l. c. andreani, optical transitions, excitons, and polaritons in bulk and low-dimensional...

Upload: vg51

Post on 04-Feb-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    1/56

    OPTICAL TRANSITIONS, EXCITONS, AND

    POLARITONS

    IN

    BULK

    AND

    LOW-DIMENSIONAL SEMICONDUCTOR STRUCTURES

    Lucio Claudio Andreani

    Dipartimento

    di Fisica A. Volta

    Universita degli Studi di

    Pavia

    via A. Bassi, 6

    1-27100 Pavia,

    Italy

    1 INTRODUCTION

    Electrons in solids are

    subject to the

    crystal potential, as well as

    to the mutual

    electron

    electron interaction.

    The

    resulting quantum-mechanical system represents a many-body

    problem of great complexity.

    In

    weakly-correlated systems like

    the

    usual semiconduc

    tors,

    a good starting point is provided by the one-particle picture, in which the crystal

    eigenstates are approximated by Slater

    determinants

    where

    the

    electrons occupy

    the

    one-particle eigenstates called

    band

    levels. This is only an

    approximate picture,

    since

    the

    electron-electron

    interaction

    yields corrections

    to the

    excited-state

    spectrum

    of

    the

    crystal. In particular,

    two-particle excitations called excitons arise

    at

    energies below

    the band gap,

    and

    excitonic corrections are found also

    at

    energies above the band gap.

    The

    electronic

    states of

    a crystal can

    be

    probed

    using

    an external

    electromagnetic

    field.

    The study

    of

    the

    optical properties gives very precise information on

    the

    elec

    tronic structure of semiconductors. In analyzing

    the

    radiation-matter interaction,

    it

    is

    useful to distinguish between instantaneous and retarded parts of

    the

    electromagnetic

    field.

    The unretarded

    (c

    - ;

    00

    part

    describes

    the

    instantaneous Coulomb interaction

    and

    corresponds

    to

    the

    longitudinal electromagnetic field, while

    the retarded part

    is

    identified with

    the

    transverse electromagnetic field, i.e., with

    the

    physical photons. For

    one-particle states, interaction with

    the

    transverse electromagnetic field gives rise to

    interband

    and

    intraband

    transitions. For two-particle (excitonic) states,

    interaction

    with

    the

    longitudinal

    part

    of

    the

    electromagnetic field corresponds

    to

    the

    electron-hole

    exchange interaction, while

    interaction

    with

    the

    transverse electromagnetic field gives

    rise

    to

    polariton effects.

    The

    above picture applies to bulk semiconductors as well as to mesoscopic struc

    tures

    like

    quantum

    wells, wires

    and

    dots. However

    both the

    electronic states

    and

    onfined Electrons and Photons

    Edited by

    E.

    Burstein and C. Weisbuch, Plenum Press, New York, 1995

    57

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    2/56

    the radiation-matter

    interaction are modified by

    the

    reduced dimensionality. One of

    the

    most important modifications concerns

    the

    polaritonic effect. In infinite crystals,

    the

    conservation

    of

    crystal

    momentum

    implies

    the

    formation of quasi-stationary

    states

    called excitonic polaritons. In confined systems, the lack of crystal momentum con

    servation

    in

    one

    or

    more directions implies that an exciton

    interacts with

    a

    continuum

    of photon states, thereby

    providing a mechanism for intrinsic radiative decay of free

    excitons.

    In these lecture notes we give an outline of the theory of linear optical properties in

    semiconductors, considering

    both

    bulk

    and

    confined systems. We consider one-particle

    properties

    (interband and

    intraband transitions) as well as exciton

    and

    polariton effects.

    One main point is

    the

    modification of exciton states

    and

    polari ton effects in going from

    bulk

    to

    quantum-well systems, considering in particular

    the

    exciton radiative lifetime.

    In

    Sec. 2

    we treat

    optical properties

    in

    bulk semiconductors. After a review

    of

    the

    classical theory

    of

    dielectric properties, including

    the

    Lorentz-oscillator model

    and

    Kramers-Kronig relations, we describe

    the

    calculation of

    the

    dielectric

    constant

    by a

    semiclassical theory

    of

    the

    radiation-matter

    interaction. We

    then

    give an outline

    of

    the quantum theory

    of excitons

    and

    polaritons. In Sec. 2.5 we summarize

    the present

    understanding of the radiative recombination of excitons in bulk semiconductors, which

    is a basic problem but still not completely understood. In Sec. 3 we discuss optical

    properties

    in

    confined systems, particularly in

    the

    quantum-well geometry. Following a

    similar scheme, we first review

    the

    theory of one-particle transitions,

    and

    then consider

    exciton

    states and

    polariton effects. In Sec. 3.4 we discuss

    the

    radiative lifetime of

    excitons in going from bulk

    to

    confined systems. In Sec. 4 we summarize the

    main

    points

    treated

    in these lectures.

    In the Appendix

    we give rules for converting from the

    Gaussian cgs)

    to the

    SI MKSA) system of units.

    2 OPTICAL PROPERTIES IN BULK

    SEMICONDUCTORS

    2.1 Classical Theory

    of

    Dielectric

    Properties

    Classical electromagnetic

    theory.

    Maxwell equations in

    matter

    must be supple

    mented by constitutive relations. We neglect magnetic effects,

    and

    consider only linear

    response.

    In

    a homogeneous

    and

    isotropic medium, the Fourier components

    of the

    polarization

    and

    displacement fields are related to

    the

    electric field by Gaussian units)

    P

    D

    XE

    E + 47rP

    =

    tE,

    1)

    2)

    where t(w,k) = 1 47rX w,k) is

    the

    frequency-

    and

    wavevector-dependent dielectric

    function.

    The

    wavevector dependence of t is referred to as spatial dispersion

    [1]. t

    can often be neglected since

    the

    wavelength of radiation is much larger than

    the lattice

    spacing. In a crystal, which is invariant only under translations by lattice vectors, the

    macroscopic dielectric function contains local-field effects

    [2]. In

    anisotropic

    media, the

    dielectric function becomes a 3 x 3 tensor, which reduces

    to

    a scalar only for cubic

    crystals in

    the

    long-wavelength limit.

    The

    displacement

    current

    appearing

    in

    Maxwell equations is given

    by

    J _ ~ a

    d - 47r t

    In the

    presence of free charges, an induced

    current

    appear according

    to Ohm s

    law,

    J

    ind

    = erE

    58

    3)

    4)

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    3/56

    Both

    currents

    (3),(4) can be taken

    into

    account by introducing a complex dielectric

    function,

    5)

    At finite frequencies, the conductivity u(w)

    is

    also complex. Thus it is a matter

    of

    convention

    to attribute the

    dispersive properties

    of

    a

    medium

    to

    a real dielectric function

    plus a conductivity, or

    to

    a complex dielectric function E = E1 E2

    The index of refraction,

    N = n K

    is also complex. The absorption coefficient

    is

    obtained as

    w

    a =

    -K .

    c

    When E2

    E1,

    Eq. (7)

    can

    be approximated by

    6)

    7)

    8)

    This is a good approximation for semiconductors, but

    it

    may

    fail for metals (see Sec.

    2.2).

    Oscillator model. A simple classical picture for the dielectric function is provided

    by the Drude-Lorentz

    model

    [3, 4J. The electrons

    in

    a crystals are represented by a

    collection of

    damped

    harmonic oscillators, which respond

    to

    an applied electric field

    according

    to the

    equation

    [

    2 J E

    mo x

    ijX

    W j X = - e loc,

    9)

    where Eloc is

    the

    local field acting on the oscillator and mo is the free-electron mass.

    The

    induced

    dipole moment for each oscillator is

    Pj = eXj =

    2 2 . ).

    mo Wj - W - t i j W

    10)

    The macroscopic polarization

    is

    found by multiplying Pj by the number of oscillators

    per unit volume, which we denote by Ii V, and summing over all resonances. f the

    local field is identified with the applied field (i.e., if local-field corrections

    are

    neglected),

    the resulting dielectric function is

    47re

    2

    1 Ii

    E

    w

    =

    1

    mo

    V L

    ~ _ w

    2 _ ii W

    J J J

    (11)

    The

    oscillator strength

    Ii

    is a dimensionless quantity, which represents

    the

    number of

    classical oscillators

    at

    frequency

    Wj

    Since

    the total

    number of oscillators equals

    the

    number

    N of electrons in the crystal,

    the

    oscillator strengths

    must

    satisfy

    the sum

    rule

    L i =

    N

    12)

    j

    This

    sum rule

    can be

    derived

    by

    comparing the high-frequency limit of Eq. (11) with

    the

    known limiting form of

    the

    dielectric function,

    13)

    59

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    4/56

    where w;

    =

    47rNe

    2

    /(moV)

    is

    the plasma frequency.

    Equation 11) yields

    the

    real and imaginary parts of the dielectric function,

    the

    index

    of refraction, absorption coefficient, etc. Close

    to

    a resonance frequency, the real

    part l W) shows anomalous dispersion, while

    2 W)

    yields resonant absorption

    [3,4].

    Free charges

    contribute

    a frequency-dependent conductivity u(w)

    = ne

    2

    /(m*(( - iw)),

    where

    m

    is the effective mass.

    The

    relation between 11) and

    the

    electronic structure

    is given by a microscopic calculation

    of the

    parameters

    Wj,

    Ii

    and i j In

    a solid,

    in

    the energy region

    of

    interband transitions, the frequencies W j form a continuum. A

    quantum-mechanical calculation of

    the

    dielectric function, with a suitable definition of

    the oscillator strength, will

    be

    described in Sec. 2.2.

    Dispersion

    relations

    nd sum rules. General properties of

    the

    optical constants

    follow from the causality principle. The displacement field is related

    to

    the electric field

    by

    D t) = E t)+47r l O x r)E t-r)dr,

    14)

    where the response function

    x r)

    can be taken to vanish for

    r

    ,

    bD

    h

    Ql

    I i

    Ql

    0

    HH1

    x

    o

    -30

    LH1

    -90

    HH3

    LH2

    (0)

    1 5 0 0 L L ~ 1 L ~ ~ 2 ~

    k L

    x

    =

    3.0

    k or

    HH1

    LH1

    HH2

    b)

    o

    1

    2

    k L

    Figure 7. Dispersion

    of

    the valence subbands in a 80 A wide GaAs-Gao.6Alo.4As quantum well a)

    with no applied stress b) with an applied stress X = 3.0 kbar along the growth direction.

    absorption probability for a single

    quantum

    well is a

    pure number,

    which can

    be

    defined

    as

    ) _ energy absorbed/unit time surface

    w w - incident

    energy/unit time

    surface

    91)

    The absorption probability can

    be

    calculated

    by

    time-dependent perturbation theory,

    as in Sec. 2.2. The result is

    92)

    where kll is

    the

    in-plane wavevector. Note that

    w w)

    is dimensionless, while

    the

    ab

    sorption coefficient 30)

    has the

    dimensions of em 1 .

    It

    is important

    to

    realize that

    the

    absorption coefficient 29) of a

    plane

    wave for

    a sample containing a single

    quantum

    well is zero in

    the thermodynamic

    limit. This

    implies

    that

    a plane wave propagating along the layer planes

    is

    not attenuated. There

    is no contradiction with experiment, since the case of an infinite plane wave is never

    realized

    in

    practice. In the real case of an incident light beam with diameter

    C

    the

    absorption per

    unit length

    is of

    the

    order of w w)

    /

    C

    (apart

    from a factor of order

    unity which depends on the electric field profile).

    Thus

    the attenuation of a light beam

    propagating

    in the layer planes can

    be

    increased

    by

    decreasing the

    spot

    diameter. A

    finite absorption of order w w)/C is also obtained for light propagating in a waveguide

    configuration,

    and

    in this case

    the

    relevant length C is the thickness of the waveguide

    [75] This is

    to

    be

    contrasted

    with

    the

    case of light propagating along

    the

    growth

    direction, where

    the

    dimensionless absorption probability 91) is

    independent

    of

    the

    spot size.

    81

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    26/56

    Multiple

    quantum

    wells MQWs) are usually employed in order to increase

    the

    optical efficiency. For a sample with

    N

    periods, the transmittance

    T

    is related to

    the

    dimensionless absorption probability 91) by

    T

    = 1 - wt =

    exp N

    .log 1 -

    w))

    exp -Nw ,

    93)

    and

    the usual

    formula for exponential decay is recovered. Macroscopically,

    the

    absorp

    tion coefficient per unit length is obtained as

    N Nw w

    a = - - l og 1 - w

    -

    = ,

    w

    b

    94)

    where w b is

    the MQW

    period

    and

    =

    N Lw L

    b

    )

    is

    the

    thickness of

    the

    sample.

    This derivation shows that the

    proper

    length

    by

    which the dimensionless absorption

    probability

    has to

    be

    divided

    is

    the period

    w b not

    the well

    width

    Lw).

    A

    MQW

    sample can also be considered as a uniaxial crystal with a unit cell of size

    w b along

    the

    growth direction. The macroscopic absorption coefficient can again

    be

    defined as in Eq. 29), and

    can be

    calculated by

    the

    procedure

    of

    Sec. 2.2.

    When

    the

    electronic states are localized along the growth direction,

    the

    analog of Eq. 28) is

    E E 2 E = S E E

    r

    2dkll

    =

    1 E

    r

    2dk

    l

    l

    V .

    cv

    k V cv JBZ

    (27r)2 w

    + b cv

    JBZ (27r)2

    J J

    (95)

    where E

    j

    denotes

    the

    sum over

    the

    identical QWs and S is the

    area

    of the sample.

    Thus we see that formula 94) gives the absorption coefficient of a MQW sample for

    any

    direction of propagation.

    The

    physical meaning of Eq. 94) is that

    the

    attenuation

    of a light beam in a MQW sample is proportional to the concentration of absorbing

    regions,

    and

    therefore

    must

    depend

    on

    both

    well and barrier widths.

    A

    limiting case

    of

    Eq. 94) is the case of a single quantum well, when

    the

    barrier width b - - t and the

    macroscopic absorption coefficient vanishes.

    A description

    of

    optical absorption in quantum well

    structures

    which is

    based

    on

    the absorption probabil ity has two advantages: first, the absorption probability for

    interband

    transition

    is essentially

    independent

    of the well

    width

    see below),

    without

    the artificial

    1/

    w dependence which

    is

    sometimes

    reported

    in the literature as a result

    of dividing the

    optical density by

    the

    well width

    [76J.

    Second,

    the

    absorption probability

    integrated

    over

    the

    excitonic peak is directly related

    to the

    oscillator

    strength per

    unit

    area , which is

    the

    basic

    quantity

    character izing quasi-two-dimensional excitons see Sec.

    3.2).

    Interband

    transitions. For transitions between valence and conduction subbands,

    the

    momentum matrix element can be evaluated by keeping only the lowest-order

    term

    in Eq. 89), and by neglecting

    the

    gradient of

    the

    slowly-varying envelope function.

    The

    result is

    (96)

    Selection rules come from

    the

    envelope-function as well as from

    the

    Bloch-function

    part.

    In the approximation of infinite barrier height,

    the

    envelope functions of valence

    and

    conduction subbands

    are

    identical, which leads

    to the

    selection rule

    D.n

    = 0

    [77J.

    This is

    only approximately

    true

    for a well of finite barrier height, since particles with different

    effective masses have different penetration in

    the

    barriers. In any case, transitions with

    D.n = 1,3, . . .

    are forbidden by parity. Selection rules for

    the matrix

    element between

    Bloch functions can be found using

    the

    Wigner-Eckart theorem

    [70J,

    or, in a more

    82

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    27/56

    elementary

    way, by observing that heavy holes

    at

    kll

    =

    0 have an angular

    momentum

    s =

    3/2,

    while light holes

    and

    electrons have s = 1/2. For light polarized in

    the

    layer planes, both heavy- and light-hole transitions are allowed. However, for light

    polarized along

    the

    growth direction,

    the

    operator

    f

    . p does

    not

    change

    the

    angular

    momentum and only light-hole transi tions are allowed. The ratios of oscillator strengths

    for

    interband

    transitions at kll = 0 are ~ ~ l

    :

    ~ ' J )

    : frJ)

    = 4 : 3 : 1.

    Equation

    (96) shows

    that

    the interband momentum matrix

    element

    between

    Bloch

    functions is essentially

    the

    same for bulk and QW systems. Therefore, the oscillator

    strength

    (34) is also

    the

    same. Differences in

    the

    optical absorption come from

    the

    joint

    density of states. In an ideal two-dimensional system,

    the

    joint density of states has a

    staircase form [77].

    Absorption

    of a single quantum well is low, being of

    the

    order of a percent. This can

    be easily estimated from Eqs. (92)

    and

    (96), which yield (for the heavy-hole transition in

    the

    case of infinite barrier heights) w

    1Ie

    2

    /

    nne ) fl,fm:).

    The

    absorption probability

    is

    to

    a first approximation independent of

    the

    well thickness, since

    the

    two-dimensional

    density of states P D = p,/ In2 is a constant. This has been verified experimental ly [78].

    Corrections

    to

    this result

    can

    come from leaking

    of

    the

    wavefunction

    in the

    barriers,

    which reduces

    the

    overlap integral,

    and

    from

    the

    slight variation of

    the

    Sommerfeld

    factor

    with

    the well width (see below).

    The quantitative

    evaluation of

    interband

    absorption in QWs is

    made

    complicated

    by valence band mixing. At kll 0, heavy and light hole states become mixed and

    transitions between all pairs of conduction and valence subbands are possible. The

    ones which

    are

    forbidden at kll

    =

    0 acquire a larger oscillator

    strength

    at those values

    of kll for which

    the

    corresponding valence

    subband

    is strongly mixed

    with

    the sub

    band of an allowed transition. The negative curvature of a subband also increases the

    absorption,

    because

    it

    gives a large joint density of states.

    An

    example

    of

    calculated

    interband

    absorption for transitions

    to the

    first conduction

    subband is

    shown in Fig. 8.

    A phenomenological width

    r

    = 2 meV has been introduced.

    The

    peak at the energy

    of

    the LHI-CBl transition

    comes from

    the

    negative curvature of

    the LHI subband

    (see

    Fig. 7).

    t must be stressed that

    the interband

    absorption shown in Fig. 8 is

    not

    a measurable

    quantity,

    since

    the measured

    absorption contains excitonic effects.

    The

    Sommerfeld

    factor in the strict two-dimensional limit has been calculated by Shinada

    and

    Sugano

    [79],

    with the

    result

    S w)

    _

    - 1

    +

    e-

    (97)

    This factor is two

    at the

    subband edge, and decreases slowly with an energy scale given

    by the

    effective Rydberg. For realistic

    quantum

    well excitons,

    the

    enhancement

    factor

    is calculated to

    be

    between

    1.3

    and

    1.4

    at

    the

    absorption edge

    [80].

    Including

    the

    excitonic effect,

    the

    absorption probability in a GaAs-Gal_xAlxAs single

    quantum

    well

    is

    w 0.75 for

    the

    first HH transition, and w 1.0 for

    the

    lowest HH and LH

    transitions together

    (see Fig. 8 and Ref. [78]).

    A schematic plot of the absorption in two-dimensional systems is shown in Fig.

    9. As for three-dimensional excitons, the absorption probability corrected for excitonic

    effects is continuous

    at the

    interband absorption edge [79]. Recently, analytical formulas

    for

    the absorption

    lineshape in

    quantum

    wells including

    bound and continuum

    exciton

    states

    have been developed [81, 82], which are based on

    the idea that the

    excitonic

    spectrum

    has a dimensionality intermediate between

    2D

    and 3D.

    83

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    28/56

    1.5

    LH1

    HH2

    HH3

    LH2

    t

    '

    1.0

    e z

    -

    c

    a

    :.;:;

    Q..

    e x

    a

    0.5

    f)

    L l

    0

    80 130

    18

    E - Egop

    meV)

    Figure 8. Interband absorption probability in a 80 Awide GaAs-Gao.6Alo.4As quantum well,

    without

    excitonic effects.

    n=l n=2 n=3

    E

    Figure 9 Schematic picture

    of

    the absorption probability in two-dimensional systems,

    without

    exci

    tonic effects dashed line) and with excitonic effects solid line).

    84

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    29/56

    Intraband transitions. In bulk crystals,

    intraband

    transitions can only occur in

    the

    presence of some scattering mechanism (see Sec. 2.2). In quantum-confined struc

    tures, intraband (or intersubband) transitions are made possible by

    the

    discretization

    of the energy levels. Such transitions can only be observed in doped or photoexcited

    samples [70, 83, 84].

    The

    dipole

    matrix

    element for

    an intersubband transition

    between

    states of the

    form (89) describes

    what

    is called

    an

    envelope-state transition.

    The term

    containing

    the

    derivative of

    the

    Bloch function now gives no contribution, since

    the momentum

    matrix element between Bloch functions of either the conduction or

    the

    valence band

    vanishes.

    The term with the

    gradient of

    the

    envelope function is of

    the

    same

    order

    as

    that coming from the correction

    to

    the lowest-order wavefunction on

    the

    r.h.s. of Eq.

    (89). This second

    term

    in fact dominates,

    and

    has

    the

    effect of multiplying

    the

    matrix

    element by

    maim:. The

    result is

    (98)

    For light polarized along

    the

    growth direction, transitions are allowed only between

    conduction (or valence) states of opposite parities. For the transition between

    the

    two lowest conduction levels,

    the matrix

    element

    in

    the

    case of infinite barriers

    is

    8/3) -in/Lw . The

    oscillator

    strength

    (34) for

    the

    lowest intersubband

    transition

    is then evaluated as

    fz 256

    ma

    0 96

    ma

    ,

    271r2

    m; m;

    (99)

    i.e., it is nearly identical to that for interband transitions [85 86]. However

    the joint

    density of states

    is quite different: for parabolic

    subbands

    with

    the

    same effective mass,

    the transition

    energy would be

    the

    same

    at

    all wavevectors, resulting in a 5-function

    line where all the oscillator

    strength

    is concentrated. Line broadening comes from

    conduction band nonparabolicity, as well as from many-body effects [87]. Assuming a

    Lorentzian broadening,

    the intersubband

    absorption probability can be calculated from

    (92)

    and

    (98), with the result

    (100)

    where

    ns

    is

    the

    areal density of electrons in

    the

    first conduction

    subband. Intersubband

    absorption is smaller

    than

    interband absorption by a factor v fi

    2

    n

    s

    / m:,)

    [70,86].

    When

    the

    transition

    matrix

    element is evaluated in

    the

    gauge

    eE

    r,

    the

    so-called

    giant electric dipole is found [85]. Of course

    the

    physical results are independent of

    the

    choice of

    the

    gauge,

    and

    in fact the same expression (99) for

    the

    oscillator strength is

    found in

    the

    two gauges.

    The

    physical meaning of

    the

    giant electric dipole is that an

    oscillator strength comparable to that of interband transitions is found in

    the

    infrared,

    i.e., at a much lower energy as compared

    to interband

    transitions.

    For light polarized in

    the

    planes, intersubband transitions become second-class (i.e.,

    the matrix

    element is linear in

    the

    wavevector).

    In the

    valence

    band they

    become

    allowed due

    to

    valence

    band

    mixing. A theoretical analysis of

    intersubband

    transitions

    in the valence band has been given in Ref. [88].

    3.2 Excitons in Confined Systems

    Thin-film and QW regimes: overview. In

    this Section we discuss a few aspects

    of the theory of excitons in

    thin

    layers [89, 90]. Two regimes can be distinguished,

    85

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    30/56

    according to the relative values of

    the

    well width L and

    the

    exciton radius a o In

    the

    limit

    L

    ai3, which we shall call

    the thin-film regime, the

    excitonic

    Rydberg R*

    is

    much

    larger than the quantization energy

    2

    / m* L2), and the exciton is only weakly

    perturbed

    by confinement.

    The

    internal electron-hole wavefunction is

    undistorted,

    but

    the

    motion of the center of mass of

    the

    exciton

    is

    quantized [91, 92]. In

    the

    limit

    L

    v ai3 which we call

    the quantum-well regime), the

    excitonic Rydberg is smaller than

    the

    quantization

    energy

    of

    the

    subbands,

    and

    separate quantization

    of

    the

    electron

    and

    hole subbands occurs.

    The

    distortion of

    the

    internal exciton wavefunction

    due

    to

    a decrease of the average electron-hole separation leads to

    an

    increase

    of

    the binding

    energy

    and

    of

    the

    oscillator strength per unit area as the well width is reduced [93,94].

    Excitons in semiconductor heterostructures are usually described within

    the

    envelope

    function scheme, with the electron and hole confinement potentials being added to the

    effective-mass Hamiltonian. Due

    to

    translational invariance in

    the

    layer planes,

    the

    exci

    ton is characterized by

    an

    envelope function F p,

    Ze,

    Zh), where p is the in-plane relative

    coordinate. A suitable description of

    the

    exciton in

    the

    thin-film regime neglecting

    the

    valence-band degeneracy and in the assumption of infinite barriers)

    is

    provided by the

    variational wavefunction of D Andrea and Del Sole [92],

    F p,Ze,Zh)

    =

    N[cos KZ) - Fc z)cosh PZ)

    +

    Fo z)sinh PZ)]e-

    r

    /

    a

    ,

    101)

    where r is

    the

    electron-hole relative coordinate,

    Z.is the

    center-of-mass coordinate along

    the

    growth direction, N

    is

    a normalization factor, and P,a are variational parameters.

    The

    functions

    Fe z), Fo z)

    are determined by

    the

    fulfillment of

    the

    no-escape boundary

    conditions

    F ze

    = ~ = F Zh = ~ = o

    The

    requirement that

    the

    wavefunction

    has a continuous derivative leads to

    the

    following quantization condition for the center

    of-mass wavevector K

    KL PL

    t a n T

    Ptanh

    T

    = o

    102)

    The quantity

    1/

    P can be

    interpreted

    as a dead-layer thickness

    [40]

    and

    is of

    the order

    of the exciton radius. For P L

    1

    the quantization condition reduces to the center-of

    mass quantization K = r / L - 2/P).

    The

    exciton levels are those of a particle of mass

    M in a

    box of

    width Lcff

    =

    L

    -

    2/P For thin layers, the center-of-mass

    quantization

    is

    only approximately true, and the full quant izat ion condition 102) must

    be

    used. The

    transition-layer thickness 1/

    P

    decreases with

    the

    well width.

    In

    the

    quantum-well regime,

    the

    exciton wavefunction is more suitably

    represented

    in a basis consisting of products of conduction

    and

    valence subbands,

    F p, Ze, Zh) = L

    Ajjle-UIPCj Ze)Vj{Zh).

    j l

    103)

    For very narrow wells, only one pair of valence/conduction subbands needs to kept. In

    the limit L 0 if the barriers are taken of infinite height, the binding energy tends

    to

    the value

    appropriate

    to the two-dimensional Coulomb problem,

    R;D =

    4R;D

    [79]

    With finite barrier height,

    the

    binding energy reaches a

    maximum and

    tends

    to the

    value appropriate to

    the

    barrier material as the well width tends to zero [95]. For wide

    wells, several pairs of subbands must be kept in

    the

    expansion 103).

    The

    wavefunctions

    101) and 103) have been shown to match for L v 3ai3 [96].

    A quantitative determination of the exciton binding energy in quantum wells is

    complicated by the interplay of several effects

    [73].

    In real

    structures,

    the finite height

    of

    the

    barriers

    and the

    variation of

    the

    band parameters from one material

    to

    the

    other

    must

    be taken into account. Due

    to the

    degeneracy of the valence band, the kinetic

    86

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    31/56

    20

    18

    :>

    16

    -;;::

    14

    \: '

    5

    12

    10

    8

    HHI-CBI Is)

    6 ~ ~ ~ ~ ~ L ~ L ~ L ~ L ~ ~

    o

    50

    100

    150 200

    well widUl

    A)

    100

    80

    ~ 6

    5

    il 40

    lS

    t

    20

    \t

    H H I C B I

    Is)

    \

    x = O . 2 S

    \

    b) \ o x=OAO

    ,

    ,

    ,

    ,

    O ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ L ~ ~ ~

    o 50 100

    150

    2

    well width

    A)

    Figure

    10. a) Binding energy of the ground-state HHI-CBl exciton and

    b)

    oscillator strengths

    per unit area

    for in-plane polarization

    of the

    ground-state

    HHI-CBl and

    LHI-CBl excitons in

    GaAs/Al Gal_ As quantum wells. From Ref.

    [73]

    energy of the holes must be described by the Luttinger Hamiltonian: thus the effective

    mass Hamiltonian must be taken to be

    2

    H =

    Ec -iVe)5

    +T

    / - iv\)

    - I e

    5

    + Ve Ze) +Vh Zh))5

    ,. 104)

    Eb

    re

    -

    rh

    When the

    off-diagonal

    terms

    of

    the

    Luttinger Hamiltonian are neglected, heavy-

    and

    light-hole excitons become uncoupled; however, coupling between

    the

    two series has an

    important

    effect

    on the

    binding energies

    and

    on

    the

    oscillator

    strengths.

    Other effects of

    a comparable size are nonparabolicity of

    the

    bulk conduction band

    and

    the difference in

    dielectric constants between well and barrier materials. All the above effects go in

    the

    direction of increasing the binding energy,

    and

    taken together result in very high binding

    energies see Fig.

    lOa), which in

    GaAs/

    AlAs

    quantum

    wells can

    be

    even higher

    than

    the

    two-dimensional limit [73]. This prediction has recently been confirmed experimentally

    [81]

    The

    combined effect

    of

    nonparabolicity

    and the

    dielectric mismatch results in a

    decrease of

    the

    critical well width, below which

    the

    exciton binding energy tends

    to the

    value appropriate to

    the

    barrier material. However a quantitative calculation of this

    effect is still lacking. A similar problem arises in short-period superlatt ices: when

    the

    superlattice

    period decreses below a critical value,

    the

    exciton wavefunction spreads

    out

    and extends

    over several layers

    [97]

    However

    the maximum

    value

    of

    the

    binding energy

    is likely

    to

    be strongly increased by nonparabolicity

    and the

    dielectric mismatch.

    In

    general,

    the

    increase of the binding energy due to

    the

    smaller dielectric constant of

    the

    barrier

    often called

    dielectric confinement

    [98]) is of great interest as a way

    to

    increase

    the stability

    of

    the

    exciton against

    thermal

    dissociation.

    Oscillator strength. The oscillator strength is defined, as usual , by Eq. 34).

    For excitons in

    thin

    layers,

    the

    oscillator

    strength

    is proportional

    to the

    area S of

    the

    87

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    32/56

    10

    8

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    33/56

    accurate calculation of

    the

    oscillator

    strength

    requires theories which take into account

    valence band mixing and coupling between excitons belonging to different subbands

    [100, 73]. For quantum wells

    of medium width

    the two-dimensional approximation

    overestimates

    the

    oscillator

    strength

    by a factor between three

    and

    four. The oscillator

    strengths

    of

    the ground-state excitons in GaAs-Gal_xAlxAs quantum wells are shown

    in Fig. lOb, and

    are

    in good agreement with absorption

    and

    reflectivity measurements.

    Excitons in QWs:

    symmetry

    properties

    and selection

    rules.

    Due to valence

    band mixing, heavy- and light-hole subbands in quantum wells are coupled for finite

    values of

    the

    in-plane wavevector. Since exciton states are built

    up

    from band states

    with wavevectors v

    lias

    valence

    band

    mixing can be expected

    to playa

    role on

    the

    selection rules for excitonic transit ions. Selection rules for excitons in

    quantum

    wells

    have been derived in a model-independent way using

    symmetry arguments

    [101, 102].

    Here we give just a few relevant examples.

    The point group Td of the zincblende structure becomes D2d in quantum wells, due

    to the

    reduction of

    symmetry.

    The

    lowest heavy-hole and conduction

    subbandshave

    r6

    symmetry

    at

    kll=O, while

    the

    lowest light-hole

    subband

    has

    r7

    symmetry.

    The

    symme

    try of exciton

    states

    is given by the product decomposition of 61). For

    ground-state

    excitons,

    the

    envelope function transforms according

    to the

    identity representation.

    Thus the ground-state HH1-CB1 exciton transforms as r6 r6 = r

    1

    9 r

    2

    9 r

    s

    , while

    the ground-state

    LH1-CB1 exciton has multiplicity

    r6

    r7

    = r3 E9 r

    4

    E9 rs The

    representations r

    1

    r

    2

    r

    3

    r

    4

    are nondegenerate, while rs is twofold degenerate. Since

    the

    z-component of

    the

    dipole operator has

    4

    symmetry, while the x components

    transform according

    to

    the

    rs

    representation, we see that for in-plane polarization both

    heavy- and light-hole excitons are allowed, while for light polarized along

    the

    growth

    direction only

    the

    light-hole exciton is allowed. t can also be shown that the oscillator

    strength for

    the

    z-polarized LH exciton is four times that for in-plane polarization.

    Although

    the zincblende lattice is not invariant

    under

    space inversion, effects re

    lated to the violation of inversion symmetry are small and difficult to observe in III-V

    semiconductors. Therefore, parity selection rules hold for excitons in

    quantum

    wells to

    a very good approximation. For example, excitons like LH1-CB2 or HH2-CB1 which

    belong to conduction and valence sub bands of opposite parities are forbidden in the

    s-like

    ground

    state. Such excitons can be observed only in p-like excited

    states.

    Parity

    symmetry is broken in the presence of

    an

    electric field applied

    or

    built-in). We refer

    to

    [101, 102,

    73]

    for details.

    Parity

    selection rules for two-photon

    transitions

    are op

    posite to those for one-photon transitions, thereby allowing very useful complementary

    information to be obtained

    by

    two-photon spectroscopy [102].

    The above selection rules maintain their validity also when valence band mixing

    is taken into account.

    In

    fact, selection rules for ground-state excitonic transitions

    are identical with those for

    the

    subbands

    at

    kll=O, since excitons

    at

    kcx=O transform

    according to irreducible representations of the crystal point group.

    The

    role of valence

    band mixing is to give a finite oscillator strength to some excitons not in s-states like

    LH1-CB2

    2p).

    Also, valence

    band

    mixing changes

    the

    ratio of

    the

    oscillator strengths

    of the in-plane polarized ground-state HH1-CB1 and LH1-CB1 excitons from

    the

    value

    3:1 characterist ic of

    the

    subbands at kll=O) to

    about

    two [73]

    Excitons in QWs:

    exchange

    interaction. The electron-hole exchange interac

    tion splits the exciton states corresponding to different irreducible representations of

    the

    point group. t is proportional to the singlet component of a given exciton state

    via the spin-orbit

    factor g

    The spin-orbit factors for

    QW

    excitons are: for the r

    89

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    34/56

    LHI-CBl

    HHI-CBl

    4 1)

    r

    /

    /

    /

    /

    5

    (2)

    / ~

    / /

    1

    :;

    __

    _X_I6_ 4_)--1.L____ _ 1_)

    5

    (2)

    It

    x I6 4)

    x.y

    x.y

    Figure

    12. Schematic representation

    of

    the effect

    of the

    exchange interaction on

    the

    ground-state

    excitons

    in quantum

    wells. Numbers

    in

    parentheses denote the degeneracy

    of the

    state.

    representation = 1

    and

    =

    1/3

    for HH

    and

    LH excitons, respectively; for

    the

    r

    4

    representation = 4/3 for the light-hole exciton. All other states have = 0 and

    no exchange contribution to the energy. The exchange splittings of the

    ground-state

    heavy-

    and

    light-hole excitons in QWs are illustrated in Fig. 12.

    In

    bulk semiconductors,

    the

    dipolar part of

    the

    exchange interaction

    is

    nonanalytical

    for k

    ex

    -- 0 and gives rise to a splitting between longitudinal and transverse states with

    the symmetry of

    the

    dipole see Sec. 2.3). For an isolated quantum well, we use

    the

    following terminology: taking

    the

    exciton wavevector k

    ex

    =

    kxx

    along

    the

    x-axis, we

    denote

    by

    T mode

    the

    rs exciton with polarization vector E ii we call L-mode

    the

    other

    rs

    state

    with polarization E

    x

    and we call

    Z mode the

    r

    4

    exciton with Ell z.

    The

    Z mode exists only for the light-hole exciton.

    Thus

    the longitudinal-transverse spli tting

    is defined

    to

    be

    the

    exchange splitting between

    Land

    modes with rs symmetry.

    The exchange interaction for QW excitons has been calculated in Ref.

    [26]

    with

    the

    Wannier-function formalism, and in Ref. [103] with the k-space methods.

    The

    two procedures yield essentially

    the

    same results.

    In

    Fig. 13 we show

    the

    long-range

    part of the exchange interaction for the light-hole exciton as a function of the in-plane

    exciton wavevector k

    ex

    I t can be seen that

    the

    longitudinal-transverse splitting vanishes

    linearly in k

    ex

    for k

    ex

    --) O

    This is an effect of reduced dimensionality,

    and

    it can be

    understood

    from

    the

    fact

    that the

    T

    splitting

    at

    k

    ex

    =

    0 is proportional

    to

    the

    oscillator

    strength per unit volume see Eq. 68)): for a single QW, the dimensionless oscillator

    strength is proportional

    to the

    area

    and

    the oscillator strength per

    unit

    volume vanishes

    [26].

    The situation should be different in the case of multiple quantum wells, since

    the

    oscillator strength per unit volume is finite and a nonvanishing T splitting between

    rs excitons

    at

    kex=O is expected to occur [104].

    t

    can be seen from Fig. 13 that

    the

    z-polarized r4) light-hole exciton at k

    ex

    = 0 is

    shifted upwards with respect

    to the

    rs exciton by an energy of the order of a meV. Such

    a splitting, which is expected from symmetry arguments, arises from the depolarization

    field for a polarization perpendicular to

    the

    interfaces. The

    ZT

    splitting of quantum well

    90

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    35/56

    x=O.4

    1.0

    LH exciton

    g

    >

    r

    .5

    L

    0.0

    T

    0

    10 15

    kL

    Figure 13.

    Long-range exchange energies

    of the L,

    and

    Z

    modes for

    the

    ground-state

    LHI-CBl

    excitons in a

    60

    A

    wide GaAs/Alo.4Gao.6As

    quantum

    well. From Ref.

    [26].

    excitons has been observed experimentally [105, 106].

    Other

    features

    of the k-dependent

    exchange interaction

    are

    much more difficult to observe, since they are overwhelmed by

    spatial

    dispersion.

    t

    has already been mentioned in Sec. 2.3 that taking into account

    the nonanalytic

    part of the exchange interaction corresponds to including the depolarization field in the

    dielectric response, i.e, to solve Maxwell equations in the instantaneous limit

    [1,

    29].

    Thus it

    is no surprise that

    the

    dispersion of

    the

    exciton taking into account

    the

    dependent

    exchange interaction coincides with

    that

    found from

    the

    solution of

    the

    electrostatic

    equations [107].

    3.3 Polaritons

    in

    Confined

    Systems

    Division into non

    radiative

    and

    radiative

    modes. Polariton effects are defined

    to

    be

    the

    effects coming from

    the

    interaction between excitons

    and the retarded

    trans

    verse)

    part

    of

    the

    electromagnetic field.

    There

    is a

    fundamental

    difference between

    polariton

    effects in bulk

    and

    in confined systems see Fig. 14).

    In

    bulk crystals, due

    to

    the

    conservation of crystal

    momentum

    an exciton with a given wavevector k

    ex

    interacts

    with

    only one

    photon

    with

    the

    same wavevector

    and

    polarization: there is no density

    of states

    for radiative decay, and

    the

    mutual

    interaction

    gives rise

    to

    the

    stationary

    po

    lariton states. In quasi-two-dimensional systems, due to the breaking of translational

    invariance along

    the

    growth direction, an exciton with in-plane wavevector kll interacts

    with

    photons with the same in-plane wavevector

    but

    with all possible values of k

    z

    Thus

    there is a density of states for radiative decay, given by

    ~ V n 2 i w

    p(kll w) == L,o(/iw - k ~ + kn = sCi: ~ B k o -

    kll ,

    kz 7r C V o

    -

    ll

    108)

    where

    n

    is

    the

    index

    of refraction,

    o

    =

    nw/

    c is

    the

    wavevector of light

    in

    the

    sample,

    and

    B a::)

    is

    the

    Heavyside function

    B a::)

    =

    1 for

    a

    >

    0,

    B a::)

    =

    0 for

    a

    .(z)p>.(z ),

    >

    110)

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    37/56

    where A are

    quantum numbers

    of

    the

    excited

    states

    of

    the

    crystal . We consider only

    the

    resonant

    terms, and

    incorporate

    the

    contribution of all nonresonant

    terms

    in

    the

    background dielectric constant. For excitons in thin layers, we have

    k

    _ g.xl u

    c

    l

    er

    lu

    v

    } 2

    X x W - . )

    Ii Wk

    - W - 2

    111)

    p.x z) = F.x p =

    O,z,z), 112)

    where F.x p, Ze Zh) is

    the

    exciton envelope function.

    Thus the

    semiclassical description

    of polariton effects in thin layers consists of the following three steps [111, 92, 112]:

    1.

    Determine the

    excitonic levels

    liw.x and

    wavefunctions F.x p,

    Ze,

    Zh)

    by

    solving

    the

    Schrodinger equation for

    the

    exciton;

    2. Identify

    the

    resonant terms

    and

    calculate

    the

    nonlocal susceptibili ty 111);

    3. Solve Maxwell

    equations

    with

    the

    constitut ive relation 109).

    Before presenting a few applications, let us discuss some features of

    the

    nonlocal theory.

    III The formalism can be applied to both the thin-film and quantum-well regimes for

    the

    exciton. For

    QWs

    only one resonant level needs

    to be

    considered, while for

    thin

    films one has

    to sum

    over

    the

    quantized levels of

    the

    center of mass.

    III The

    expression 110) has the important feature that each

    term

    in

    the

    sum is

    separable in z, z .

    This

    allows Maxwell equations to be solved analytically in

    many

    cases [111].

    III

    In the

    thin-film regime,

    the

    formalism is intrinsically ABC-free. Additional bound

    ary

    conditions are embodied in

    the

    no-escape boundary conditions for electrons

    and

    holes separately. Also,

    the

    formalism can account for

    phenomena

    related

    to

    the distortion of

    the

    excitonic wavefunction close to the surface, like a decrease

    of

    the

    dead-layer

    depth

    for

    thin

    layers.

    III The

    formalism takes full account of polariton effects, including the radiative broad-

    ening i.e.

    both

    real and imaginary parts of

    the

    self-energy are found).

    In

    expres

    sion 111), , is a

    nonradiative

    broadening

    term: the

    radiative width arises auto

    matically from

    the

    solution of Maxwell equations with retardation,

    and

    manifests

    itself as

    an

    additional broadening

    of

    absorption or reflectivity peaks.

    At

    this point we should mention that an alternative procedure has been developed,

    known as Stahl's coherent wave approach [41] in which

    the

    Schrodinger equation for

    the

    excitons is formulated in a density-matrix scheme

    and

    is solved in

    the

    presence of

    the

    electromagnetic field.

    Stahl's

    approach gives equivalent results for linear propert ies,

    but

    can also be applied

    to

    nonlinear phenomena and

    to study

    polariton effects at frequencies

    above the band edge.

    Polariton effects in thin

    films:

    polariton

    interference

    vs.

    e quantization.

    Interference of polaritons was observed

    in

    CdS

    and

    CdSe slabs [55],

    in

    CuCI slabs [113],

    and, most

    recently, in GaAs

    thin

    layers [114]. An example

    is

    given

    in

    Fig. 15.

    The traditional explanation

    makes use of

    the

    dielectric function with

    spatial

    dis

    persion, which gives the two polariton branches

    j

    = 1,2 as the solutions of

    c

    2

    k

    _ [ WLTWT ]

    - Eoo 2

    w

    2

    Wk -

    w

    2

    - i w

    J

    113)

    93

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    38/56

    I

    r

    I gl h

    b

    ll

    I I I

    t;

    I

    I

    n

    F

    a

    I

    1514 5 6 1518

    152 1522

    photon n rgy

    me-V)

    Figure

    15.

    Reflectance

    spectrum

    of

    a

    MBE

    grown

    GaAs

    thin

    layer,

    in

    the

    spectral

    region

    of the

    exciton resonance. From Ref. [114].

    The observed interference pattern can be explained as follows

    [55J:

    below the transverse

    frequency, a single period is observed, which corresponds to

    Fabry-Perot

    interference

    of

    the lower polariton according to the condition

    2klL

    = 27rm

    L

    is the

    width

    of the

    slab). The period doubling above WL is interpreted as interference between UP and LP,

    described by the condition (k

    1

    -

    k2 L

    = 27rm. Moreover, a longer period appear which

    corresponds to

    interference

    of the UP

    with itself.

    The

    reflectivity

    can be

    calculated

    imposing additional

    boundary

    conditions like Pekar s ABC

    I i

    Pi

    =

    0

    at the

    surfaces.

    For GaAs, the interference pattern is more complicated, due to the presence of two

    exciton-polariton branches (heavy and light) [114J.

    A

    related

    phenomenon which occurs in very

    thin

    layers is quantization of the center

    of

    mass

    of

    the exciton, which can be observed in luminescence and reflectivity [115, 116,

    117J. Assuming

    an

    excitonic wavefunction with CM confinement, selection rules for

    optical transitions can

    be

    derived, which are

    in

    reasonable agreement with

    experiment

    [115J.

    The two

    phenomena

    can

    be treated in

    a unified way using

    the

    nonlocal scheme

    described above. Numerical calculations of absorption and reflectivity profiles [112, 92J

    show

    that

    the

    conditions of polariton interference are recovered for frequencies

    not

    too close

    to the

    resonance region.

    When the

    polariton

    is

    excitonlike,

    the

    interference

    condition

    2kL = 27rm

    corresponds

    to the

    quantization condition

    k = 7r

    /

    L)m

    for

    the

    center-of-mass wavevector: this is the case for the LP for

    k

    ka and for

    the UP

    close

    to the

    longitudinal frequency. For very

    thin

    layers,

    the

    quantized wavevectors are

    too

    large

    to

    involve

    the

    UP. Also, as we mentioned in Sec. 3.2, quantization

    of the

    center

    of mass is only approximately described by

    the

    condition

    k = 7r

    / Lerr

    m, and the

    more

    precise quantization condition (102)

    must be

    used.

    In the

    resonance region between

    WT

    and WL

    the

    positions of absorption and reflectivity peaks

    cannot in

    general be simply

    explained

    by

    any of the two pictures; instead the full nonlocal theory must

    be

    used.

    Polariton effects manifest themselves

    both

    as shifts with respect

    to the quantized

    CM

    levels,

    and

    as additional broadenings of the peaks

    due

    to

    the

    radiative width.

    The

    nonlocal approach has been developed

    so

    far only for

    the

    case of

    normal

    inci

    dence. Polariton dispersion in

    thin

    films has been studied

    in

    Ref.

    [118J

    using a macro

    scopic

    theory

    with additional

    boundary

    conditions. The dispersion of

    stationary

    po-

    94

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    39/56

    lariton states

    has

    been

    calculated for

    both T-modes

    s-polarization)

    and L, Z

    modes

    p-polarization).

    The

    case of

    p

    polarization is more complicated, since

    the

    longitudinal

    mode

    is also excited

    in

    layers with L

    A

    and

    therefore

    the

    ABC have two components.

    We refer

    to

    [118,

    36]

    for details.

    Polaritons

    in

    quantum

    wells:

    nonradiative

    modes.

    In

    the

    quantum-well limit,

    the nonlocal formalism can be used

    to

    find the dispersion of nonradiative waveguide)

    modes by taking only one resonant

    term

    in

    the

    nonlocal susceptibility [26, 107, 119]

    the

    full quantum-mechanical theory is developed in Ref. [120]).

    The quantity

    p,\ z) =

    FQw O)c z)v z), where

    FQw p)

    is

    the

    in-plane electron-hole wavefunction. We take a

    propagation vector

    k

    =

    kxx

    along

    the

    x

    axis. We enclose

    the quantum

    well

    in

    a large

    box

    of width

    d,

    such that

    the

    response function is essentially zero for Izl

    Iz l > d/2.

    We

    assume an electric field which decays exponentially far from

    the

    well:

    114)

    where

    the

    decay

    constant

    is given by

    2

    _

    k

    2

    W

    )1/2

    a - x - Eoo .

    c

    115)

    Here we give

    an example of the

    solution of Maxwell equations for

    the T-mode

    T

    E,

    or

    s-polarization).

    The

    electric

    and

    magnetic fields are given by

    E

    B

    Maxwell equations imply

    Since for TE modes V . E

    =

    0

    the

    above equation gives, for

    Izl d/2,

    116)

    117)

    118)

    Maxwell

    boundary

    conditions require

    that Ell Ell

    are continuous

    at z = d/2:

    from 116),

    BEy/Bz is also continuous. Hence boundary conditions can be conveniently expressed

    in terms of E

    y

    , yielding

    119)

    t can

    be shown by

    direct substitution that

    a solution of Eq. 118) with

    boundary

    conditions 119) is

    Ey(z) =

    J

    z p(z )e-alz-z l.

    120)

    Substituting in

    118), we obtain

    the

    dispersion relation of

    T-mode

    polaritons in

    the

    form

    121)

    95

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    40/56

    1623

    L=60 A x=0 4

    51622

    ko the effects of retardation are negligible,

    and

    the

    dispersion of

    the

    exciton with the long-range exchange interaction Fig. 13) is

    recovered.

    Luminescence from the L-polarized QW excitonic polariton has been reported in

    Ref. [121), where coupling to the nonradiative modes

    is

    achieved by use of a grating.

    96

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    41/56

    Possible evidence for

    QW

    excitonic polaritons is given

    by the

    time-of-flight experiment

    of Ref.

    [122J,

    in which an increase of the delay

    time

    of a light pulse

    propagating

    in the

    layer planes

    is

    observed close to

    the

    exciton resonance. This could represent the analog

    of

    the

    classical group-velocity experiments

    [57J.

    I t is of interest to investigate to which extent the quantum well can be described

    within

    a local scheme with

    an

    effective dielectric function,

    WLT

    f W)

    = foo[1

    .

    J

    a -

    W

    -

    Z Y

    126)

    where WLT is a parameter which can be called effective LT splitt ing .

    The

    relation

    between

    the two approaches can be

    studied

    for the normal-incidence reflectivity. Within

    the nonlocal scheme, the reflectivity of an isolated

    quantum

    well

    surrounded

    by infinite

    barriers

    is found to

    be

    [123J

    r

    2

    R w)

    =

    a ,

    w

    -

    wa)2

    T

    ra)2

    127)

    where

    ra = ~ ~ x y

    nmaC S

    128)

    will be

    interpreted

    in

    the

    next section

    to

    be

    the

    decay rate of the electric field. Within

    the local scheme, the reflectivity of a layer of width d is calculated

    to

    be

    R w) = 1 f W) -1 1- e

    2iqd

    ) 12

    1 Vf W))2

    -

    1 -

    VE w))2e

    2iqd

    129)

    where q = V

    W w

    / c.

    The

    two approaches give identical results

    under the

    condition

    qd

    1,

    which can be written as

    [124J

    d)2

    EooWLT

    \ 1 1 1.

    A a

    -

    W

    -

    Z Y

    130)

    I f this condition is satisfied, the relation between

    the

    phenomenological

    parameter

    WLT

    and

    the

    microscopic parameters is WLT = r

    a

    / k

    a

    d . For a single

    quantum

    well, d

    coincides with

    the

    well

    width L

    w

    ,

    while for a multiple

    quantum

    well

    d must

    be identified

    with

    the MQW

    period

    Lw+Lb

    [123, 124J. A sufficient condition for

    130) to

    hold is that

    the

    width Y

    be much larger than

    the

    effective LT splitting

    [123J.

    This is usually verified

    experimentally. The parameter WLT can

    be

    expressed as

    the

    bulk LT splitting, multiplied

    by an enhancement factor related

    to

    excitonic confinement. However we

    remark

    that

    for a single quantum well WLT is not

    the

    exchange splitting between longitudinal and

    transverse

    rs excitons, which, as

    we

    have seen, vanishes for k

    ex

    ---t

    O. For

    MQWs

    on the other hand WLT is

    expected

    to coincide with

    the

    finite exchange splitting at

    k/l

    =

    O.

    A thorough comparison between local

    and

    nonlocal schemes for what concerns

    the

    exciton-polariton dispersion

    is

    done in Ref.

    [124J,

    also for the case

    of

    MQWs where

    it

    is shown that the condition for

    the

    validity of

    the

    local description is given again by

    Eq.

    130).

    The nonlocal scheme

    can be

    used to calculate the reflectivity and attenuated total

    reflection of a quantum well. Such a calculation

    is

    done in Ref. [125J for both

    sand

    p

    polarizations. Excitonic polaritons in

    the

    radiative region

    appear

    as peaks or dips in

    the reflectivity, while nonradiative polaritons manifest themselves as dips in

    the

    ATR

    spectrum

    [125J.

    In Ref. [126J the nonlocal formalism

    is

    applied to a calculation of the

    reflectivity in quantum wells, quantum wire, and quantum dot structures.

    97

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    42/56

    0.2

    0.1

    o 3

    L

    Ie)

    Figure 17. Radiative decay rate of the k =

    a

    exciton as a function of crystal thickness.

    From

    Ref.

    [130]

    3.4 Exciton Radiative Lifetime: from

    Bulk

    to Confined Systems

    Crossover

    2D--t3D:

    overview. Excitons in

    thin

    layers with thickness

    ,\

    have

    a finite radiative lifetime, as was first shown in Ref. [108]. For Frenkel excitons,

    the

    decay

    rate is

    of

    the order

    of

    r

    \2/a

    rmob where

    a is the

    lattice spacing

    and rmol

    is

    the

    decay

    rate

    of an isolated molecule.

    The

    decay

    is

    called

    superradiant ,

    because

    the

    decay

    rate is proportional to the

    number of molecules within a wavelength, which

    contribute

    in phase

    to the

    decay [127].

    The

    superradiant

    decay of excitons

    in

    anthracene

    films was observed

    in

    Ref.

    [128].

    For Wannier-Mott excitons,

    the

    decay

    is

    still often

    called

    superradiant, although

    it

    is not

    meaningful anymore

    to

    speak

    of

    an enhancement

    with respect

    to the

    isolated molecule. I prefer

    the

    picture of Fig. 14, according

    to

    which

    the

    intrinsic radiative decay of a free exciton in low-dimensional systems

    is

    due

    to

    coupling with a continuum of photon states.

    For excitons in

    thin

    layers,

    the

    superradiant decay

    must go

    over

    to the

    stable

    polariton behavior as

    the

    film thickness becomes

    >.. The

    behavior of

    the

    decay

    rate

    as a function of thickness

    is

    calculated in Refs.

    [129, 130]

    (see Fig.

    17,

    which is

    calculated for Frenkel excitons). For

    L

    \,

    the

    decay

    rate

    increases as

    r

    )

    L:

    for

    Wannier-Mott excitons, this corresponds

    to the

    regime where

    the

    center of

    mass

    of

    the

    exciton is quantized,

    and the

    oscillator

    strength

    per

    unit

    area increases linearly with

    the

    well thickness (see Fig. 11). For >.

    the

    decay

    rate

    has an oscillatory behavior and

    decreases as

    1/

    due

    to the

    reduced overlap between exciton

    and photon

    wavefunctions.

    In fact, for

    \

    the

    physical picture is expected

    to

    coincide with

    the

    bulk polariton

    behavior, for which the decay

    rate

    is an escape rate which goes as 1/ (see Eq. (85)).

    This

    crossover from superradiant decay

    to

    bulk polaritons is therefore well described

    by

    the

    results of Refs.

    [129, 130].

    In Ref. [130]

    it

    is also shown that

    the

    bulk polariton

    behavior is

    obtained at

    all wavevectors, when proper approximations

    to

    the

    radiative

    self-energy are

    made.

    The

    QW

    regime

    as

    \,

    where

    the

    oscillator

    strength (and

    therefore

    the

    decay

    rate)

    increases as

    the

    well width is reduced,

    is not

    described in Refs.

    [129, 130].

    98

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    43/56

    Radiative lifetime

    of

    QW excitons. Now we calculate

    the

    radiative decay

    rate

    of a free exciton in an isolated QW,

    under

    the assumption that

    the

    in-plane wavevector

    kll

    = kxx is conserved [131, 119, 123, 132]. We use the Golden Rule (20), with the

    interaction

    Hamiltonian being given by

    (21). The

    initial

    state

    consists of

    the

    exciton

    state with polarization vector e

    and

    no photons present, while

    the

    final

    state

    is

    the

    crystal

    ground

    state

    plus a

    photon

    with wavevector k

    =

    (qll, kz)

    and

    polarization

    >..

    Expressing

    the

    vector

    potential in

    second quantization according

    to (81),

    the

    matrix

    element is calculated

    to

    be

    (131)

    The

    squared

    matrix

    element summed over

    the

    photon polarizations can be expressed

    in terms of

    the

    exciton oscillator strength as

    2

    11 e

    2

    li,2

    LI(il.C fW = 2 -L E l e . g A )1

    2

    A

    n ma

    V

    A

    (132)

    Now

    the

    Golden Rule gives

    (133)

    Evaluating the

    one-dimensional density of states as

    in (108)

    gives

    the

    decay

    rate in

    terms

    of

    the

    oscillator

    strength

    per

    unit area

    as

    r kx) = 211

    ka

    L E Ie. g A)1

    2

    0(k

    a

    - k

    x

    ,

    n

    mackz

    A

    S

    (134)

    where

    now

    kz = Jk6 - ki.

    We must now specify

    the

    exciton polarization vector e

    and

    the photon

    polarization vectors g A). For a given in-plane wavevector

    kll

    = kxx,

    the

    two

    orthogonal

    photon

    polarization vectors can be chosen as

    (135)

    For the T-exciton, I:A f,le. g A)1

    2

    = fxy, where fxy is the oscillator strength for in-plane

    polarization.

    2

    For

    the

    L-exciton, we

    obtain

    I:A

    Ele.

    g A)1

    2

    =

    fxy(kz

    /k

    a

    )2,

    while for

    the

    Z-exciton

    I:A Ele .

    g A)1

    2

    =

    fz k

    x

    /k

    a

    2.

    Thus we obtain the

    radiative widths

    of the

    T

    L, and Z excitons for kx ka as follows:

    211 e

    2

    fxy

    ko

    n mac S k

    z

    211 e

    2

    fxu kz

    ---:;;: mac

    o

    211 e

    2

    fz k;

    n

    mac

    S

    kokz

    136)

    (the

    factor of two was missing in Ref.

    [123]). The

    decay

    rate

    vanishes for kx

    >

    ko For

    the

    light-hole exciton

    fxy =

    4fz. Optically inactive

    states

    obviously have zero radiative

    width.

    2Note

    that

    for a given exciton polarization vector

    e, the

    oscillator

    strength

    can

    be thought to be

    summed over all possible polarizations, since only the vector i =

    e

    contributes. Thus the oscillator

    strengths

    fxy, fz are

    exactly those calculated e.g. in Ref.

    [73], and

    reported in Fig. 10.

    99

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    44/56

    At kll

    0, Land

    T

    modes have

    the

    same decay

    rate 2ro

    27re2/ nmoc)) fxlJ/S).

    Taking an oscillator strength

    fx,,/ S

    = 50

    .10-

    5

    A 2 which is

    appropriate

    for

    the

    HH1-

    CB1 exciton in

    GaAs/

    AlxGal_xAs quantum wells of

    about

    100 Awidth [73], we obtain

    firo 0.026 meV.

    The

    decay time of an exciton

    state

    is

    TO 1/ 2r

    o

    ) 12 ps [132].

    This coincides with the result found from

    the

    reflectivity calculation

    with the

    nonlocal

    susceptibility Sec. 3.3):

    the

    decay

    time

    of

    the

    electric field is

    l / r

    o, while

    the

    decay

    time

    of

    the

    intensity is twice as short. Note that

    the

    Golden Rule gives directly the decay

    rate

    of

    the

    exciton

    state.

    This lifetime is much longer than found in Ref. [131], where

    the

    index of refraction is not considered

    and

    the two-dimensional limit for the exciton

    is assumed. Observation of such short lifetimes has been

    reported

    in Refs. [133, 134].

    The

    radiative

    width

    136) represents

    the

    imaginary part of the exciton self-energy

    due

    to interaction with

    photons.

    The

    real

    part of the

    self-energy, calculated e.g.

    in

    Refs.

    [131, 119, 132], is a small effect.

    The largest polariton effect or quantum-well excitons

    is the radiative lifetime.

    Note that

    the

    decay

    rate

    diverges as

    kx

    - - t

    o

    for the

    T

    and

    Z

    modes.

    This

    divergence, which can be

    traced

    back

    to

    the

    density

    of

    states 108), is

    integrable and disappears when

    the

    thermal average is taken. A slight broadening of

    the

    wavevector due

    to

    inelastic scattering would wash

    out

    this divergence, which we

    believe to have no physical consequences.

    The

    intrinsic

    radiative

    decay of free excitons is calculated assuming conservation

    of

    the

    in-plane wavevector, thereby disregarding

    the

    effects of interface roughness and

    acoustic phonon scattering. Wavevector conservation is likely

    to be

    a good

    assumption

    when the coherence

    length

    of

    the

    exciton is longer than the wavelength of light. Conse

    quently,

    the

    short intrinsic lifetimes can only

    be

    observed in carefully selected samples

    at low

    temperature

    [133, 134].

    In

    general, several effects can change this simple picture.

    At

    low

    temperature,

    excitons can

    be

    bound to

    impurities

    or

    interface defects. Also,

    interface roughness

    acts

    as a disordered

    potential

    for

    the

    exciton motion,

    and

    produces

    a mobility edge within

    the

    inhomogeneously broadened exciton line [135]: below

    the

    mobility edge

    the

    exciton is localized by

    the

    disorder, while above

    the

    mobility edge

    the

    exciton is mobile

    and

    interface roughness acts as a dephasing mechanism. Finally,

    scattering with acoustic phonons has to be taken into account at finite

    temperature.

    The

    interplay

    between all these effects constitutes a complicated problem, which is only

    partly

    understood

    at time of writing. In the following I shall discuss a few models which

    have been proposed [136].

    Effect ofthermalization.

    Thermalization is due

    to

    inelastic scattering

    with

    acous

    tic phonons [53], which changes

    the

    exciton wavevector but does not change the

    total

    exciton population scattering with optical phonons is not expected to be relevant for

    thermal

    energies smaller than

    36

    meV, which is

    the

    LO phonon energy in GaAs).

    The

    key point is comparing the scattering rate with

    the

    radiative lifetime. The scattering

    rate

    by acoustic phonons is measured to be linear in with a coefficient I 5 pe V /

    K

    for a QW of 135 Awidth [53] Thus thermalization processes are faster than radiative

    decay for

    >

    10

    K. There

    is also

    other

    evidence for this conclusion, coming from

    a high-energy Boltzmann tail of

    the

    exciton lines [137].

    Thus it

    can be assumed as a

    working

    assumption

    that

    thermalization processes are faster

    than

    radiative recombina

    tion, i.e., that excitons always have a thermal distribution while they decay radiatively.

    However the possibility of a failure of this assumption must be kept in mind, particu

    larly for thermalization between dipole-allowed

    and

    triplet exciton states see Fig. 12),

    since spin-flip scattering is likely

    to

    be slower. Also, having a thermalized distribution

    depends

    on the

    conditions of excitation resonant or nonresonant)

    and

    on

    the

    time of

    observation [134].

    100

  • 7/21/2019 L. C. Andreani, OPTICAL TRANSITIONS, EXCITONS, AND POLARITONS IN BULK AND LOW-DIMENSIONAL SEMICONDU

    45/56

    In the

    assumption

    of a rapid thermalization,

    the

    decay

    rate

    of

    the

    luminescence is

    given by the

    thermal

    average of

    the

    decay rate (136).

    The

    two characteristic energies

    are the

    thermal energy kBT, and

    the

    kinetic energy of excitons which decay radiatively:

    the latter is at

    most

    El =

    2

    k5 (2M),

    where

    M is the

    exciton mass. Using

    M =

    0.25

    mo,

    we find

    1i2k6/(2M)

    1.1 K. This means that, for

    T

    1

    K,

    only a small fraction of

    excitons occupy

    the

    states

    with

    kx