Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

Embed Size (px)

Citation preview

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    1/136

    Tel Aviv UniversitySchool of Chemistry

    Subject:

    Developing methods for transportcalculations in heterojunctions with

    Coulomb interactions

    Thesis submitted for the degree ofDoctor of Philosophy

    by

    Tal J. Levy

    Submitted to the Senate of Tel-Aviv UniversityJuly 2014

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    2/136

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    3/136

    This work was carried out under the supervision of

    Professor Rabani Eran

    School of Chemistry, Tel-Aviv University

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    4/136

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    5/136

    Dedications and Acknowledgements

    This thesis is dedicated to my parents, who throughout the years have given me much more

    than I can ever express here; to my brothers, Gad and Ran, who always find the time for abeer and a laugh when I am down, and to my mischievous cats, who always remind me that

    there is an easier way.

    I also dedicate this thesis to my beautiful partner and companion through life, Ariel, who

    recently gave me the best present of all - my daughter Phoenix.

    Last but not least, I dedicate this thesis to my advisor, Prof. Rabani Eran, whose only

    concern was my well-being (personally and scientifically). I really do mean it when I say that

    without your help and guidance, all of this would have never happened. Thank you.

    Finally, I would like to thank my past and present mentors, collaborators, group members,

    teachers and friends. Every little thing you have said or done contributed to my success.

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    6/136

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    7/136

    Abstract

    The demand for smaller, faster, and lower cost electronics has sent the industry and the scientific

    community looking for different paradigms. It almost naturally led to the question of whether

    it is possible to use single molecules and quantum dots as active elements in nanocircuits fora variety of applications. Developments in nanofabrication techniques have made possible the

    old dream of contacting individual molecules and quantum dots to macroscopic electrodes and

    explore their electronic transport properties. Moreover, it has been shown that molecules can

    indeed mimic the behavior of some of the more commonly used microelectronic components

    e.g., diodes, switches, and transistors. These achievements have given rise to what is now known

    as Molecular Electronics.

    There are still many experimental issues to overcome before molecular electronics will turn

    from an idea into a realizable technology, regardless, the exploration of molecular-scale circuits

    has already led to the discovery of many fundamental effects. In this sense, molecular elec-

    tronics has become a new interdisciplinary field of science, in which knowledge from traditional

    disciplines like physics, chemistry and engineering is combined to understand the electrical

    and thermal conduction at the molecular scale. Nevertheless, theory still faces several impor-

    tant challenges as the description of transport via an interacting nanoscaled region is a grave

    problem.

    Generally speaking, transport is a many-body, nonequilibrium phenomena, and its exact

    treatment requires a formalism explicitly designed to work out-of-equilibrium. While the equi-

    librium properties of systems similar to those to be treated in this thesis are quite well un-derstood and can be accurately solved for, the development of a general approach suitable for

    the treatment of fully nonequilibrium, many-body systems still remains a formidable task. An

    exact theoretical treatment of such systems is rather sparse to this day and includes only a

    small class of over-simplified model problems.

    We gather that in order to exploit the limitations and potential of the aforementioned

    novel electronic it is compulsory to carry out a comprehensive theoretical study of the physical

    properties that govern these systems. To this end, improvement of existing computational tools

    and formalisms, and the development of new ones are mandatory. The limitations of todaysformally/numerically exact solutions are often too severe for them to be useful in practice

    (usually, exponential scaling of computational resources and/or computational time with system

    size), thus, most theoretical treatments of quantum transport rely on approximations of some

    sort.

    In this thesis we study, analyze, and develop two methods for transport calculations in

    nanoscaled heterojunctions, described by model Hamiltonians (impurity models), with electron-

    electron interactions. The methods are: (a) the equations-of-motion technique for the nonequi-

    librium Green function and (b) a semiclassical approach.

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    8/136

    The nonequilibrium Green function formalism is considered as one of the pillars of current

    approaches to quantum transport which have been implemented in both model Hamiltonian

    formulations and first-principle methodologies. This approach is able to deal with a broad

    variety of physical problems related to quantum transport at the nanoscale. It can deal with

    strong non-equilibrium situations and it can also include interaction effects (e.g., electron-electron, electron-phonon) in a systematic way (diagrammatic perturbation theory).

    In this work we elaborate on the equations-of-motion technique which is one of the more basic

    and more powerful methods to calculate the Green functions of interacting quantum systems.

    In spite of its simplicity, it gives the appropriate results for strongly correlated nanosystems,

    describing qualitatively and in some cases quantitatively important transport phenomena, such

    as, Coulomb blockade and the Kondo effect in quantum dots.

    Be that as it may, we prove that Green functions calculated using the latter method may

    break the symmetries and relations that the correlation functions should fulfill by definition.Consecutively, the various expectation values calculated with these objects may turn to be

    unphysical. Thereafter, we suggest a strategy to restore the symmetries lost in the process

    of deriving the equations-of-motion and advise to add this step as an essential part of the

    equations-of-motion technique.

    Illustrations are then provided for two impurity models: the Anderson model (which serves

    as the simplest model to incorporates electron-electron interactions) and the double Anderson

    impurity model (an extended Hubbard model which includes inter-site Coulomb repulsion on

    top of the on-site electron-electron interactions). In addition, we develop two closures for the

    equations-of-motion obtained for the double Anderson model based on physical justifications

    in the regime studied, that outperform the commonly used closures found in the literature.

    On a completely different note, we explore and develop a semiclassical approach, which al-

    lows for the study of nonequilibrium quantum transport in molecular junctions using Hamiltons

    classical equations-of-motion, where the number of equations scales linearly with the number

    of the degrees-of-freedom. The dynamics of Fermions has always provided a challenge for semi-

    classical methods, since Fermions have a particularly notable non-classical behavior, which is

    the exchange antisymmetry of Fermions (which can be stated as the Pauli exclusion principle).

    The key idea behind the three methods presented here is to transform a general second quan-tized many-electron Hamiltonian into a classical one by defining a prescription which maps the

    electronic operators into classical functions that correctly account for the anti-commutativity

    of the Fermion operators. The three different mappings depicted in this work are just a nat-

    ural evolution of the same basic idea, where each transformation addresses the flaws of its

    predecessor.

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    9/136

    Contents

    1 Introduction 11

    1.1 Molecular Electronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

    1.2 Theoretical treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

    1.3 The model Hamiltonians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

    1.3.1 The single resonant level model . . . . . . . . . . . . . . . . . . . . . . . 16

    1.3.2 The Anderson model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

    1.3.3 The Double Anderson model . . . . . . . . . . . . . . . . . . . . . . . . . 17

    1.4 Thesis outline and goal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    2 Equations-of-Motion technique for the Nonequilibrium Green functions in

    Quantum Transport 19

    2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

    2.2 Nonequilibrium Green Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . 202.2.1 Dyson equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

    2.2.2 Analytic continuation and Langreth rules . . . . . . . . . . . . . . . . . . 28

    2.2.3 Equations-of-motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    2.2.3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    2.2.3.2 The equations-of-motion method . . . . . . . . . . . . . . . . . 31

    2.2.3.3 The single resonant level model - EOM treatment . . . . . . . . 33

    2.2.4 NEGF symmetry breaking within the EOM technique . . . . . . . . . . . 36

    2.2.4.1 The Anderson model . . . . . . . . . . . . . . . . . . . . . . . . 36

    2.2.4.2 The double Anderson model . . . . . . . . . . . . . . . . . . . . 40

    2.2.5 NEGF symmetry restoration . . . . . . . . . . . . . . . . . . . . . . . . . 46

    2.2.5.1 The Anderson model: symmetric-EOM approach . . . . . . . . 48

    2.2.5.2 The double Anderson model: symmetric-EOM approach . . . . 49

    2.3 Steady state conductance in a double quantum dot array:

    Assessing the symmetric-EOM technique for the NEGF . . . . . . . . . . . . . . 52

    2.3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

    2.3.2 Different closures of the EOM . . . . . . . . . . . . . . . . . . . . . . . . 53

    9

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    10/136

    2.3.2.1 Approximation1 . . . . . . . . . . . . . . . . . . . . . . . . . . 55

    2.3.2.2 Approximation2 . . . . . . . . . . . . . . . . . . . . . . . . . . 56

    2.3.2.3 Approximation3 . . . . . . . . . . . . . . . . . . . . . . . . . . 56

    2.3.2.4 Approximation4 . . . . . . . . . . . . . . . . . . . . . . . . . . 57

    2.3.3 Master equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 572.3.4 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

    2.3.4.1 Symmetric bridge . . . . . . . . . . . . . . . . . . . . . . . . . . 59

    2.3.4.2 Asymmetric bridge . . . . . . . . . . . . . . . . . . . . . . . . . 66

    2.4 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

    3 Semiclassical approaches to quantum transport 73

    3.1 overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

    3.1.1 Action-angle mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

    3.1.2 Cartesian mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

    3.2 The Hubbard mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

    3.2.1 Hubbard Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

    3.2.2 Reformulating the Hamiltonian in terms of Hubbard operators . . . . . . 80

    3.2.3 Classical mapping for the Hubbard operators . . . . . . . . . . . . . . . . 81

    3.3 Quantum transport in impurity models:

    Assessing the Hubbard mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

    3.3.1 Resonant level model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

    3.3.2 Anderson impurity model . . . . . . . . . . . . . . . . . . . . . . . . . . 853.3.3 Double Anderson model . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

    3.4 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

    4 Summary and Outlook 95

    A Full derivation of the broken symmetry in the Anderson model 97

    B Brief summary of the derivation of the NEGFs EOM: the double Anderson

    model 103

    C Calculating the poles of the retarded NEGF 109

    D Complimentary information for the Hubbard mapping 113

    Bibliography 117

    10

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    11/136

    Chapter 1

    Introduction

    1.1 Molecular ElectronicsThe demand for smaller, faster, and lower cost electronics, is by no means new. Already back

    in 1956 Arthur von Hippel formulated the basis of the bottom-up approach that he called

    molecular engineering [1]:

    Instead of taking prefabricated materials and trying to devise engineering applications con-

    sistent with their macroscopic properties, one builds materials from their atoms and molecules

    for the purpose at hand ...

    Of course he was not alone, in 1959the physicist Richard Feynman discussed the possibility

    of devices of extremely small dimensions in his lecture entitled Theres plenty of room at thebottom [2]:

    I dont know how to do this on a small scale in a practical way, but I do know that

    computing machines are very large; they fill rooms. Why cant we make them very small, make

    them of little wires, little elements and by little I mean little. For instance, the wires should

    be 10 or 100 atoms in diameter, and the circuits should be a few thousand angstroms across. [.

    . .] There is plenty of room to make them smaller. There is nothing that I can see in the laws

    of physics that says the computer elements cannot be made enormously smaller than they are

    now.The concept of molecular engineering introduced by Von Hippel led to the first notion of

    molecular electronics. It eventually resulted in a collaboration between Westinghouse Electric

    and the U.S. Air Force at the end of the 1950s. At that time, the U.S. Air force was open to

    new ideas and alternatives to the recently introduced integrated circuits and thus, organized

    a conference on molecular electronics which included scientists and engineers from military

    and private research labs. It was there that colonel C.H. Lewis, director of Electronics at the

    Air Research and Development Command, coined the term Molecular Electronics:

    Instead of taking known materials which wil l perform explicit electronic functions, and

    11

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    12/136

    reducing them in size, we should build materials which due to their inherent molecular struc-

    ture will exhibit certain electronic property phenomena. We should synthesize, that is, tailor

    materials with predetermined electronic characteristic. Once we can correlate electronic prop-

    erty phenomena with the chemical, physical, structural, and molecular properties of matter, we

    should be able to tailor materials with predetermined characteristics. We could design and create

    materials to perform desired functions. Inherent dependability might eventual ly result. We call

    this more exact process of constructing materials with predetermined electrical characteristics

    Molecular Electronics.

    However, molecular electronics, as we understand it today, started at the end of the1960s

    and the beginning of1970s. Thereupon, different groups started to investigate, experimentally,

    electron transport through molecular mono-layers. For instance, Hans Kuhn and coworkers

    studied new ways of fabricating mono-layers of organic materials (LangmuirBlodgett films) [3],

    which they were able to sandwich between metal electrodes and measure the electrical con-ductivity of the resulting junctions. A few years later (1974), Arieh Aviram and Mark Ratner

    published a now-famous paper on molecular rectifiers [4] in which they described how a mod-

    ified charge-transfer salt could operate as a traditional diode in an electrical circuit. The idea

    was considered for a long time a theoretical curiosity that could not be tested experimentally,

    and in this sense, it did not have much impact in the scientific community back then.

    Much has changed with the invention of the scanning tunneling microscope (STM) [5, 6]

    in 1981, the introduction of mechanically controllable break junction (MCBJ) technique [7]

    in 1985, and the development of controllable single molecule junctions [8, 9, 10, 11, 12, 13,

    14, 15, 16, 17, 18, 19, 20, 21]. The STM was the first tool that provided a practical way

    to see and manipulate matter at the atomic scale and very quickly it became clear that it

    could provide a realistic way to address single molecules and to study their electronic transport

    properties [22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35].

    With these techniques a large amount of interesting phenomena have been observed. For

    example: rectification [17], negative differential conductance [9, 33, 36], Coulomb blockade [22,

    10, 11, 14, 15, 20], Kondo effect [11, 37], vibrational effects [24, 29, 30, 32, 34], and nanoscale

    memory effects [38, 39, 40] just to name a few.

    Many engineering challenges, such as robustness and stability of these molecular devices,still remain to be addressed before molecular electronics can be practical, however, these

    results clearly show the potential of molecule-based electronics. Mechanical challenges are not

    the only thing that needs to be attended to. In that regard the theoretical treatment of current

    and heat flowing through a nanoscaled junction is far from complete.

    12

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    13/136

    1.2 Theoretical treatment

    The growing interest in the properties of both mesoscopic and nanometric devices and the new

    physics and phenomena encountered in the various labs has raised fundamental and conceptual

    questions. Theory faces several important challenges and the development of a general approachsuitable for the treatment of fully nonequilibrium many-body systems still remains a formidable

    task. In this regard the main challenges facing theory are:

    Understanding the coupling of individual molecules to macroscopic electrodes under

    nonequilibrium conditions [41, 42, 43, 44].

    Characterization of the temperature dependence on conductance as well as the role of

    molecular vibrations are of crucial importance. This would facilitate understanding ofelastic and inelastic transport phenomena, where the nature of both coherent and inco-

    herent transport is paramount for a complete picture [45, 46, 47, 48].

    Electron-electron correlations in the macroscopic electrodes can often be neglected by the

    application of an effective band-structure description of the leads, such correlations, on the

    other hand, are important within the molecule and thus strongly affect the conduction [49,

    50].

    Electron-hole excitations (excitons [51]) formed in the heterojunction, can remarkably

    affect the current through it. While in case of noninteracting electrons the electron-hole

    interaction leads to reduction in the current (exciton blocking), in the case of strong

    Coulomb repulsion, conduction exists even when the electronic connectivity does not

    exist [52]. Also, electronhole pair creation processes can explain vibrational excitations

    characteristic of a single-molecule contact [53].

    Characterizing transport behavior in systems driven by weak and strong electromagnetic

    fields and the optical properties of such junctions is another difficult issue which must be

    overcome by theory [54, 55, 56, 57].

    Although noteworthy advances have been made, there is still a discrepancy between experimen-

    tally measured and theoretically calculated values [42, 18, 58]. While recent work has been more

    encouraging regarding these discrepancies, the problem clearly remains unsolved. Thus, the de-

    velopment of a practical, general approach suitable for the treatment of fully nonequilibrium,

    many-body systems is desirable.

    13

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    14/136

    In general, many-body, interacting, out-of-equilibrium problems cannot be solved exactly

    but for a few simple cases [59, 60, 61, 62]. Excluding recent developments based on brute-

    force approaches such as time-dependent numerical renormalization-group techniques [63, 64,

    65, 66, 67, 68], iterative [69] or stochastic [70, 71, 72, 73, 74] diagrammatic techniques to

    real- time path integral formulations, wave function based approaches [75], or reduced dynamicapproaches [76, 77, 78, 79], all suitable to relatively simple model systems, most theoretical

    treatments of quantum transport rely on approximations of some sort.

    From a more formal standpoint, there are roughly two main theoretical frameworks that

    can be used to study quantum transport in nanosystems at finite voltage: the nonequilibrium

    Green function (NEGF) formalism [80, 81, 82] otherwise known as the Keldysh NEGF or the

    SchwingerKeldysh formalism [83, 84], and generalized master equations methods [85, 86].

    In this work we also introduce an emerging paradigm suitable to tackle quantum transport

    at the nanoscale; a semiclassical approach [62, 87, 88, 89, 90], where the quantum Hamiltonian,and the quantum operators are mapped onto classical functions and degrees-of-freedom that

    follow Hamiltons equations of motion.

    For the sake of completeness we mention two more emerging and improving theoretical

    frameworks applicable to open nonequilibrium systems: (a) methods based on density functional

    theory (DFT) [91, 92, 93, 94, 95, 96, 97, 98], and (b) time dependent multiconfigurational

    methods [99, 100, 101].

    In the next section we introduce the model Hamiltonians used throughout this thesis.

    1.3 The model Hamiltonians

    Up till 1970, in most treatments of electron tunneling through a metal-insulator-metal (MIM)

    junction it was accustomed to transform an initial Hamiltonian (of the total system) into an

    effective Hamiltonian of the form:

    Heff= HL +HR +T , (1.3.1)

    where HL and HR represents the metals in the (MIM) junction (as accurately as possible),

    i.e., the left and the right electrodes, while the transfer term Tdescribes the probability of

    an electron to tunnel through the junction. Although, the tunneling is energy dependent,

    this dependency was neglected from the transfer term, thus, making this theory unsuitable to

    describe many-body effects, such as electron-electron and electron-phonon interactions. The

    main reason for using effective Hamiltonians was the belief that a description of a system as a

    whole will not allow for the calculation of the current, which is a nonequilibrium process.

    In1970Caroliet alproposed the prototype of model Hamiltonians [102, 103, 104, 105] used

    14

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    15/136

    ever since in the context of transport. The Hamiltonian they proposed was designed to permit

    direct calculation of the tunneling current through a MIM junction. This was achieved by

    describing the system in terms of localized atomic functions instead of quasi-free electron wave

    functions, which resulted naturally in a three-parts Hamiltonian describing the electrodes, the

    system, and the interaction between these parts. The general form of the Hamiltonian is givenby:

    H = Hsys +Hbath +Hint, (1.3.2)

    where Hbath describes the macroscopic leads (left and right electrodes), Hsys describes the

    system of interest, and Hint is the interaction Hamiltonian between the system and the leads.

    The contacts (leads) are usually modeled as infinite noninteracting Fermion baths [106, 107,

    108], they are assumed to be each at their own equilibrium characterized by chemical potentials

    L and R, where the difference = LRe is the applied voltage bias across the junction(e is the absolute electronic charge of an electron), and are described by a grand canonical

    distribution. The leads Hamiltonian in second quantization is given by:

    Hbath=

    K{L,R}

    ,kK

    kckck , (1.3.3)

    where k is the energy of a free electron in the left (L) or right (R) lead, in momentum state

    kand spin , with the operators ck and ck being annihilation and creation operators of such

    an electron. The density matrix of the left/right lead at equilibrium is given by:

    L/R= e

    ,kL/R kckckL/RNL/R

    Tr

    e

    ,kL/R kckckL/RNL/R

    , (1.3.4)

    where= (KBT)1 is the reciprocal of the temperatureTand the Boltzmann constantKB, the

    trace (Tr [ ])should be taken as the sum over a complete basis and NL/R=

    ,kL/R ckck

    is the occupation operator of the left/right lead.

    The interaction between the system and the contacts is simply given by the tight-binding

    Hamiltonian [102]:

    Hint=

    K{L,R}

    ,kK

    m

    tkmckdm+ h.c.. (1.3.5)

    The transition matrix element tkm represents the coupling strength between the system (site

    m, spin ) and the leads (electron in state k and spin ). It depends on the specific contact

    geometry between the bridge and the leads. The corresponding level width functions are given

    15

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    16/136

    by:

    L/Rnm () = 2

    ,kL/R(tkn)

    tkm( k) . (1.3.6)

    These functions describe the broadening of the energy levels on the molecular bridge due to

    this coupling.The form chosen for Hsys depends on the system studied. In the following subsections we

    describe the different system Hamiltonians in detail.

    1.3.1 The single resonant level model

    This simple model was introduced by Ugo Fano [109] in 1961when he worked on the effect of

    auto-ionization on excitation spectra. In this model the system has one spin-less level:

    Hsys = n. (1.3.7)

    Here n= dd is the number operator of the electron occupying the level with free-energy .Since the system has only one level, and no spin, we can simplify the expressions of the bath

    and interaction Hamiltonians:

    Hbath=

    K{L,R}

    kK

    kckck, (1.3.8)

    Hint=

    K{L,R}

    kK

    tkckd + h.c.. (1.3.9)

    1.3.2 The Anderson model

    The same year Fano introduced his single level Hamiltonian, Philip W. Anderson was seeking

    for conditions required for the stability of localized magnetic moments like those of iron, cobalt

    or nickel dissolved in non-magnetic metals [110]. In the model he proposed the system isrepresented by one electronic level that can accommodate up to 2 interacting electrons (spin

    up and spin down electrons) [111, 112]. The systems Hamiltonian for the Anderson model is

    given by:

    Hsys=

    {,}n+ U nn. (1.3.10)

    Here n =dd is the number operator of the spin electron with energy and U is therepulsion energy between two electrons on the same site with opposite spins (on-site repulsion).

    16

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    17/136

    The interaction Hamiltonian takes the form:

    Hint=

    K{L,R}

    ,kK

    tkckd+ h.c.. (1.3.11)

    1.3.3 The Double Anderson model

    This model is an extension of the Anderson model to systems that include more than just

    one electronic level. Here we have 2 electronic levels each of which can accommodate up to

    2electrons. The model incorporates on-site and inter-site Coulomb repulsions and a hopping

    term that allows electrons to hop from one site to the other. The Hamiltonian in second

    quantization is given by [113, 114]:

    Hsys =

    ,m{,}mnm+

    m

    Umnmnm

    +,

    V

    nn +

    hd

    d+ h.c.

    . (1.3.12)

    The first two terms on the R.H.S are similar to the Anderson impurity model Hamiltonian

    (extended to 2sites,and ),V

    is the repulsion energy between two electrons on different

    sites (inter-site repulsion), and his the coupling strength for electron hopping between the

    two sites. In the calculations performed and results presented throughout this thesis the twosites are assumed to be in a serial configuration, thus, the interaction Hamiltonian takes the

    form:

    Hint=

    ,kLtkc

    kd+

    ,kR

    tkckd+ h.c.. (1.3.13)

    That is, site is connected to the left lead while site is attached to the right lead.

    1.4 Thesis outline and goal

    The primary objective of the research program presented here has been to analyze and develop

    existing and novel theoretical and computational methods tailored to investigate charge (and

    spin) transport at nanometer length scales. The theoretical examination of transport at the

    nanoscale requires the use of advanced and powerful techniques able to deal with the dynamical

    properties of the relevant physical systems, to explicitly include out-of-equilibrium situations

    typical for electronic transport as well as to take into account interaction effects. In particular,

    we have been interested in techniques suitable for large electronic systems, i.e. low scaling tech-

    17

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    18/136

    niques that describe the system at hand beyond a mean-field approximation. In that regard we

    focused our attention on two methods for transport calculations in nanoscaled heterojunctions,

    described by model Hamiltonians (impurity models), with electron-electron interactions:

    1. The equations-of-motion technique for the nonequilibrium Green function.

    2. A semiclassical approach.

    The nonequilibrium Green functions formalism is introduced in chapter 2 where we elaborate on

    the equations-of-motion technique. We prove that Green functions calculated using the latter

    may at different levels of closures and decoupling schemes break the symmetries and relations

    that the correlation function should fulfill by definition, consecutively, the various expectationvalues calculated with these objects may turn to be unphysical.

    Thereafter, we suggest a strategy to restore the symmetries lost in the process of deriving

    the equations-of-motion and advise to add this step as a mandatory part of the aforementioned

    technique. Illustrations are then provided for two impurity models: the Anderson model (which

    serves as the simplest model to incorporates electron-electron interactions) and the double

    Anderson impurity model (an extended Hubbard model which includes inter-site Coulomb

    repulsion on top of the on-site electron-electron interactions). In addition, we develop two

    closures for the equations-of-motion obtained for the nonequilibrium Green function of the

    double Anderson model based on physical justification in the regime studied, that outperform

    the commonly used closures found in the literature.

    In chapter 3, we present three mappings of a semiclassical approach developed in collabora-

    tion with the Miller group at the University of California, Berkeley U.S.A. These methods allow

    for the study of nonequilibrium quantum transport in molecular junctions using Hamiltons

    classical equations-of-motion, where the number of equations scale linearly with the number

    of degrees-of-freedom. The key idea behind these methods is to transform a general second

    quantized many-electron Hamiltonian into a classical one. This is achieved by defining a pre-

    scription which maps the electronic operators into classical functions that correctly accountfor the anti-commutativity of the Fermion operators. The three different mappings are just a

    natural evolution of the same basic idea, where each transformation addresses the flaws of

    its predecessor.

    Out of the three mappings we center on the most recent Hubbard mapping and asses its

    validity by employing the method on the three model Hamiltonians described in the previous

    section.

    18

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    19/136

    Chapter 2

    Equations-of-Motion technique for the

    Nonequilibrium Green functions in

    Quantum Transport

    2.1 Overview

    In this chapter we outline the theory of the nonequilibrium Green function (NEGF) formalism,

    which is widely used to describe transport phenomena in nanojunctions. As will be clear shortly,

    the NEGF method is able to deal with a very broad spectrum of physical problems related to

    transport at the nanoscale [80, 82, 59, 115, 116, 117]. For one, as it name implies, it is tailoredto take on out-of-equilibrium situations, and interaction effects can be dealt with in a well

    defined manner (diagrammatic perturbation theory [118], functional derivatives technique [80,

    119]). The NEGF formalism, initiated by Schwinger [83], Kadanoff and Baym [80] and later

    Keldysh [84] allows one to:

    Study the time evolution of a many-particle quantum system.

    Calculate time-independent and time-dependent expectation values such as currents, den-sities, electron addition and removal energies and the total energy of the system.

    Calculate the spectral function which gives access to the local density of states.

    Describe dissipative processes and memory effects in transport that occur due to electron-

    electron interactions and coupling of electronic to nuclear vibrations.

    In a series of papers [102, 103, 105, 104] Caroli et al presented a general formalism for the

    calculation of the current through an interacting system. An exact expression for the current

    19

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    20/136

    was later reformulated by Meir and Wingreen [120] in terms of the systems NEGF and the

    self-energies which represent the effects of the external baths on the system.

    The NEGF formalism is widely used nowadays to describe electron and hole transport in a

    large variety of devices and materials, such as: III Vresonant tunnel diodes [121, 122, 123,124, 125, 126], electron waveguides (i.e., electrons in 2D) [127], Silicon tunnel diodes [128, 129],carbon nanotubes [130, 131, 132, 133], metal wires [92, 134], organic molecules [135, 136, 137,

    138, 139, 140, 141, 142], and magnetic leads [143, 144].

    Physics that have been included are open-system boundaries [102], full-band structure [145,

    146, 128, 129], the self-consistent Hartree potential [123, 147], exchange-correlation potentials

    within a density functional theory (DFT) approach [148, 138, 149, 92], photon absorption and

    emission [150, 151, 54, 152], energy and heat transport [153, 45], acoustic, optical, intra-valley,

    inter-valley, and inter-band phonon scattering [125, 126, 154, 128, 129, 151], single-electron

    charging and nonequilibrium Kondo systems [155, 156, 157, 158, 112], shot noise [159, 160, 161],alternate Current [121, 162, 163, 164], and transient response [165].

    Excluding simple noninteracting cases, the calculation of the systems NEGF required to

    obtain the current (or any other observable) is far from trivial. Most applications are based

    on a perturbation expansion (and diagrammatic techniques) to obtain the systems NEGF.

    Alternatively, one can use the equations-of-motion (EOM) approach, which allows to deduce

    the systems NEGF by deriving the corresponding equations-of-motion.

    This chapter continues with a brief description of the NEGF [117]. We shall focus on the

    equations-of-motion technique, outline its advantages and disadvantages, and elaborate on the

    single-particle NEGF symmetry breaking [166] due to the way this technique is carried out.

    We show that this symmetry/relation breaking can lead to solutions which are not physical

    and suggest a scheme to restore it [166]. Finally, we formulate a couple of closures, one of

    which outperforms the commonly used closures found in the literature, for the EOM of the

    NEGF for the double Anderson model and use the symmetric-EOM technique also developed

    by us to calculate the current and differential conductance through the double quantum-dot

    array [114, 167] (the double Anderson model).

    2.2 Nonequilibrium Green Functions

    In nonequilibrium theory one usually divides the full Hamiltonian into three parts:

    H (t) = H0 +V H

    +Hex (t) = H+Hex (t) , (2.2.1)

    20

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    21/136

    where H0 is the single-particle, noninteracting part, Vcontains all the interactions (electron-

    electron, electron-phonon, impurity scattering, etc.) and Hex (t) contains all external pertur-

    bations (which may be time-dependent) driving the system out of equilibrium.

    A standard device is to assume that the external perturbations are turned on at a certain

    time t0, while prior to that time, the system is at equilibrium, i.e., Hex (t) = 0 for t < t0,and is described by the thermal equilibrium density operator which is a function of the time-

    independent interacting Hamiltonian H:

    = 1

    ZeH, Z=Tr

    eH

    . (2.2.2)

    Before introducing the single-particle, nonequilibrium Green function, which is a correlation

    function between two Fermion/Boson field operators at different locations and times, we will

    formulate the nonequilibrium problem. The task at hand, is the calculation of the expectation

    value of an observable at time t > t0(associated to the quantum mechanical operator O):

    O (t)

    =Tr

    OH (t)

    , (2.2.3)

    where the subscript H indicates that the time dependence of the Heisenberg operator is governedby the full Hamiltonian (equation (2.2.1)) and is given in equation (2.2.2).

    The general plan of attack is to transform the extremely complicated time dependence of

    the operator O (t)to a much simpler form, specifically, have its time evolution governed by the

    noninteracting Hamiltonian H0. Furthermore, to allow a perturbation expansion and the use ofWicks theorem [168, 169] or the Feynman diagrams [169, 170, 118], which are just a graphical

    way representing the results of Wicks decomposition, one also needs to transform the density

    operator such that it becomes a single-particle density operator. This is usually achieved by

    two transformations [115]. The first transformation is due to the following identity:

    OH (t) = Sex (t0, t)OH(t)Sex (t, t0) , (2.2.4)

    with

    OH(t) = e ihHt Oe ih Ht , (2.2.5)

    Sex (t, t0) =T

    exp

    i

    h

    tt0

    dHexH ()

    , (2.2.6)

    and

    HexH (t) = eihHtHex (t) e

    ih Ht . (2.2.7)

    The time-order operator T{}in equation (2.2.6) equals either the chronological time-orderoperator, Tf{}, which moves the operator with the later time to the left, if t > t0, or the

    21

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    22/136

    anti-chronological time-order operator, Tb {}, which moves the operator with the later timeto the right.

    At this point the time-contour (proposed by Schwinger [83] and later by Keldysh [84] and

    Craig [171]) and contour-ordered quantities [82] are introduced. By joining the two exponentials

    of equation (2.2.4) we get:

    OH (t) = Sex (t0, t)OH(t)Sex (t, t0) ,

    = Tb

    exp

    i

    h

    t0t

    dHexH ()

    OH(t) Tf

    exp

    i

    h

    tt0

    dHexH ()

    ,

    = TC

    exp

    i

    h

    t0t0

    dHexH()

    OH(t) ,

    = TC

    exp

    i

    h

    C

    dHexH ()

    OH(t) . (2.2.8)

    The integral in the exponent is now a contour integral along the contour C(depicted in fig-

    ure 2.2.1) and TC{}is the contour time-order operator, which moves operators with timesfurther along the time-contour Cto the left. By defining:

    SexC = TC

    exp

    i

    h

    C

    dHexH ()

    , (2.2.9)

    equation (2.2.3) can be rewritten as:

    O (t) =TrOH (t)

    =Tr

    TC

    SexCOH(t)

    . (2.2.10)

    In order to define the fundamental object of nonequilibrium many-body theory, i.e. the

    NEGF, one extends the notion of the time variable (t) to contour time variables (). We

    Figure 2.2.1: The time-contourC, starts and ends at t0with contour variable1on theforward/upper branch(C+)and 2 on thebackward/lower branch(C). By definition, any point lying on the backward branch comesafter a point lying on the forward branch, meaning us, TC

    A(1)B (2)

    = B (2)A(1). The minus sign

    originates from the permutation of two Fermion operators.

    not in passing that the extension is not unique as a time t can be mapped into a contour

    time variable on the forward or the backward branches. Finally, the contour-ordered Green

    22

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    23/136

    function is given by:

    G(r2, 2,r1, 1) = ih

    TC

    H (r2, 2)

    H (r1, 1)

    , (2.2.11)

    where (r, ) / (r, ) is the annihilation/creation field operator which annihilates/createsa particle at position r and time . Henceforth, our discussion will only involve Fermion field

    operators.

    After applying the transformation described we get:

    G(r2, 2,r1, 1) = i

    h

    TC

    SexC

    H(r2, 2)H(r1, 1)

    TC

    SexC

    ,

    =

    i

    h

    Tr

    TC

    SexC

    H(r2, 2)H(r1, 1)

    Tr

    TCSexC . (2.2.12)The time-contour Cmay be elongated so that it would run beyond the latest time (t1 or t2);

    usually the contour is elongated to t . In this form the contour-ordered Green function (GF)does not admit Wicks decomposition [172] and the use of Feynman diagrams, since the field

    operators and the density operator depend on the (still complicated) interacting Hamiltonian

    Hand not on the noninteracting Hamiltonian H0. To circumvent this problem one more

    transformation is performed. The full derivation can be found in Ref. [82], but basically one

    uses the identity:

    =Z0

    Z0S

    v (i, 0) , (2.2.13)

    0= 1

    Z0eH0, Z0=Tr

    eH0

    , (2.2.14)

    Sv (i, 0) =Ti

    exp

    i

    h

    i0

    dtVH0 (t)

    , (2.2.15)

    with

    VH0 (t) = ei

    hH0tV e

    i

    hH0t. (2.2.16)

    The operator Ti is the time-order operator, ordering the operators along the imaginary-time

    segment [i, 0] (see figure 2.2.2 for more details) such that the time closest toiis to the

    23

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    24/136

    left. Putting it all together one gets:

    G(r2, 2,r1, 1) = i

    h

    Tr0TCi

    SexH0,C

    SvCiH0 (r2, 2)H0 (r1, 1)

    Tr

    0TCi

    SexH0,CSvCi

    ,

    = ih

    Tr0TCi SCiH0 (r2, 2)H0 (r1, 1)

    Tr0TCi

    SCi

    . (2.2.17)where

    H0 (r, ) =eihH0 (r) e

    ihH0,H0 (r, ) =e

    ihH0 (r) e

    ihH0,(2.2.18)

    the contour time-order TCi operator, orders the operators along the contour Ci, and

    SexH0,C= TCexp ih CdHexH0 () , (2.2.19)

    SvCi= TCi

    exp

    i

    h

    Ci

    dVH0 ()

    , (2.2.20)

    SCi=SexH0,C SvCi , (2.2.21)

    with

    HexH0 () = eih H0Hex (t) e

    ihH0. (2.2.22)

    The contour-ordered Green function in equation (2.2.17) accepts Wicks decomposition and

    the perturbation expansion is possible. As already mentioned at the beginning of section 2.2,

    Figure 2.2.2: The Kadanoff-Baym contour Ci, starts at t0 and ends at t0 i, with contour variable 1 onthe forward branch and2 on the backward branch. By definition any point on the vertical branch comes after

    the points on either of the horizontal branches, thus, TCi {A(1)C(3)B (2)} = C(3)B (2)A(1).

    all interactions are present in V. If one does not care about initial correlations (which is

    a conceivable assumption when studying steady-state transport for example), one can take

    24

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    25/136

    the limit t0 . With reference to the equilibrium theory, interactions are switched onadiabatically, such that at t0= the system is not interacting and H= H0. Under thisassumption = 0, therefore, the transformation in equation (2.2.13) is not called for and the

    addition of the imaginary segment to the time-contour C is not required [173]. The resulting

    time-contour is known as the Schwinger-Keldysh time-contour, which we will denote byC(seefigure 2.2.3). In this regard, the contoured-ordered Green function takes a somewhat simpler

    form:

    G(r2, 2,r1, 1) = ih

    TC

    H (r2, 2)

    H (r1, 1)

    , (2.2.23)

    which can be reformulated in the interaction representation:

    G(r2, 2,r1, 1) = ih

    Tr0TC

    SCH0 (r2, 2)

    H0 (r1, 1)

    Tr0TC SC , (2.2.24)

    with

    SC= SexC

    SvC, (2.2.25)

    SexC= TC

    exp

    i

    h

    C

    dHexH0 ()

    , (2.2.26)

    SvC= TC

    exp

    i

    h

    C

    dVH0 ()

    . (2.2.27)

    In this study we are interested in steady-state transport, therefore, in all that follows, we

    confine ourselves exclusively to the Keldysh approach [174], i.e., we neglect initial correlations

    and work on the Schwinger-Keldysh contour C. As already mentioned, by extending the time

    Figure 2.2.3: The Schwinger-Keldysh contour C. When initial correlations are neglected, one ignores the

    imaginary segment of the time-contour Ci and elongates the time-contour C from t0= to t = and back.

    variables into contour time variables, the uniqueness of these variables is lost. Such being the

    case one should keep track of which branch is in question. With two time variables that can

    25

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    26/136

    be located on either one of the two branches, four real-time Green functions are introduced:

    GT(r2, t2,r1, t1) = ih

    Tf

    H (r2, t2)H (r1, t1)

    1, 2 C+,

    G(r2, t2,r1, t1) =

    ih H (r2, t2)H (r1, t1) 1 C+, 2 C,

    GT(r2, t2,r1, t1) = ih

    Tb

    H (r2, t2)H (r1, t1)

    1, 2 C,

    (2.2.28)

    where GT is the time-ordered Green function,GT is the anti-time-ordered Green function, Gis the greater Green function. As opposed to its equilibrium

    counter-part, the single-particle NEGF is a 2 2tensor [175]:

    G(r2, 2,r1, 1) = GT(r2, t2,r1, t1) G

    (r2, t2,r1, t1) GT(r2, t2,r1, t1) . (2.2.29)

    Usually one defines two other real-time Green function, the retarded Green function

    GR

    and

    the advanced Green function

    GA

    :

    GR(r2, t2,r1, t1) = GT(r2, t2,r1, t1) G(r2, t2,r1, t1)

    = i

    h (t1 t2)

    H (r2, t2) ,

    H (r1, t1)

    , (2.2.31)

    where { , } is the anti-commutator. With these definitions, the single-particle NEGF tensorcan also be presented as follows [84]:

    G(r2, 2,r1, 1) =

    GR(r2, t2,r1, t1) G are all connected via the fluctuation-dissipation

    theorem [176] and only one Green function is needed to describe a system, this is not longer the

    case in nonequilibrium situations. Nevertheless, one easily observes that the real-time NEGFs

    26

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    27/136

    are not all independent and the following relations and symmetries hold by definition:

    GR(r2, t2,r1, t1) =

    GA (r1, t1,r2, t2)

    ,

    G(r2, t2,r1, t1) =

    G (r1, t1,r2, t2)

    ,

    GR(r2, t2,r1, t1) GA(r2, t2,r1, t1) = G>(r2, t2,r1, t1) G(r2,r1; ) G

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    28/136

    to interactions between the particle and the system it is part of, one obtains Dysons equation:

    Gr, , r,

    = G0

    r, , r,

    +

    d1d2G0 (r, ,2) (2,1) G

    1,r,

    ,

    = G0 r, , r, + d1d2G (r, ,2) (2,1) G

    01,r, ,(2.2.36)

    where 1 (r1, 1)and

    d1 dr1 C d1.Dysons equation can be easily expressed in matrix notation (see equations (2.2.32)):

    Gr, , r,

    = G0

    r, , r,

    +

    d1d2G0 (r, ,2) (2,1) G

    1,r,

    ,

    = G0 r, , r, + d1d2G (r, ,2) (2,1) G01,r

    , ,(2.2.37)

    only now the contour integrals are transformed into real-time integration, i.e.,

    d1 dr1 dt1.Of course, usually, one does not know the exact form of the self energy, and approximated ex-

    pressions are used to approximate the Green function.

    2.2.2 Analytic continuation and Langreth rules

    When writing down the EOM for the NEGF or when considering the different terms in the

    perturbation expansion we encounter contour quantities with the following structures:

    F1 (2, 1) =

    C

    A (2, ) B (, 1) d, (2.2.38)

    F2 (2, 1) =

    C

    A (2, ) B

    ,

    C

    , 1

    dd, (2.2.39)

    F3, = A, B , , (2.2.40)F4

    ,

    = A

    ,

    B

    ,

    , (2.2.41)

    where we suppressed all variables but the temporal ones for clarity and brevity. The contour

    integration and contour variable are not practical for calculations and one can replace the

    contour time variables and the contour integrals with real-time variables and integrals (analytic

    continuation [115]) using Langreth theorem [81, 115, 117]. The rules provided by Langreth

    theorem are summarized in table 2.1. For example, using Langreth rules and Dysons equation

    28

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    29/136

    for the NEGF, the equations for the retarded and lesser Green functions are given by:

    GRr, t,r, t

    = GR0

    r, t,r, t

    +

    d1d2GR0 (r, t,2)

    R (2,1) GR1,r, t

    ,

    = GR

    0 r, t,r, t+ d1d2GR

    (r, t,2) R

    (2,1) GR

    0 1,r, t ,(2.2.42)

    G 1 on the contour

    otherwise . (2.2.49)

    31

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    32/136

    The EOM [180] for Gij (2, 1)can be written as:

    ih

    2Gij (2, 1) = C (2 1)

    ai, a

    j

    +

    TC

    2ai (2) a

    j (1)

    ,

    = C (2 1)ai, aj+TC ih H (2) , ai (2)aj (1) ,= C (2 1) ij+

    TC

    i

    h

    H0, ai (2)

    aj (1)

    +

    TC

    i

    h

    V(2) , ai (2)

    aj (1)

    . (2.2.50)

    Since H0is the time-independent, noninteracting, quadratic Hamiltonian, the term

    H0, ai (2)

    equals p pap (2), whereps are parameters of the noninteracting Hamiltonian, such as thekinetic energy. Using this and the definition in equation (2.2.48) we can rewrite:

    TC

    i

    h

    H0, ai (2)

    aj (1)

    =p

    pGpj (2, 1) . (2.2.51)

    The term

    TC

    ih

    V(2) , ai (2)

    aj (1)

    results in higher order Green functions (depending

    on the form ofV(t), these can be 2-, 3- or even higher particle Green functions). We shall

    denote those as G(2, 1). Finally we can rewrite:

    ih

    2 i

    Gij (2, 1) = C (2 1) ij+p=i

    pGpj (2, 1)

    +p

    TpGp (2, 1) , (2.2.52)

    where once again Tps are parameters of the interaction Hamiltonian

    V(t)

    , such as the

    Coulomb repulsion energy. Define the Green functiongi (2, ):

    ih

    2 i

    gi (2, ) =C (2 ) , (2.2.53)

    and equation (2.2.52) now takes the form:

    Gij (2, 1) = gi (2 1) ij+p=i

    p

    C

    gi (2 ) Gpj (, 1)d

    +p

    Tp

    C

    gi (2 )Gp (, 1)d. (2.2.54)

    Depending on the Hamiltonian, the newly generated Green functions (Gpj and Gp) can

    involve lead (bath) operators as well as system operators. Except for very simple cases, where

    an exact closure can be obtained (see sub-subsection 2.2.3.3), writing the EOM for Gpj

    (2

    , 1

    )

    32

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    33/136

    and Gp (2, 1) will produce new and/or higher order Green functions that need to be eval-

    uated. This leads (in principle) to an infinite/ intractable set of equations. The idea of the

    EOM method is therefore, to truncate this hierarchy of equations making a mean-field like ap-

    proximation for the higher-order Green functions through lower order functions. This is the

    Achilles heel of this method as there is no systematic way to close the equations and no theoryto back-up such a procedure. Usually the approximations have physical meaning within the

    regime of the problem at hand [189, 190, 111, 167].

    As an illustration of the method we next calculate the NEGF of a single resonant level [109,

    62]described in subsection 1.3.1.

    2.2.3.3 The single resonant level model - EOM treatment

    Define the systems contour ordered Green function:

    G

    ,

    = ih

    TC

    d () d

    . (2.2.55)

    The equation of motion ofG (, )is:

    ih

    G

    ,

    = C

    +

    TC

    dd

    d () d

    = C

    i

    hTC d () d

    ih

    K{L,R}

    kK

    tk

    TC

    ck () d

    = C

    + G

    ,

    +

    K{L,R}

    kK

    tkFk

    ,

    , (2.2.56)

    whereFk (, ) = ih

    TC

    ck () d ()

    , is a new Green function generated in the procedure.

    To obtain G (, )we now derive the EOM ofFk (, ):

    ih Fk , = C ck, d+TCddck () d

    = ih

    k

    TC

    ck () d i

    htk

    TC

    d () d

    = kFk

    ,

    + tkG

    ,

    . (2.2.57)

    The last equation can be rearranged:

    ih

    k

    Fk

    ,

    = tkG

    ,

    . (2.2.58)

    33

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    34/136

    Define the leads noninteracting Green function through the following equation:

    ih

    k

    gk ( 1) = C ( 1) , (2.2.59)

    so equation (2.2.58) takes the form:

    Fk

    ,

    =

    C

    gk ( 1) tkG

    1, d1. (2.2.60)

    At this point the equations close. To finish our derivation, substitute equation (2.2.60) in

    equation (2.2.56):

    ih

    G

    ,

    = C

    +

    K{L,R} kKC

    |tk|2 gk ( 1) G

    1,

    d1.

    (2.2.61)

    Define the levels noninteracting Green function through the following equation:

    ih

    g ( 1) =C ( 1) , (2.2.62)

    so finally the equation for the systems NEGF is:

    G, = g + K{L,R}

    kK

    C

    g ( 2) |tk|2 gk (2 1) G1, d1d2.(2.2.63)

    The Green function G (, ) contains the information on the resonant level in an explicitway (via g ( )), while the effect of all other (infinitely many) levels of the left and rightleads appear only in the sum:

    K{L,R} kK|tk|2 gk (2 1) , (2.2.64)

    which we recognize as the self energy:

    (2 1) =

    K{L,R}

    kK

    |tk|2 gk (2 1) . (2.2.65)

    Applying Langreth rules one can obtain the retarded, advance, lesser and greater Green func-

    tions:

    34

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    35/136

    GR/A

    t, t

    = gR/A

    t t

    +

    gR/A (t t2) R/A (t2 t1) GR/A

    t1, tdt1dt2.

    (2.2.66)

    G

    t, t

    = g

    t t

    +

    gR (t t2) R (t2 t1) G

    t1, tdt1dt2

    +

    gR (t t2) (t2 t1) GA

    t1, tdt1dt2

    +

    g (t t2) A (t2 t1) GA

    t1, tdt1dt2. (2.2.67)

    In steady state the NEGF is a function of one time variable [115] (the time difference). In

    this case it is simpler to express the NEGFs in Fourier space. To simplify the notation we

    denote the Fourier transform ofG (t)as G (), i.e., functions with an argument are Fourier

    transforms of their time-domain counterparts. The resulting equations are:

    GR/A () =gR/A () + gR/A() R/A () GR/A () , (2.2.68)

    G () = g () + gR () R () G ()

    +gR

    ()

    () GA

    () + g

    () A

    () GA

    () , (2.2.69)

    Here:

    gR/A () = 1

    h i , 0+, (2.2.70)

    with the + signs corresponding to the retarded Green function,

    g< () = 2i(h )

    dd0

    , (2.2.71)

    g

    >

    () = 2i(h )1 dd0 , (2.2.72)R/A() =

    R/AL () +

    R/AR () =

    K{L,R}

    kK

    |tk|2h k i

    , 0+, (2.2.73)

    < () = () = 2i K{L,R}

    kK

    |tk|2 (h k) (1 f(k K)) , (2.2.75)

    wheref()is the Fermi-Dirac distribution and dd0represents the occupancy of the resonant35

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    36/136

    level in steady state.

    An exact expression for the stationary current through a system coupled to large noninter-

    acting metallic leads in terms of the systems Green function can be derived [120]:

    I = ie

    4h

    d

    Tr(f( L)L () f( R)R ())GR () GA ()

    +Tr(L () R ())G< ()

    , (2.2.76)

    where L () = 2Im

    RL()

    , or equivalently

    I = e

    2h

    dTr

    L() >R ()

    G< ()

    . (2.2.77)

    The resonant level model serves as a great toy model, nevertheless, it is too simple as it

    involves no interactions. The NEGF and the self-energies can be calculated exactly, and noapproximations are needed. As already mentioned when interactions are involved, one needs

    to truncate the infinite/intractable set of equations and decouple the higher order correlation

    functions in terms of the lower order ones, the resulting single-particle NEGF does not nec-

    essarily obey the symmetries and relations of equation (2.2.34). To demonstrate this we turn

    to the Anderson model at the Kondo regime and the double Anderson model described in

    subsection 1.3.2 and subsection 1.3.3 respectively.

    2.2.4 NEGF symmetry breaking within the EOM technique

    2.2.4.1 The Anderson model

    Following the derivation in Refs. [158, 115] we define the following contour ordered Green

    function:

    G

    t,

    = ih

    TC

    d () d

    , (2.2.78)

    G, = i

    h TC n () d () d , (2.2.79)

    where is the opposite spin of. Various approximate decoupling procedures can be applied to

    the many-particle Green functions [182] that are generated during the procedure of writing the

    EOM for the single-particle NEGF. Here we follow the approximation scheme used in Ref. [158]:

    1. All electronic correlations containing at mostone lead operator, are not decoupled and

    their EOM are calculated.

    36

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    37/136

    2. Higher order Green functions involving (opposite) spin correlations in the leads are set to

    zero, that is:

    ih

    TC

    Ak () Bq () Cs () d

    = 0

    where A and B are either creation or annihilation operators of the leads and Cs is asystem operator (dor dwith spin s).

    3. The remaining higher order Green functions involving lead and system degrees of freedom

    are decoupled such that

    F

    ,

    = i

    h

    TC

    ck () d () cq () d

    = kqf(k K) G

    ,

    .

    Given these approximations the EOM for G (, )can now close (full derivation can be foundin Ref. [115] pages 172-176). The resulting EOM (in steady state) are:

    GR/A () =gR/A () + g

    R/A () UG

    R/A() , (2.2.80)

    GR/A () = g

    R/A2 ()n gR/A2 () R/A1 () GR/A () , (2.2.81)

    G () = g () + g

    R () UG

    () + g

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    38/136

    gR/A2 () =

    1

    h U R/A4 (), (2.2.86)

    g () = gR ()

    0 () g

    A() , (2.2.87)

    g

    2 () = gR

    2 ()

    4 () gA

    2 () . (2.2.88)Also,

    R/A0 () =

    R/A0L () +

    R/A0R () =

    K{L,R}

    kK

    |tk|2h k i

    , 0+, (2.2.89)

    is the exact retarded/advanced self-energy for the noninteracting case. The self-energies due to

    the tunneling of the electron are R/A1 ()and R/A3 ():

    R/Aj () = R/AjL () + R/AjR () ,

    =

    k{L,R}A(j)k |tk |2

    1

    h + k U i

    + 1

    h k + i

    , j= 1, 3and 0+

    (2.2.90)

    withA(1)k = f(k K),A(3)k = 1. The lesser self energies are defined as in Ref. [186]:

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    39/136

    where

    PR/A = 1

    1 + gR/A2

    R/A1 g

    R/A U

    . (2.2.94)

    Applying the principle of reductio ad absurdum we assume G

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    40/136

    0 0.2 0.4 0.6 0.8 1

    e/U

    0

    0.1

    0.2

    0.3

    n

    Re nIm n

    Figure 2.2.4: One of the main flaws of the EOM approach for the Anderson impurity model, if the symmetrybreaking of the lesser Green function, which leads to a complex occupation numbers. Here we plot the value ofn

    (occupation of the spin up electron) as a function of the bias voltage (e/U). The most notable effect is

    the appearance of an imaginary portion ton

    (red line) as the bias voltage is increased. To obtain the results

    only the real part ofn was used to converge the self-consistent equations for the NEGF (equations (2.2.80)to (2.2.83)). Parameters used (in units ofU): L,= R,= 0.3, L,= R,= 0.05, = 0.2, = 0.2 and= 4. The wide band limit approximation was used for the calculation of the self-energies.

    This simple approximation does not violet the symmetry relations of the single-particle Green

    function (see appendix A), but as pointed above, it does not reproduce the Kondo peaks at low

    temperatures.

    2.2.4.2 The double Anderson model

    For the double Anderson model, described in subsection 1.3.3, we follow the derivation given

    in Ref. [114], and define the following contour ordered Green functions:

    G

    , =

    i

    h TC d () d

    , (2.2.98)G

    s

    ,

    = i

    h

    TC

    ns () d () d

    , (2.2.99)

    where , s = {,}and is the spin opposite to , and= {, }. In what follows we will usethe naming convention summarized in table B.1 in appendix B. The approximations used in

    Ref. [114] are:

    1. Neglect the simultaneous hopping of electron pairs to and from the system. i.e., neglect

    40

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    41/136

    all2-particle Green functions Gqrst(, ), Fqrstk(,

    )and Fqrstk(, ).

    2. Assume that

    Fsjk, =

    i

    h TC njs () ck () d

    i

    h

    dt1gk ( 1)

    TC

    d (1) njs (1) d

    ,

    wherenjs (t)is the number operator of the electron on site j and spins, and we assumed

    k L, and ih

    k

    gk ( 1) = C ( 1) .

    This is obtained by writing the EOM ofFsjk(, )and assuming that njs ()is constant,

    i.e.

    njs () = 0, which is the case in steady state. This assumption is equivalent totreating the coupling to the leads up to the second order with respect to tkm. It neglects

    processes necessary to qualitatively capture the Kondo effect [191, 192, 193], yet results

    are predicted to be reliable for temperatures above the Kondo temperature (TK) [158, 60].

    3. Higher order Green functions (3-particle) of the form

    ih

    TC

    nq() nr () d () d (

    )

    are decoupled to:

    ih

    nq()TC

    nr () d () d

    i

    hnr (t)

    TC nq() d () d

    .

    Here, = {, } .

    These approximations lead to the following equations, written in Fourier space, where ()

    was omitted for brevity (for a brief summary of the derivation of the EOM ofG(, ) see

    appendix B):

    h

    R/A0

    1

    G

    R/A= + h

    G

    R/A+ UG

    R/A

    +VG

    R/A+ V

    G

    R/A,

    (2.2.100)

    G

    R/A=

    h U V

    n

    V

    n

    R/A0

    1

    hG

    R/A+ n

    V

    G

    R/A+ V

    G

    R/A,

    (2.2.101)

    41

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    42/136

    G

    R/A=

    h U

    n

    Vn V R/A0

    1n + h

    G

    R/A+ n

    U

    G

    R/A+ V

    G

    R/A,

    (2.2.102)

    G

    R/A=

    h U

    n

    Vn V R/A0

    1n + h

    G

    R/A+ n

    U

    G

    R/A+ V

    G

    R/A,

    (2.2.103)

    GR/A

    = h U n V n V

    R/A0

    1

    hG

    R/A+

    n

    UG

    R/A+ V,

    G

    R/A,

    (2.2.104)

    G

    R/A=

    h U n V

    n

    V R/A0

    1

    dd,

    + hG

    R/A

    +n,UGR/A + V,, GR/A ,(2.2.105)

    G

    R/A=

    h U Vn Vn R/A0

    1

    n

    + hG

    R/A+

    n

    V, G

    R/A+ V,

    G

    R/A

    .

    (2.2.106)

    We now show that given this set of equations, the symmetry relationG()

    R=

    G ()A

    is not satisfied. For simplicity we prove that for the case where

    42

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    43/136

    Vsij = 0. Define

    gR/Ai () =

    1

    h i R/A0 (), (2.2.107)

    (gui)R/A () = 1

    h i Ui R/A0 (), (2.2.108)

    guni

    R/A() =

    1

    h i Ui ni R/A0 (), (2.2.109)

    R/A0 () =

    kL

    |tk|2h k i

    , 0+, (2.2.110)

    R/A

    0 () = kR tk

    2

    h k i , 0+

    . (2.2.111)

    Simple substitutions now yield the following set of equations:

    G

    R=

    1 gR

    h2 gR1

    gRhg

    R+ g

    RU

    G

    R+ gRh

    g

    RU

    G

    R,

    (2.2.112)

    G

    R=

    1 (gu)R

    gun

    R h2

    1

    guR h2 gunR

    1

    h2 (gu)R guR gunR gunR nnUU

    1

    (gu)R h

    gun

    R n + (gu)R guR hgunR n

    n

    U

    1

    gu

    R

    h

    2

    gun

    R1

    , (2.2.113)

    43

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    44/136

    G

    R=

    1

    gu

    R gun

    R h2

    1

    gunR h2 (gu)R

    1

    h

    2

    (gu)R

    guR

    gunR

    gunR nnUU

    1

    guR n+ h2 (gu)R guR gunR gunR n

    n

    U

    1

    gun

    Rh (g

    u)

    R h

    1, (2.2.114)

    and

    G

    A=

    1 gA

    h2 gA1

    gAhgA+ gAUGA

    + g

    A

    h

    g

    A

    UGA ,

    (2.2.115)

    G

    A=

    1

    gu

    A gun

    A h2

    1 (gu)Ah 2gunA

    1

    h 2 guA (gu)A gunA gunAnnUU

    1

    guA

    h gun

    A

    n

    +

    gu

    A(gu)

    A h gun

    A

    nnU

    1 (gu)A

    h 2 gunA1 , (2.2.116)

    G

    A=

    1 (gu)A

    gun

    A h2

    1

    gunA h 2 guA

    1

    h 2 guA (gu)A gunA gunAnnUU

    1

    (gu)

    A

    n

    + h

    2

    guA

    (gu)A gun

    A

    gunA

    nnU

    1

    gun

    A h2 guA1

    , (2.2.117)

    Taking the conjugate of equation (2.2.115) and using the fact that,

    gRi=

    gAi

    i {, } , (2.2.118)

    44

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    45/136

    we obtain:

    G

    A=

    1 gR

    h2 gR1

    gR

    h

    gR

    + gR

    U

    GA

    + gR

    h

    gR

    UG

    A

    ,(2.2.119)

    Comparing the result with equation (2.2.112) and since this should hold for any value ofU, it

    is obvious that for the identity

    G()R

    =

    G ()A

    to be true, the following should

    hold:

    gRUG

    R= gRh

    g

    RU

    G

    A, (2.2.120)

    or equivalently:

    G

    R = gRh GA . (2.2.121)Using equations (2.2.113) and (2.2.117) and eliminating equal terms on both sides of the equal-

    ity, we get:

    L.H.S = h

    gunR n + UguR hgunR nn ,

    (2.2.122)

    R.H.S = gRh

    (n

    +Uh2 guR gunR gunR nn

    .

    (2.2.123)

    Note that by definition:

    (gui)R =

    (gui)

    A

    i {, } , (2.2.124)

    guni

    R= g

    uni

    A

    i

    {,

    }. (2.2.125)

    Obviously G

    R = gRh

    G

    A, (2.2.126)

    hence, the relation

    G()R

    =

    G ()A

    is not satisfied under these aforementioned

    approximations and truncation of the EOM.

    45

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    46/136

    The same can be done (not presented) to show that:

    G()

    = G ()

    , (2.2.127)

    and G()

    R G()A = G()> G()< . (2.2.128)As a result the single-particle density matrix, which is obtain from the lesser Green function, is

    not Hermitian (see figure 2.2.5), the occupation of levels is a complex number, and the value of

    the stationary current calculated via equation (2.2.76) or via equation (2.2.77) yields different

    results (figure 2.2.6).

    0 0.1 0.2 0.3 0.4

    e/U

    realpart

    0 0.1 0.2 0.3 0.4 0.5

    e/U

    imaginarypart

    0 0.1 0.2 0.3 0.4 0.5

    Coherences

    imaginarypart

    Figure 2.2.5: The single-particle density matrix should be Hermitian, i.e., =

    , or in other words

    Im

    = Im

    and Re

    =Re

    with = ih2

    G ()

    G< and B = GR GA. Definethe difference anti-Hermitian matrix C=AB, and redefine the retarded and advancedGreen functions GR = GR +C/2, and GA = GA C/2.

    The resulting Green functions (GR, GA, G< and G>) obey all symmetry relations of equa-

    tions (2.2.33) and (2.2.34) by construction. Note that if the original Green functions obeyed

    the symmetry relations to begin with, our symmetrization procedure will not alter them in any

    way.

    We now return to our calculations for both the Anderson and double Anderson models. For

    both models we use the closures described above. For each set of calculations we have applied

    the above symmetrization scheme and compared the results to those obtained without restoring

    symmetry.

    2.2.5.1 The Anderson model: symmetric-EOM approach

    The closure used in Ref. [158] is sufficient to describe the appearance of the Kondo resonances

    at low temperatures, as seen in figure 2.2.7, where we plot the density of states as a function

    of energy for several temperatures (all calculations are done with symmetry restoration). The

    development of Kondo peaks in the density of states as the temperature decreases is clearly

    evident, signifying a regime of strong correlations which is qualitatively captured by the simple

    EOM approach when symmetry is restored.

    In figure 2.2.8 we plot the value of

    n

    as a function of the bias voltage with and without

    symmetry restoration. The most notable effect is the appearance of an imaginary portion ton

    as the bias voltage is increased. By applying the symmetrization scheme proposed in

    subsection 2.2.5 to the lesser Green function, we restore the relation G() =

    G ()

    .

    This is sufficient to obtain a real value for

    n

    , as clearly shown in figure 2.2.8. All other

    symmetry relation are not violated here and thus, our symmetrization procedure does not

    affect them at all. Interestingly, taking only the real part of

    n

    provides identical results

    48

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    49/136

    when compared to the results obtained after the full symmetrization procedure. However, this

    is only true for the simple case of the single site impurity model and does not hold for more

    complex systems.

    -5 -4 -3 -2 -1 0 1 2 3

    /

    0

    0.1

    0.2

    DOS(1/)

    Figure 2.2.7: Density of states (DOS) of the spin up electron as calculated using the symmetric-EOM approachfor different temperatures;= 200(red line),= 50(blue line) and= 5(black dashed line). The development

    of Kondo peaks in the density of states as the temperature decreases is clearly evident. The DOS of the spin

    up electron calculated from the unsymmetrized Green function is identical provided only the real part ofn

    was used in the calculations. Parameters used are similar to those used in Refs. [158, 186], exceptU that was

    chosen to be large (U= 10) but not infinite. Other parameters (in units of = L + R): L= 3/10,R= 0,

    ,= 2. The bands are modeled as a Lorentzian with a half bandwidth 100.

    2.2.5.2 The double Anderson model: symmetric-EOM approach

    While for the case of a single site Anderson model only the relation G() =

    G ()

    breaks down, in the double Anderson model we find that all 3symmetries described by equa-

    tion (2.2.34) are violated. This can be traced to the more complex form of the Hamiltonian for

    the double Anderson model, where each site is only coupled to one of the leads and transportin enabled by the direct hopping term between the two sites.

    Similar to the case of the Anderson model, as a result of symmetry breaking the occupation

    of the levels n is a complex number. In addition, the coherences,= ih2

    G()

    G()< is

    preserved. Indeed, in the case of a single site Anderson model, even if symmetry is not restored,

    this relation holds and the two calculations yield identical values for the current. However, in

    the present case, all 3 symmetry relations are broken and thus, equations (2.2.76) and (2.2.77)

    give different results for the current, as clearly evident in figure 2.2.10. More significantly is

    the fact that equation (2.2.76) produces a finite value for the current even when the bias iszero, indicating the break down of the fluctuation dissipation relation. When symmetry is

    restored (black line) the two calculations are identical, as they should be, and the violation of

    the fluctuation dissipation relation is also resolved.

    The symmetrization scheme proposed in subsection 2.2.5 is not a magic cure and, in fact,

    does not resolve all issues of mater. It is well known that the lesser and greater Green functions

    should obey a simple sum rule where the integral over the difference of their diagonal elements

    should always sum to 1:

    50

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    51/136

    imag

    inarypart

    0 0.1 0.2 0.3 0.4 0.5

    e/U

    0 0.1 0.2 0.3 0.4

    e/U

    realpart

    Coherences

    Figure 2.2.9: The imaginary (upper panels) and real (lower panels) parts of (blue line) and (red line)calculated before (right panels) and after (left panels) symmetry was restored. The solid black thin line in the

    upper panels marks the zero axis. As expected, after symmetry restoration (left panels), Im

    = Im

    and Re

    =Re

    , while before symmetry restoration (right panels) these equalities are violated. See

    figure 2.2.5 for parameters.

    S=ih

    2

    d

    (G ())

    < (G ())>= 1. (2.2.129)

    In figure 2.2.11 we plot the sum rule as given by equation (2.2.129) for the double Anderson

    model where symmetry has been restored. A similar plot for the single site Anderson model

    yields a value of 1 regardless of whether symmetry has been restored or not within the clo-

    sure discussed above. However, in the case of the more evolved double Anderson model, even

    when symmetry is restored and the Green functions obey all 3 relations described in equa-

    tion (2.2.34), the sum rule is violated. Nonetheless, the sum m{,}

    {,}

    Sm=Ne, where

    Ne= 4is the total number of electrons in the system at maximal occupancy, is indeed preserved

    when symmetrization is restored.

    In the remaining of the chapter, we compare 4 different closures for the EOM of the NEGF

    for the double Anderson model, two of which are developed by us while the remaining two

    can be found in the literature. We use the Green function obtained via the symmetric-EOM

    approach to calculate the differential conductance of the model. We note in passing that this is

    the simplest theoretical model that can describe transport phenomena in a donor-acceptor com-

    plex or heterojunctions, which are the working principle of many electronic and optoelectronic

    devices [194, 195].

    51

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    52/136

    0 0.1 0.2 0.3 0.4e/U

    0

    2

    4

    6

    8

    10

    12

    Stationary

    current

    (e/h_)

    Eq. (1)

    Eq. (2)

    Sym

    10-2

    Figure 2.2.10: IVcurves calculated using equations (2.2.76) and (2.2.77) before (blue and red lines respec-tively) and after (black line) applying the symmetry procedure suggested in subsection 2.2.5. See figure 2.2.6

    for parameters.

    2.3 Steady state conductance in a double quantum dot

    array:

    Assessing the symmetric-EOM technique for the NEGF

    2.3.1 Overview

    We study the role of different approximate closures to the EOM of the NEGF on steady state

    properties (namely, the differential conductance) for a double quantum dot (QD) array, coupled

    to two macroscopic leads [196] (the double Anderson model [113, 114]). Although general, this

    model can be used to study transport through single diatomic molecules (where each atom is

    represented by its conducting orbital [197] and even larger molecules and complexes [198, 199,

    200] or structures [201].

    Four closures are examined; two already proposed [114, 183] (we will refer to these closuresas approximations1and 4) and two developed by us (denoted as approximations 2and 3). The

    results obtained from the different closures were compared to the results attained using a many-

    particle Master Equation (ME) approach [202] adequate for weak hybridization (system-leads

    couplings) and high temperatures [203, 204].

    We find that in contrast to the simple case of a single site model (Anderson model) in which

    different closures beyond the simplest Hartree approximation scheme [179] yield very similar

    transport results [182] (steady state current and differential conductance as a function of the

    applied bias voltage) at high temperatures, the double Anderson model yield very different

    52

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    53/136

    0 0.1 0.2 0.3 0.4

    e/U

    0.9

    1.0

    1.1

    S

    S

    Figure 2.2.11: S (blue line) andS (red line) calculated from the symmetrized lesser and greater Greenfunctions as a function of the bias voltage. The exact result should have been 1 (as marked by the solid black

    line). Parameters used for the simulations in units ofU=U = U are: L=

    L =

    R =

    R= 0.0025,

    R= R=

    L=

    L= 0, h

    =

    h= 0.25, V = 0.1, = = 0.1, = = 0.175and = 80.

    stationary currents and differential conductance curves for the different closures studied.

    The performance of the different closures is analyzed in terms of the poles of the unperturbed

    retarded Green function in comparison to the exact many-body result for the isolated system.

    We find that one of the closures developed by us provides the most accurate description of the

    poles and also the best overall agreement with the ME approach for all parameters studied inthis work. While these results are encouraging, a word of caution is in place. It is clear that

    the conclusions drawn from the performance of the different closures for the single site model

    cannot be extended directly to the two site model, in analogy, a suitable closure for the two

    site model may fail in the larger systems. Thus, the study of larger arrays of QDs will require

    either analysis along the lines sketched here or employ a different formalism. In what follows

    we embrace the notation for the different Green functions described in appendix B.

    2.3.2 Different closures of the EOM

    The double Anderson model is described in subsection 1.3.3 and the relevant Green functions

    can be found in appendix B. Here we chose U= U= U,V = V = V and h

    = h

    = h.

    53

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    54/136

    The EOM in Fourier space of the retarded Green function is:

    G()

    R=

    h (0 ())R

    1+ h

    G()

    R+ V

    G

    ()

    R+VG()

    R+ UG()

    R

    , (2.3.1)where 0/ ()(c.f. equations (2.2.110) and (2.2.111)) is the tunneling (from the left/right

    lead) self energy (the label 0 refers to the limits U, V 0). Calculating the EOM for thehigher order Green functions will give rise to other 2-particle and 3-particle correlations:

    G

    ()

    R=

    h V

    0 ()

    R1n + h

    G

    ()

    R,

    hG

    ()

    R+ h

    G

    ()

    R+ V

    G()

    R+ U

    G()

    R+

    k

    tk Fk()R tk Fk()R + tkFk()R .(2.3.2)

    Henceforth, the superscript R and the implicit dependence on are omitted for brevity.

    G = (h U 0)1

    n + hG hG

    +hG+ VG+ VG

    +k

    tkF

    k tkFk+ tkFk

    , (2.3.3)and

    G =

    h V 0

    1 n

    dd

    +hG+ UG

    + VG

    + k tkF

    k

    tkF

    k+ t

    kF

    k

    . (2.3.4)

    To continue, one has to formulate the equations for these new Green functions, which in turn

    will lead to other (higher order) Green functions. This infinite hierarchy of equations needs to

    be truncated at a certain level, a process which is referred to as closure. In general, closures

    cannot be improved systematically. Furthermore, it is often difficult to assess, a priori, the

    accuracy of a given closure. We now discuss several different closures which are physically

    motivated, tractable, and some are commonly used in the context of transport.

    54

  • 8/10/2019 Developing methods for transport calculations in heterojunctions with Coulomb interactionsFinal Full(Heb)

    55/136

    2.3.2.1 Approximation1

    Following the derivation given in Ref. [183], the following approximations are made:

    1. All3-particle Green functions set to zero.2. Simultaneous tunneling of electrons of opposite spins are neglected, i.e., all Fqrstk, F

    qrstk

    and Gqrstare also set to zero.

    3. Remaining Green functions mixing leads and system operators are decoupled so

    Fsk

    ,

    = i

    h

    TC

    ns () ck () d

    ,

    ih

    d1gLk( 1)

    TC

    d (1) ns (1) d

    .

    See sub-subsection 2.2.4.2, second assumption, for a short discussion regarding this ap-

    proximation.

    4. The remaining2-particle Green functions of the form Gs(, )are decoupled so Gs(,

    ) =ns ()G(, ).

    The resulting equations (for the retarded Green functions) are given by:

    G = (h 0)1