Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Quantum Wells, Wires and Dots: Theoretical and Computational Physics of Semiconductor Nanostructures
Quantum Wells, Wires and Dots: Theoretical and Computational Physics of Semiconductor Nanostructures
Quantum Wells, Wires and Dots: Theoretical and Computational Physics of Semiconductor Nanostructures
Ebook1,052 pages8 hours

Quantum Wells, Wires and Dots: Theoretical and Computational Physics of Semiconductor Nanostructures

Rating: 2 out of 5 stars

2/5

()

Read preview

About this ebook

Quantum Wells, Wires and Dots, 3rd Edition is aimed at providing all the essential information, both theoretical and computational, in order that the reader can, starting from essentially nothing, understand how the electronic, optical and transport properties of semiconductor heterostructures are calculated. Completely revised and updated, this text is designed to lead the reader through a series of simple theoretical and computational implementations, and slowly build from solid foundations, to a level where the reader can begin to initiate theoretical investigations or explanations of their own.
LanguageEnglish
PublisherWiley
Release dateSep 26, 2011
ISBN9781119964759
Quantum Wells, Wires and Dots: Theoretical and Computational Physics of Semiconductor Nanostructures
Author

Paul Harrison

Paul Harrison is a UK-based writer and editor of fiction and nonfiction books for children.

Read more from Paul Harrison

Related to Quantum Wells, Wires and Dots

Related ebooks

Physics For You

View More

Related articles

Reviews for Quantum Wells, Wires and Dots

Rating: 2 out of 5 stars
2/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Quantum Wells, Wires and Dots - Paul Harrison

    ABOUT THE AUTHOR(S)

    Paul Harrison is currently working in the Institute of Microwaves and Photonics (IMP), which is a research institute within the School of Electronic and Electrical Engineering at the University of Leeds in the United Kingdom. He can always be found on the web, at the time of writing, at:

    http://www.ee.leeds.ac.uk/homes/ph/

    and always answers e-mail. Currently he can be reached at:

    p.harrison@leeds.ac.uk or p.harrison@physics.org

    Paul is working on a wide variety of projects, most of which centre around exploiting quantum mechanics for the creation of novel opto-electronic devices, largely, but not exclusively, in semiconductor Quantum Wells, Wires and Dots. Up-to-date information can be found on his web page. He is always looking for exceptionally well-qualified and motivated students to study for a PhD degree with him—if interested, please don’t hesitate to contact him.

    Zoran Ikonic was a Professor at the University of Belgrade and is now also a researcher in the IMP. His research interests and experience include the full width of semiconductor physics and optoelectronic devices, in particular, band structure calculations, strain-layered systems, carrier scattering theory, non-linear optics, as well as conventional and quantum mechanical methods for device optimization.

    Vladimir Jovanovic completed his PhD at the IMP on physical models of quantum well infrared photodetectors and quantum cascade lasers in GaN- and GaAs-based materials for near-, mid- and far-infrared (terahertz) applications.

    Marco Califano is a Royal Society University Research Fellow based in the IMP at Leeds whose main interests focus on atomistic pseudopotential modelling of the elec­tronic and optical properties of semiconductor nanostructures of different materials for applications in photovoltaics

    Craig A. Evans completed his PhD on the optical and thermal properties of quantum cascade lasers in the School of Electronic and Electrical Engineering, University of Leeds in 2008. He then worked as a Postdoctoral Research Assistant in the IMP working in the field of rare-earth doped fibre lasers and integrated photonic device modelling and has now joined the staff of the school.

    Dragan Indjin is an Academic Research Fellow in the IMP and has research interests in semiconductor nanostructures, non-linear optics, quantum computing and spintronics.

    ABOUT THE BOOK

    99.5% of this book was produced with ‘open source’ not-for-profit software. The text was prepared with LATEX2ε using Wiley’s own style (class) files. It was input by hand initially with the aid of the excellent ‘VI Like Emacs’ (vile) and then with the superb ‘Vi IMproved’ (vim/gvim). The schematic diagrams were prepared using ‘xfig’, the x-y plots with ‘xmgr/xmgrace’ and three-dimensional molecular models with ‘RasMol’. ‘BibView’ was used to maintain a BIBTEX database in the earlier editions. Manuscript preparation has been under all three of the most popular operating systems by now!

    Further information about the book, including errata and the software from the first edition is available on the book’s web page, which is currently:

    http://www.imp.leeds.ac.uk/qwwad/

    INTRODUCTION

    Since their discovery/invention by Esaki and Tsu in the 1970s, semiconductor quantum wells and superlattices have evolved from scientific curiosities to a means of probing the fundamentals of quantum mechanics, and more recently into wealth-creating semicon­ductor devices.

    In this work, a brief resume of quantum theory and solid state physics is given before launching into the main body of the book—the theoretical and computational framework of semiconductor heterostructures. The first chapter introduces the concepts of effective mass and envelope function approximations, which are two cornerstone theories, from a quite different perspective. Usually these two techniques are introduced from rigorous mathematics with some approximations—see, for example, the works by Bastard and Burt. The motivation behind this approach is to introduce these concepts from very simplistic intuitive, and at times, graphical arguments. The range of validity of such approximations is not being challenged; they are good theories and used many times in this book. This present approach is merely to reinforce these ideas in ‘pictures rather than words’.

    The aim of this book is to provide solid foundations in the theoretical methods neces­sary for calculating some of the basic electronic and optical properties of semiconductor quantum wells, wires and dots. Some background knowledge will be required; I often cite Ashcroft and Mermin [1] and Blakemore [2] for support material in solid state physics, and works such as Eisberg [3] and Weidner and Sells [4] for quantum theory. This present treatise should be considered to complement existing books in the field. I thoroughly recommend the books by Jaros [5], Davies and Long [6], Kelly [7], Turton [8], Ivchenko and Pikus [9], Shik [10], and Basu [11]. These texts will provide the literature reviews, and descriptive introductions, and also, in some cases, detailed theoretical treatments.

    CHAPTER 1

    SEMICONDUCTORS AND HETEROSTRUCTURES

    1.1 THE MECHANICS OF WAVES

    De Broglie (see reference [4]stated that a particle of momentum p has an associated wave of wavelength λ given by the following

    (1.1) c01_img01.jpg

    Thus, an electron in a vacuum at a position r and away from the influence of any electromagnetic potentials, could be described by a state function , which is of the form

    (1.2) c01_img02 .jpg

    where t is the time, ω the angular frequency and the modulus of the wave vector is given by:

    (1.3) c01_img03.jpg

    The quantum mechanical momentum has been deduced to be a linear operator [12] acting upon the wave function ψ , with the momentum p arising as an eigenvalue, i.e.

    (1.4) c01_img04.jpg

    where

    (1.5) c01_img05.jpg

    which, when operating on the electron vacuum wave function in equation (1.2), would give the following:

    (1.6) c01_img06.jpg

    and therefore

    (1.7)

    c01_img07.jpg

    (1.8)

    c01_img08.jpg

    Thus the eigenvalue:

    (1.9) c01_img09.jpg

    which not surprisingly can be simply manipulated c01_img70.jpg to reproduce de Broglie’s relationship in equation (1.1).

    Following on from this, classical mechanics gives the kinetic energy of a particle of mass m as

    (1.10) c01_img10.jpg

    Therefore it may be expected that the quantum mechanical analogy can also be represented by an eigenvalue equation with an operator:

    (1.11) c01_img11.jpg

    i.e.

    (1.12) c01_img12.jpg

    where T is the kinetic energy eigenvalue, and given the form of V in equation (1.5) then:

    (1.13) c01_img13.jpg

    When acting upon the electron vacuum wave function, i.e.

    (1.14) c01_img14.jpg

    then

    (1.15)

    c01_img15.jpg

    Thus the kinetic energy eigenvalue is given by:

    (1.16) c01_img16.jpg

    For an electron in a vacuum away from the influence of electromagnetic fields, then the total energy E is just the kinetic energy T. Thus the dispersion or energy versus momentum (which is proportional to the wave vector k) curves are parabolic, just as for classical free particles, as illustrated in Fig. 1.1.

    Figure 1.1 The energy versus wave vector (proportional to momentumcurve for an electron in a vacuum

    c01_img17.jpg

    The equation describing the total energy of a particle in this wave description is called the time-independent Schrödinger equation and, for this case with only a kinetic energy contribution, can be summarized as follows:

    (1.17) c01_img18.jpg

    A corresponding equation also exists that includes the time-dependency explicitly; this is obtained by operating on the wave function by the linear operator i himg.jpg ∂/∂t, i.e.

    (1.18)

    c01_img19.jpg

    i.e.

    (1.19) c01_img20.jpg

    Clearly, this eigenvalue c01_img71.jpg is also the total energy but in a form usually associated with waves, e.g. a photon. These two operations on the wave function represent the two complimentary descriptions associated with wave–particle duality. Thus the second, i.e., time-dependent, Schrödinger equation is given by the following:

    (1.20) c01_img21.jpg

    1.2 CRYSTAL STRUCTURE

    The vast majority of the mainstream semiconductors have a face-centred cubic Bravais lattice, as illustrated in Fig 1.2. The lattice points are defined in terms of linear combinations of a set of primitive lattice vectors , one choice for which is:

    (1.21)

    c01_img22.jpg

    The lattice vectors then follow as the set of vectors:

    (1.22) c01_img23.jpg

    Figure 1.2 The face-centred cubic Bravais lattice

    c01_img24.jpg

    where α1, α2 and α3 are integers.

    The complete crystal structure is obtained by placing the atomic basis at each Bravais lattice point. For materials such as Si, Ge, GaAs, AlAs, InP, etc., this consists of two atoms, one at ( c01_img25.jpg ) and the second at ( c01_img26.jpg ), in units of A0.

    Figure 1.3 The diamond (left) and zinc blende (right) crystal structures

    c01_img27.jpg

    For the group IV materials, such as Si and Ge, as the atoms within the basis are the same then the crystal structure is equivalent to diamond (see Fig. 1.3 (left)). For III–V and II–VI compound semiconductors such as GaAs, AlAs, InP, HgTe and CdTe, the cation sits on the ( c01_img26.jpg ) site and the anion on ( c01_img28.jpg ); this type of crystal is called the zinc blende structure, after ZnS, see Fig 1.3 (right). The only exception to this rule is GaN, and its important Inx Ga1–xN alloys, which have risen to prominence in recent years due to their use in green and blue light emitting diodes and lasers (see for example [13]) these materials have the wurtzite structure (see reference [2] p. 47).

    Figure 1.4 Schematic illustration of the ionic core component of the crystal potential across the {001} planes—a three-dimensional array of spherically symmetric potentials

    c01_img29.jpg

    From an electrostatics viewpoint, the crystal potential consists of a three-dimensional lattice of spherically symmetric ionic core potentials screened by the inner shell electrons (see Fig. 1.4), which are further surrounded by the covalent bond charge distributions that hold everything together.

    1.3 THE EFFECTIVE MASS APPROXIMATION

    Therefore, the crystal potential is complex; however using the principle of simplicity∗ imagine that it can be approximated by a constant! Then the Schrödinger equation derived for an electron in a vacuum would be applicable. Clearly though, a crystal isn’t a vacuum so allow the introduction of an empirical fitting parameter called the effective mass, m ∗. Thus the time-independent Schrödinger equation becomes:

    (1.23) c01_img30.jpg

    and the energy solutions follow as:

    (1.24) c01_img31.jpg

    This is known as the effective mass approximation and has been found to be very suitable for relatively low electron momenta as occur with low electric fields. Indeed, it is the most widely used parameterisation in semiconductor physics (see any good solid state physics book, e.g. [1,2,7]). Experimental measurements of the effective mass have revealed it to be anisotropic—as might be expected since the crystal potential along say the [001] axis is different than along the [111] axis. Adachi [14] collates reported values for GaAs and its alloys; the effective mass in other materials can be found in Landolt and Bornstein [15].

    In GaAs, the reported effective mass is around 0.067 m0, where m0 is the rest mass of an electron. Fig. 1.5 plots the dispersion curve for this effective mass, in comparison with that of an electron in a vacuum.

    Figure 1.5 The energy versus wave vector (proportional to momentumcurves for an electron in GaAs compared to that in a vacuum

    c01_img32.jpg

    1.4 BAND THEORY

    It has also been found from experiment that there are two distinct energy bands within semiconductors. The lower band is almost full of electrons and can conduct by the movement of the empty states. This band originates from the valence electron states which constitute the covalent bonds holding the atoms together in the crystal. In many ways, electric charge in a solid resembles a fluid, and the analogy for this band, labelled the valence band is that the empty states behave like bubbles within the fluid—hence their name holes.

    In particular, the holes rise to the uppermost point of the valence band and just as it is possible to consider the release of carbon dioxide through the motion of beer in a glass, it is actually easier to study the motion of the bubble (the absence of beer), or in this case the motion of the hole.

    In a semiconductor, the upper band is almost devoid of electrons. It represents excited electron states which are occupied by electrons promoted from localized covalent bonds into extended states in the body of the crystal. Such electrons are readily accelerated by an applied electric field and contribute to current flow. This band is therefore known as the conduction band.

    Fig. 1.6 illustrates these two bands. Notice how the valence band is inverted—this is a reflection of the fact that the ‘bubbles’ rise to the top, i.e. their lowest energy states are at the top of the band. The energy difference between the two bands is known as the bandgap , labelled as Egap on the figure. The particular curvatures used in both bands are indicative of those measured experimentally for GaAs, namely effective masses of around 0.067 m0 for an electron in the conduction band, and 0.6 m0 for a (heavy-)hole in the valence band. The convention is to put the zero of the energy at the top of the valence band. Note the extra qualifier ‘heavy’. In fact, there is more than one valence band, and they are distinguished by their different effective masses. Chapter 15 will discuss band structure in more detail; this will be in the context of a microscopic model of the crystal potential which goes beyond the simple ideas introduced here.

    Figure 1.6 The energy versus wave vector curves for an electron in the conduction band and a hole in the valence band of GaAs

    c01_img33.jpg

    1.5 HETEROJUNCTIONS

    The effective mass approximation is for a bulk crystal, which means the crystal is so large with respect to the scale of an electron wave function that it is effectively infinite. Within the effective mass approximation, the Schrödinger equation has been found to be as follows:

    (1.25) c01_img34.jpg

    When two such materials are placed adjacent to each other to form a heterojunction , then this equation is valid within each, remembering of course that the effective mass could be a function of position. However the bandgaps of the materials can also be different (see Fig. 1.7).

    The discontinuity in either the conduction or the valence band can be represented by a constant potential term. Thus the Schrödinger equation for any one of the bands, taking the effective mass to be the same in each material, would be generalized to:

    (1.26)

    c01_img35.jpg

    Figure 1.7 Two dissimilar semiconductors with different bandgaps joined to form a heterojunction; the curves represent the unrestricted motion parallel to the interface

    c01_img36.jpg

    Figure 1.8 The one-dimensional potentials V (z) in the conduction and valence band as might occur at a heterojunction (marked with a dashed linebetween two dissimilar materials

    c01_img37.jpg

    In the above example, the one-dimensional potentials V (z) representing the band discontinuities at the heterojunction would have the form shown in Fig. 1.8, noting that increasing hole energy in the valence band is measured downwards.

    1.6 HETEROSTRUCTURES

    Heterostructures are formed from multiple heterojunctions, and thus a myriad of possibilities exist. If a thin layer of a narrower-bandgap material ‘A’ say, is sandwiched between two layers of a wider-bandgap material ‘B’, as illustrated in Fig. 1.9 (left) then they form a double heterojunction. If layer ‘A’ is sufficiently thin for quantum properties to be exhibited, then such a band alignment is called a single quantum well.

    If any charge carriers exist in the system, whether thermally produced intrinsic or extrinsic as the result of doping, they will attempt to lower their energies. Hence in this example, any electrons (solid circlesor holes (open circleswill collect in the quantum well (see Fig. 1.9). Additional semiconductor layers can be included in the heterostructure, for example a stepped or asymmetric quantum well can be formed by the inclusion of an alloy between materials A and B, as shown in Fig. 1.9 (right).

    Figure 1.9 The one-dimensional potentials V (z) in the conduction and valence bands for a typical single quantum well (left) and a stepped quantum well (right)

    c01_img38.jpg

    Figure 1.10 The one-dimensional potentials V (z) in the conduction and valence band for typical symmetric (left) and asymmetric (right) double quantum wells

    c01_img39.jpg

    Figure 1.11 The one-dimensional potentials V (z) in the conduction and valence band for a typical multiple quantum well or superlattice

    c01_img40.jpg

    Still more complex structures can be formed, such as symmetric or asymmetric double quantum wells , (see Fig. 1.10 and multiple quantum wells or superlattices (see Fig. 1.11). The difference between the latter is the extent of the interaction between the quantum wells; in particular, a multiple quantum well exhibits the properties of a collection of isolated single quantum wells, whereas in a superlattice the quantum wells do interact. The motivation behind introducing increasingly complicated structures is an attempt to tailor the electronic and optical properties of these materials for exploitation in devices. Perhaps the most complicated layer structure to date is the chirped superlattice active region of a mid-infrared laser [16].

    Figure 1.12 The one-dimensional potentials V (z) in the conduction and valence bands for a typical Type-I single quantum well (left) compared to a Type-II system (right)

    c01_img41.jpg

    All of the structures illustrated so far have been examples of Type-I systems. In this type, the bandgap of one material is nestled entirely within that of the wider-bandgap material. The consequence of this is that any electrons or holes fall into quantum wells which are within the same layer of material. Thus both types of charge carrier are localized in the same region of space, which makes for efficient (fastrecombination. However other possibilities can exist, as illustrated in Fig. 1.12.

    Figure 1.13 The one-dimensional potentials V (z) in the conduction and valence bands for a typical Type-I superlattice (left) compared to a Type-II system (right)

    c01_img42.jpg

    In Type-II systems the bandgaps of the materials, say ‘A’ and ‘C’, are aligned such that the quantum wells formed in the conduction and valence bands are in different materials, as illustrated in Fig. 1.12 (right). This leads to the electrons and holes being confined in different layers of the semiconductor. The consequence of this is that the recombination times of electrons and holes are long.

    1.7 THE ENVELOPE FUNCTION APPROXIMATION

    Two important points have been argued:

    1. The effective mass approximation is a valid description of bulk materials.

    2. Heterojunctions between dissimiliar materials, both of which can be well represented by the effective mass approximation, can be described by a material potential which derives from the difference in the bandgaps.

    The logical extension to point 2 is that the crystal potential of multiple heterojunctions can also be described in this manner, as illustrated extensively in the previous section.

    Once this is accepted, then the electronic structure can be represented by the simple one-dimensional Schrödinger equation that has been aspired to:

    (1.27)

    c01_img43.jpg

    The envelope function approximation is the name given to the mathematical justification for this series of arguments (see for example works by Bastard [17,18] and Burt [19,20]). The name derives from the deduction that physical properties can be derived from the slowly varying envelope function, identified here as ψ(z), rather than the total wave function ψ(z)u(z) where the latter is rapidly varying on the scale of the crystal lattice. The validity of the envelope function approximation is still an active area of research [20]. With the line of reasoning used here, it is clear that the envelope function approximation can be thought of as an approximation on the material and not the quantum mechanics.

    Some thought is enough to appreciate that the envelope function approximation will have limitations, and that these will occur for very thin layers of material. The materials are made of a collection of a large number of atomic potentials, so when a layer becomes thin, these individual potentials will become significant and the global average of representing the crystal potential by a constant will breakdown, for example [21]. However, for the majority of examples this approach works well; this will be demonstrated in later chapters, and, in particular, a detailed comparison with an alternative approach which does account fully for the microscopic crystal potential will be made in Chapter 16.

    1.8 THE RECIPROCAL LATTICE

    For later discussions the concept of the reciprocal lattice needs to be developed. It has already been shown that considering electron wave functions as plane waves (ei k.r), as found in a vacuum, but with a correction factor called the effective mass, is a useful method of approximating the electronic bandstructure. In general, such a wave will not have the periodicity of the crystal lattice; however, for certain wave vectors it will. Such a set of wave vectors G are known as the reciprocal lattice vectors with the set of points mapped out by these primitives known as the reciprocal lattice.

    If the set of vectors G did have the periodicity of the lattice, then this would imply that:

    (1.28) c01_img44.jpg

    i.e. an electron with this wave vector G would have a wave function equal at all points in real space separated by a Bravais lattice vector R. Therefore:

    (1.29) c01_img45.jpg

    which implies that

    (1.30) c01_img46.jpg

    Now learning from the form for the Bravais lattice vectors R given earlier in equation (1.22), it might be expected that the reciprocal lattice vectors G could be constructed in a similar manner from a set of three primitive reciprocal lattice vectors , i.e.

    (1.31)

    c01_img47.jpg

    With these choices then, the primitive reciprocal lattice vectors can be written as follows:

    (1.32) c01_img48.jpg

    (1.33) c01_img49.jpg

    (1.34) c01_img50.jpg

    It is possible to verify that these forms do satisfy equation (1.30):

    (1.35)

    c01_img51.jpg

    Now b1 is perpendicular to both a2 and a3, and so only the product of b1 with a1 is non-zero, and similarly for b2 and b3; hence:

    (1.36)

    c01_img52.jpg

    and in fact, the products biai = 2π ; therefore:

    (1.37)

    c01_img53.jpg

    Clearly β1α1 + β2α2 + β3α3 is an integer, and hence equation (1.30) is satisfied. Using the face-centred cubic lattice vectors defined in equation (1.21), then:

    (1.38)

    c01_img54.jpg

    which gives:

    (1.39)

    c01_img55.jpg

    (1.40)

    c01_img56.jpg

    (1.41) c01_img57.jpg

    Therefore, the first of the primitive reciprocal lattice vectors follows as:

    (1.42)

    c01_img58.jpg

    (1.43)

    c01_img59.jpg

    (1.44) c01_img60.jpg

    A similar calculation of the remaining primitive reciprocal lattice vectors b2 and b3 gives the complete set as follows:

    (1.45)

    c01_img61.jpg

    which are, of course, equivalent to the body-centred cubic Bravais lattice vectors (see reference [1], p. 68). Thus the reciprocal lattice constructed from the linear combinations:

    (1.46)

    c01_img62.jpg

    is a body-centred cubic lattice with lattice constant 4π/A0. Taking the face-centred cubic primitve reciprocal lattice vectors in equation (1.45), then:

    (1.47)

    c01_img63.jpg

    (1.48)

    c01_img64.jpg

    The specific reciprocal lattice vectors are therefore generated by taking different combinations of the integers β1, β2, and β3. This is illustrated in Table 1.1.

    It was shown by von Laue that, when waves in a periodic structure satisfied the following:

    (1.49) c01_img65.jpg

    then diffraction would occur (see reference [1], p. 99). Thus the ‘free’ electron dispersion curves of earlier (Fig. 1.5), will be perturbed when the electron wave vector satisfies equation (1.49). Along the [001] direction, the smallest reciprocal lattice vector G is (0,0,2) (in units of 2π/A0). Substituting into equation (1.49 gives:

    (1.50) c01_img66.jpg

    Table 1.1 Generation of the reciprocal lattice vectors for the face-centred cubic crystal by the systematic selection of the integer coefficients β1, β2, and β3

    c01_img67.jpg

    Figure 1.14 Comparison of the free and nearly free electron models

    c01_img68.jpg

    This then implies the electron will be diffracted when:

    (1.51) c01_img69.jpg

    Fig. 1.14 illustrates the effect that such diffraction would have on the ‘free-electron’ curves. At wave vectors which satisfy von Laue’s condition, the energy bands are disturbed and an energy gap opens. Such an improvement on the parabolic dispersion curves of earlier, is known as the nearly free electron model.

    The space between the lowest wave vector solutions to von Laue’s condition is called the first Brillouin zone. Note that the reciprocal lattice vectors in any particular direction span the Brillouin zone. As mentioned above a face-centred cubic lattice has a body-centred cubic reciprocal lattice, and thus the Brillouin zone is therefore a three-dimensional solid, which happens to be a ‘truncated octahedron’ (see, for example reference [1], p. 89). High-symmetry points around the Brillouin zone are often labelled for ease of reference, with the most important of these, for this work, being the k = 0 point, referred to as ‘Γ’, and the < 001 > zone edges, which are called the ‘X’ points.

    ∗Choose the simplest thing first; if it works use it, and if it doesn’t, then try the next simplest!

    CHAPTER 2

    SOLUTIONS TO SCHRÖDINGER’S EQUATION

    2.1 THE INFINITE WELL

    The infinitely deep one-dimensional potential well is the simplest confinement potential to treat in quantum mechanics. Virtually every introductory level text on quantum mechanics considers this system, but nonetheless it is worth visiting again as some of the standard assumptions often glossed over, do have important consequences for one-dimensional confinement potentials in general.

    The time-independent Schrödinger equation summarizes the wave mechanics analogy to Hamilton’s formulation of classical mechanics [22], for time-independent potentials . In essence this states that the kinetic and potential energy components sum to the total energy; in wave mechanics, these quantities are the eigenvalues of linear operators, i.e.

    (2.1) c02_img01.jpg

    where the eigenfunction ψ describes the state of the system. Again in analogy with classical mechanics the kinetic energy operator for a particle of constant mass is given by the following:

    (2.2) c02_img02.jpg

    where P is the usual quantum mechanical linear momentum operator:

    (2.3)

    c02_img03.jpg

    By using this form for the kinetic energy operator Τ, the Schrödinger equation then

    (2.4)

    c02_img04.jpg

    where the function V(x, y, z) represents the potential energy of the system as a function of the spatial coordinates. Restricting this to the one-dimensional potential of interest here, then the Schrödinger equation for a particle of mass m in a potential well aligned along the z- axis (as in Fig. 2.1) would be:

    (2.5)

    c02_img05.jpg

    Figure 2.1 The one-dimensional infinite well confining potential

    c02_img06.jpg

    Outside of the well, V(z) = ∞, and hence the only possible solution is ψ(z) = 0, which in turn implies that all values of the energy E are allowed. Within the potential well, the Schrödinger equation simplifies to:

    (2.6) c02_img07.jpg

    which implies that the solution for ψ is a linear combination of the functions f(z) which when differentiated twice give –f(z). Hence try the solution:

    (2.7) c02_img08.jpg

    Substituting into equation (2.6) then gives:

    (2.8)

    c02_img09.jpg

    (2.9) c02_img10.jpg

    Consideration of the boundary conditions will yield the, as yet unknown, constant k. With this aim, consider again the kinetic energy term for this system, i.e.

    (2.10) c02_img11.jpg

    which can be rewritten as

    (2.11) c02_img12.jpg

    The mathematical form of this implies that, as a minimum, ψ(z) must be continuous. If it is not, then the first derivative will contain poles that must be avoided if the system is to have finite values for the kinetic energy. Given that ψ(z) has already been deduced as zero outside of the well, then ψ(z) within the well must be zero at both edges too.

    Figure 2.2 Solutions to the one-dimensional infinite well confining potential

    c02_img13.jpg

    If the origin is taken as the left-hand edge of the well as in Fig. 2.2, then ψ(z) as defined in equation (2.8) can contain no cosine terms, i.e. B = 0, and hence ψ(z) = A s in kz. In addition, for ψ (0) = ψ(lw) = 0:

    (2.12) c02_img14.jpg

    where n is an integer, representing a series of solutions. Substituting into equation (2.9), then the energy of the confined states is given by:

    (2.13) c02_img15.jpg

    The only remaining unknown is the constant factor A, which is deduced by considering the normalization of the wave function; as ψ*(z)ψ(z) represents the probability of finding the particle at a point z, then as the particle must exist somewhere:

    (2.14) c02_img16.jpg

    which gives c02_img17.jpg and therefore

    (2.15) c02_img18.jpg

    Under the effective mass and envelope function approximations, the energy of an electron or hole in a hypothetical infinitely deep semiconductor quantum well can be calculated by using the effective mass m∗ for the particle mass m0f equation (2.13).

    Figure 2.3 First three energy levels versus well width for an electron in a GaAs infinite potential well

    c02_img19.jpg

    Fig. 2.3 displays the results of calculations of the lowest three energy states of an electron in a GaAs well of width lw surrounded by hypothetical infinite barriers (for these and all material parameters see Appendix A). All three states show the same monotonic behaviour, with the energy decreasing as the well width increases.

    The sine function solutions derived for this system are completely standard and found extensively in the literature. Although it should be noted that the arguments developed for setting the boundary conditions, i.e. ψ(z) continuous, also implied that the first derivative should be continuous too, use is never made of this second boundary condition. The limitations of solution imposed by this are avoided by saying that not only is the potential infinite outside the well, but in addition the Schrödinger equation is not defined in these regions—a slight contradiction with the deduction of the first boundary condition. This point, i.e. that there is still ambiguity in the choice of boundary conditions for commonly accepted solutions, will be revisited later in this chapter.

    2.2 IN-PLANE DISPERSION

    If the one-dimensional potential V(z) is constructed from alternating thin layers of dissimilar semiconductors, then the particle, whether it be an electron or a hole, can move in the plane of the layers (see Fig 2.4).

    In this case, all of the terms of the kinetic energy operator are required, and hence the Schrödinger equation would be as follows:

    (2.16)

    c02_img20.jpg

    Figure 2.4 A GaAs/Ga1-x Alx as layered structure and the in-plane motion of a charge carrier

    c02_img21.jpg

    As the potential can be written as a sum of independent functions, i.e. V = V (x) + V (y)+V (z), where it just happens in this case that V (x) = V (y) = 0, the eigenfunction of the system be written as a product:

    (2.17) c02_img22.jpg

    Using this in the above Schrödinger equation, then:

    (2.18)

    c02_img23.jpg

    It is then possible to identify three distinct contributions to the total energy E, one from each of the perpendicular x-, y-, and z- axes, i.e. E = Ex + Ey + Ez . It is said that the motions ‘de-couple’ giving an equation of motion for each of the axes:

    (2.19) c02_img24.jpg

    (2.20) c02_img25.jpg

    (2.21)

    c02_img26.jpg

    Dividing throughout, then:

    (2.22) c02_img27.jpg

    (2.23) c02_img28.jpg

    (2.24) c02_img29.jpg

    The last component is identical to the one-dimensional Schrödinger equation for a confining potential V (z) as discussed, for the particular case of an infinite well, in the last section. Consider the first and second components. Again, an eigenfunction f is sought which when differentiated twice returns –f ; however, in this case it must be remembered that the solution will represent a moving particle. Thus the eigenfunction must reflect a current flow and have complex components, so try the standard travelling wave, exp (ikxx). Then:

    (2.25)

    c02_img31.jpg

    (2.26) c02_img32.jpg

    Figure 2.5 Schematic showing the in-plane (kx,y) dispersion curves and the subband structure

    c02_img30.jpg

    which is clearly just the kinetic energy of a wave travelling along the x- axis. A similar equation follows for the y- axis, and hence the in-plane motion of a particle in a one-dimensional confining potential, but of infinite extent in the x y plane can be summarized as:

    (2.27)

    c02_img33.jpg

    Therefore, while solutions of Schrödinger’s equation along the axis of the one-dimensional potential produce discrete states of energy Ez = En, in the plane of a semiconductor quantum well there is a continuous range of allowed energies, as illustrated in Fig. 2.5. In bulk materials, such domains are called ‘energy bands’, while in quantum well systems these energy domains associated with confined levels are referred to as ‘subbands’. Therefore the effect of the one-dimensional confining potential is to remove a degree of freedom, thus restricting the momentum of the charge carrier from three-dimensions to two. It is for this reason that the states within quantum well systems are generally referred to as two-dimensional.

    Later in this text, quantum wires and dots will be considered which further restrict the motion of carriers in two and three dimensions, respectively, thus giving rise to the terms one- and zero-dimensional states.

    Summarizing then, within a semiconductor quantum well system the total energy of an electron or hole, of mass m∗, with in-plane momentum kx,y, is equal to Ez + Ex,y, which is given by:

    (2.28) c02_img34.jpg

    2.3 DENSITY OF STATES

    Therefore, the original confined states within the one-dimensional potential which could each hold two charge carriers of opposite spin, from the Pauli exclusion principle, broaden into subbands, thus allowing a continuous range of carrier momenta. In order to answer the question ‘Given a particular number of electrons (or holes) within a subband, what is the distribution of their energy and momenta?’, the first point that is required is a knowledge of the density of states, i.e. how many electrons can exist within a range of energies. In order to answer this point for the case of subbands in quantum wells, it is necessary first to understand this property in bulk crystals.

    Following the idea behind Bloch’s theorem (see reference [1] p. 133) that an eigenstate within a bulk semiconductor, which can be written as ψ = (1/ ohm.jpg ) exp (i k.r), must display periodicity within the lattice, then if the unit cell is of side L:

    (2.29)

    c02_img35.jpg

    (2.30)

    c02_img36.jpg

    (2.31)

    c02_img37.jpg

    Thus for the periodicity condition to be fulfilled, the second exponential term must be identical to 1, which implies that:

    (2.32)

    c02_img38.jpg

    where nx, ny and nz are integers. Each set of values of these three integers defines a distinct state, and hence the volume of k-space occupied by one state is (2π/L)³. These states fill up with successively larger values of nx, ny and nz, i.e. the lowest energy state has values (000), then permutations of (100), (110), etc., which gradually fill a sphere. At low temperatures, the sphere has a definite boundary between states that are all occupied followed by states that are unoccupied; the momentum of these states is called the Fermi wave vector and the equivalent energy is the Fermi energy. At higher temperatures, carriers near the edge of the sphere are often scattered to higher energy states, thus ‘blurring’ the boundary between occupied and unoccupied states. For a more detailed description see, for example, Ashcroft and Mermin [1].

    Many of the interesting phenomena associated with semiconductors derive from the properties of electrons near the Fermi energy, as it is these electrons that are able to scatter into nearby states thus changing both their energy and momenta. In order to be able to progress with descriptions of, transport, for example (later in this book), it is necessary to be able to describe the density of available states.

    The density of states is defined as the number of states per energy per unit volume of real space:

    (2.33) c02_img39.jpg

    In k-space, the total number of states N is equal to the volume of the sphere of radius k, divided by the volume occupied by one state and divided again by the volume of real space, i.e.

    (2.34) c02_img40.jpg

    (2.35) c02_img41.jpg

    where the factor 2 has been introduced to allow for double occupancy of each state by the different carrier spins. Returning to the density of states, then:

    (2.36) c02_img42.jpg

    Now equation (2.35) gives

    (2.37) c02_img43.jpg

    In addition, the parabolic bands of effective mass theory give:

    (2.38)

    c02_img44.jpg

    (2.39) c02_img45.jpg

    Which finally gives the density of states in bulk as:

    (2.40) c02_img46.jpg

    Thus the density of states within a band, and around a minimum where the energy can be represented as a parabolic function of momentum, is continual and proportional to the square root of the energy.

    The density of states in quantum well systems follows analogously; however, this time, as there are only two degrees of freedom, successive states represented by values of nx and ny fill a circle in k-space, as illustrated in Fig. 2.6. Such a situation has become known as a two-dimensional electron (or hole) gas (2DEG). Hence the total number of states per unit cross-sectional area is given by the spin degeneracy factor, multiplied by the area of the circle of radius k, divided by the area occupied by each state, i.e.

    (2.41) c02_img47.jpg

    (2.42) c02_img48.jpg

    (2.43) c02_img49.jpg

    In analogy to the bulk three-dimensional (3D) case, define:

    (2.44) c02_img50.jpg

    As the in-plane dispersion curves are still described by parabolas, then reuse can be made of equation (2.39), as follows:

    (2.45) c02_img51.jpg

    Figure 2.6 Illustration of filling the two-dimensional momenta states in a quantum well

    c02_img52.jpg

    By substituting for k in terms of the energy E, using equation (2.38) then finally the density of states for a single subband in a quantum well system is given by: in agreement with Bastard [18] p.12.

    (2.46) c02_img53.jpg

    If there are many (n) confined states within the quantum well system then the density of states ρ²D at any particular energy is the sum over all subbands below that point, which can be written succinctly as:

    (2.47) c02_img54.jpg

    where Θ is the unit step function. Fig. 2.7 gives an example of the two-dimensional density of states for a particular quantum well showing the first three confined levels. Note that the steps are of equal height and occur at the subband minima—which are not equally spaced.

    Figure 2.7 The density of states as a function of energy for a 200 ÅGaAs quantum well surrounded by infinite barriers

    c02_img55.jpg

    2.4 SUBBAND POPULATIONS

    The total number of carriers within a subband is given by the integral of the product of the probability of occupation of a state and the density of states. Given that the carriers are fermions, then clearly the probability of occupation of a state is given by Fermi–Dirac statistics; hence:

    (2.48) c02_img56.jpg

    where the integral is over all of the energies of a given subband and, of course:

    (2.49) c02_img57.jpg

    Note that EF is not the Fermi energy in the traditional sense [1]; it is a ‘quasi’ Fermi energy which describes the carrier population within a subband. For systems left to reach equilibrium, the temperature T can be assumed to be the lattice temperature; however this is not always the case. In many quantum well devices which are subject to excitation by electrical or optical means, the ‘electron temperature’ can be quite different from the lattice temperature, and furthermore the subband population could be non-equilibrium and not able to be described by Fermi–Dirac statistics. For now, however, it is sufficient to discuss equilibrium electron populations and assume that the above equations are an adequate description.

    Given a particular number of carriers within a quantum well, which can usually be deduced directly from the surrounding doping density, it is often desirable to be able to describe that distribution in terms of the quasi–fermi energy EF. With this aim, substitute the two-dimensional density of states appropriate to a single subband from

    Figure 2.8 Effect of temperature on the distribution functions of the subband populations (all equal to 1× 10¹⁰cm²) of the infinite quantum well of Fig. 2.7

    c02_img58.jpg

    Figure 2.9 Effect of temperature on the quasi–fermi energy describing the electron distribution of the ground state E1

    c02_img59.jpg

    equation (2.46) into equation (2.48), then the carrier density, i.e. the number per unit area, is given by:

    (2.50)

    c02_img60.jpg

    By putting:

    (2.51)

    c02_img61.jpg

    equation (2.50) becomes:

    (2.52)

    c02_img62.jpg

    which is a standard form (see, for example, Gradshteyn and Ryzhik [23] equation 2.313.2, p.112)

    (2.53) c02_img63.jpg

    Hence:

    (2.54)

    c02_img64.jpg

    Evaluation then gives:

    (2.55)

    c02_img65.jpg

    The minimum of integration Emin is taken as the subband minima and the maximum Emax can either be taken as the top of the well, or even EF + 10kT, say, with the latter being much more stable at lower temperatures. Given a total carrier density N, the quasi–fermi energy EF is the only unknown in equation (2.55) and can be found with standard techniques. For an example of such a method, see Section 2.5.

    Figure 2.10 Effect of electron density, N = 1, 2, 5, 10 (× 10¹⁰) cm², on the distribution function of the lowest subband of the infinite quantum well of Fig. 2.7

    c02_img66.jpg

    Fig. 2.8 gives an example of the distribution functions f FD(E) for the first three confined levels within a 200 Å GaAs infinite quantum well. As the density of carriers, in this case electrons, have been taken as being equal and of value 1× 10¹⁰cm², then the distribution functions are all identical, but offset along the energy axis by the confinement energies. As mentioned above, at low temperatures the carriers tend to occupy the lowest available states, and hence the transition from states that are all occupied to those that are unoccupied is rapid—as illustrated by the 2 K data for all three subbands. As the temperature increases, the distributions broaden and a range of energies exist in which the states are partially filled, as can be seen by the 77 and 300 K data. Physically this broadening occurs due mainly to the increase in electron–phonon scattering as the phonon population increases with temperature (more of this in Chapter 10). Fig. 2.9 displays the Fermi energy EF as a function of temperature T for the ground state of energy E1 = 14. 031 meV. At low temperatures, EF is just above the confinement energy, since the electron density is fairly low (1× 10¹⁰ cm²). As the temperature increases, EF falls quite markedly and below the subband minima. If this seems counterintuitive, it must be remembered that EF is a quasi –fermi energy whose only physical meaning is to describe the population within a subband—it is not the true Fermi energy of the complete system.

    Figure 2.11 Effect of electron density on the quasi Fermi energy describing the distribution of the ground state E1

    c02_img67.jpg

    Fig. 2.10 displays the distribution functions for a range of carrier densities, for this same ground state and at a lattice temperature of 77 K. Although not obvious from the mathematics, f FD(E) at any particular energy E appears to scale with N . The corresponding Fermi energy is illustrated in Fig. 2.11. Clearly, the Fermi energy starts below the subband minima at this mid-range temperature, as discussed above, and as expected increasing numbers of carriers in the subband increases the Fermi energy, i.e. the energy of the state whose probability of occupation is 1/2.

    2.5 FINITE WELL WITH CONSTANT MASS

    While the infinitely deep confining potential has served well as a platform for developing the physics of two-dimensional systems, more relevant to alternating layers of dissimilar semiconductors is the finite quantum well model, which under both the effective mass and envelope function approximations looks like Fig. 2.12. In particular, a layer of GaAs ‘sandwiched’ between two thick layers of Ga1-x AlxAs would form a type-I finite quantum well, where the conduction band has the appearance of Fig. 2.12, with the potential energy V representing the discontinuity in the conduction band edge between the materials.

    Figure 2.12 Solutions to the finite well potential

    c02_img68.jpg

    Again taking the simplest starting case of a constant electron mass m∗throughout the dissimilar layers, and neglecting movement within the plane of the layers, then the standard Schrödinger equation can be written for each of the semiconductor layers as follows:

    (2.56)

    c02_img69.jpg

    (2.57)

    c02_img70.jpg

    (2.58)

    c02_img71.jpg

    Considering solutions to the Schrödinger equation for the central well region, then as in the infinite well case, the general solution will be a sum of sine and cosine terms. As the potential is symmetric, then the eigenstates will also have a definite symmetry, i.e. they will be either symmetric or antisymmetric. With the origin placed at the centre of the well, the symmetric (even parity) eigenstates will then be in cosine terms, while the antisymmetric (odd parity) states will be as sine waves.

    For states confined to the well, the energy E must be less than the barrier height V, thus rearranging the Schrödinger equation for the right-hand barrier:

    (2.59) c02_img72.jpg

    Therefore, a function f is sought which when differentiated twice gives +f . The exponential function fits this description, therefore consider a sum of growing exp (+κz) and decaying exp(–κz) exponentials. In the right-hand barrier, z is positive, and hence as z increases the growing exponential will increase too and without limit. The probability interpretation of the wave function requires that:

    (2.60) c02_img73.jpg

    which further demands that:

    (2.61)

    c02_img74.jpg

    These boundary conditions for states confined in wells will be used again and again and will be referred to as the standard boundary conditions . Using this result, the growing exponential components must be rejected and the solutions are for the even parity states, which would follow as:

    (2.62) c02_img75.jpg

    (2.63) c02_img76.jpg

    (2.64) c02_img77.jpg

    Note for later that these wave functions are real, and that the eigenfunctions of this confined system carry no current and hence are referred to as stationary states. Using these trial forms of the wave function in their corresponding Schrödinger equations, gives the, as yet unknown constants:

    (2.65)

    c02_img78.jpg

    Figure 2.13 Illustration of f (E) as a function of E for the even-parity solutions; where lw =100 Å, m∗ =0.067 m0 and V =1000 meV

    c02_img79.jpg

    In order to proceed, it is necessary to impose boundary conditions. Recalling the constant mass kinetic energy operator employed in equations (2.56)–(2.58), then

    Enjoying the preview?
    Page 1 of 1