discussed in the previous section, when an electron is

    conned to a triangular atomic thick layer of graphenewith zigzag edges, its energy spectrum collapses to a shellof degenerate states at the Fermi level (the Dirac point).This is similar to the edge states in graphene ribbons[144147], but the shell is isolated from the other statesby a gap. Indeed, the zigzag edge breaks the symmetrybetween the two sublattices of the honeycomb lattice, be-having like a defect. Therefore, electronic states localizedon the zigzag edges appear with energy in the vicinity ofthe Fermi level. The degeneracy Nedge is proportional tothe edge size and can be made macroscopic. A non-trivialquestion addressed here is the specic spin and orbitalconguration of the electrons as a function of the sizeand the fractional lling of the degenerate shell of edgestates. Due to the strong degeneracy, many-body eectscan be expected to be as important as in the fractionalquantum Hall eect, but without the need for a magneticeld. Calculations based on the Hubbard approximation[134, 135] and local spin density functional theory [135,136] showed that the neutral system (i.e., at half-lling)has its edge states polarized.

    In order to study many-body eects within thecharged degenerate shell using the conguration inter-action method, we rst perform a HartreeFock calcula-tion for the charged system of N Nedge electrons, withempty degenerate shell and Nedge electrons transferredto the gate, as shown in Fig. 20 (a). The spectrum of HFquasi-particles is shown in Fig. 20 (b) with black lines.Due to the mean-eld interaction with the valence elec-trons and charged gate, a group of three states is nowseparated from the rest by a small gap of 0.2 eV. Thethree states correspond to HF quasiparticles localized inthe three corners of the triangle. The same physics oc-curs in density functional calculation within local densityapproximation (LDA), shown in the inset of Fig. 20 (b).Hence we see that the shell of almost degenerate stateswith a well dened gap separating them from the valenceand conduction bands exists in the three approaches.

    The wave functions corresponding to the shell ofnearly-degenerate zero-energy states obtained from TB-HF calculations are used as a basis set in our congura-tion interaction calculations where we add Nadd electronsfrom the gate to the shell of degenerate states. In Fig.21, total spin S of the ground state as a function ofthe lling of the degenerate shell is shown for dierentsizes of quantum dots. Three aspects of these resultsare particularly interesting: (i) for the charge neutralcase (Nadd Nedge = 0), for all the island sizes studied(Nedge = 3 7), the half-lled shell is maximally spinpolarized as indicated by red arrows, in agreement withDFT calculations [135, 136]. The polarization of the half-lled shell is also consistent with the Lieb theorem forthe Hubbard model for bipartite lattice [160]. (ii) Thespin polarization is fragile away from half-lling. If weadd one extra electron (NaddNedge = 1), magnetization

  • 348 Wei-dong Sheng, et al., Front. Phys., 2012, 7(3)

    Fig. 20 (a) Electronic density in a triangular graphene islandof 97 carbon atoms where 7 electrons were moved to the metal-lic gate at a distance of dgate. (b) S