Trova il tuo prossimo book preferito
Abbonati oggi e leggi gratis per 30 giorniInizia la tua prova gratuita di 30 giorniInformazioni sul libro
Semiconductor photonics. Principles and Applications
Di Mauro Nisoli
Azioni libro
Inizia a leggere- Editore:
- Società Editrice Esculapio
- Pubblicato:
- Dec 21, 2016
- ISBN:
- 9788893850049
- Formato:
- Libro
Descrizione
Informazioni sul libro
Semiconductor photonics. Principles and Applications
Di Mauro Nisoli
Descrizione
- Editore:
- Società Editrice Esculapio
- Pubblicato:
- Dec 21, 2016
- ISBN:
- 9788893850049
- Formato:
- Libro
Informazioni sull'autore
Correlati a Semiconductor photonics. Principles and Applications
Anteprima del libro
Semiconductor photonics. Principles and Applications - Mauro Nisoli
Index
Preface
The term photonics was introduced in the late sixties by Pierre Aigrain, who gave the following definition: Photonics is the science of the harnessing of light. Photonics encompasses the generation of light, the detection of light, the management of light through guidance, manipulation and amplification and, most importantly, its utilisation for the benefit of mankind.
In the last decades impressive progress in the field of photonics has been achieved, thanks to remarkable advances in the understanding of the physical processes at the heart of light-matter interaction in photonic applications and to the introduction of crucial technological innovations. Photonics is an extremely large field, as clearly pointed out by the above definition, since it refers to all types of technological device and process, where photons are involved.
This book does not aim to analyse all aspects of photonics: a few excellent textbooks already exist, which present several topics relevant for photonic applications. The aim of this book is to introduce and explain important physical processes at the heart of the optical properties of semiconductor devices, such as light emitting diodes (LEDs) and semiconductor lasers. It is suitable for a half-semester (or a one-semester) course in Photonics or Optoelectronics at the graduate level in engineering physics, electrical engineering or material science. It originated from a graduate course (Photonics I) which I am teaching at the Politecnico di Milano since 2006. The concepts of solid state physics and quantum mechanics which are required to understand the subjects discussed in this book are addressed in the introductory chapters. It is assumed that the reader has had courses on elementary quantum mechanics, solid-state physics, and electromagnetic theory at the undergraduate level.
The book presents a selection of topics, which I consider essential to understand the operation of semiconductor devices. It offers a relatively advanced analysis of the photo-physics of semiconductors, trying to avoid the use of exceedingly complex formalisms. Particular attention was devoted to offer a clear physical interpretation of all the obtained results. Various worked examples are added throughout all the chapters to illustrate the application of the various formulas: the solved exercises are evidenced by the coloured boxes in the text. The numerical examples are also important since they allow the reader to have a direct feeling of the order of magnitude of the parameters used in the formulas discussed in the text. The grey boxes contain concise discussions of supplementary topics or more advanced derivations of particular results reported in the main text, which may not be easily derived by the reader.
Semiconductor Photonics is organized as follows. Chapter 1 focuses on the description of a few concepts of solid state physics, which are relevant for the calculation and analysis of the band structure of semiconductors. The Bloch theorem is introduced, which describes the wavefunction of electrons in periodic structures. The tight-binding method is considered, with a few simple examples, and the k · p method, which are used to calculated the band structure of semiconductors. Chapter 2 deals with the discussion of the main properties of charged particles (electrons and holes) in intrinsic and doped semiconductors. The density of states is first calculated and the essential concepts of carrier statistics in semiconductors are discussed. Basic concepts of quantum mechanics are contained in Chapter 3. In particular, the density matrix formalism is introduced, which is used in the book for the calculation of the optical susceptibility of a semiconductor. After a very short overview of essential aspects of classical electromagnetic theory, Chapter 4 analyses the interaction of electrons with an electromagnetic field. The expressions of the interaction Hamiltonian, which are extensively used throughout the book, are derived in this chapter. Chapters 5 and 6 build on the previous chapters. In particular, Chapter 5 deals with the optical properties of bulk semiconductors, i.e., semiconductors with spatial dimensions much larger than the de Broglie wavelength of the electrons involved in the relevant physical processes. Absorption and gain coefficients are calculated and the radiative and non-radiative recombination processes in semiconductors are analysed. Chapter 6 analyses the principles of the photo-physics in semiconductor quantum wells, i.e., in semiconductor structures where the electrons are confined in one direction by a potential well, with a thickness smaller than the electron de Broglie wavelength.
In the remaining five chapters the general results obtained in the first part of the book are applied to the investigation of the main optical properties of semiconductor devices: light-emitting diodes and lasers. The general philosophy adopted in these chapters is the following: the fundamental physical processes are investigated, rather than the technological characteristics of the devices. After a short and general analysis of semiconductor lasers in Chapter 8, based on the rate equation approach, Chapter 9 contains a detailed theoretical analysis of the Distributed Feedback (DFB) lasers, based on the use of the coupled-mode equations. By using a simple perturbative approach, the threshold laser conditions are obtained. Vertical Cavity Surface Emitting Lasers (VCSELs) and Quantum Cascade Lasers are analysed in the final two chapters.
Final note: why Escher’s Sky and Water I
on the book cover? This print suggested me various connections with photonics in semiconductors. It is light which, by playing with shapes, gives rise to an evolution which transforms fishes in water into birds in the sky and the transformation is closely related to the reciprocal interaction among elements. In a similar manner, it is light which, by playing with electrons, triggers dynamical processes inside matter, with the generation of new properties and functions, where complex interactions among particles and surrounding environment play a crucial role. The analogy can be pushed even more considering that in Escher’s print the transition from sky to water starts and ends with realistic shapes of birds and fishes, respectively, and moves from one real shape to the other through a sequence of shapes that, in particular in the central portion of the print, resemble but are not exactly neither birds nor fishes. Forcing a little the analogy, we can say that typical photonic processes in semiconductors evolve from a given physical (real) state to another physical (real) state through a sequence of quantum superposition of states, which do not have an analogy in the realm of classical physics.
1. Band structure of semiconductors
1.1 Crystals, lattices and cells
A crystal is composed by a periodic repetition of identical groups of atoms: a group is called basis. The corresponding crystal lattice is obtained by replacing each group of atoms by a representative point, as shown in Fig. 1.1 . A crystal can be also called lattice with a basis. When the basis is composed by a single atom the corresponding lattice is called monoatomic. In a Bravais lattice the position R of all points in the lattice can be written as :
where a1, a2 and a3 are three non-coplanar translation vectors, called primitive vectors and n1, n2 and n3 are arbitrary (positive or negative) integers. This definition of Bravais lattice implies that this lattice looks exactly the same when viewed from any lattice point. Not only the arrangement of points but also the orientation must be exactly the same from every point in a Bravais lattice. Therefore, two points in the lattice, whose positions vectors are given by r and rʹ = r + R are completely equivalent environmentally. For example, the two-dimensional honeycomb lattice shown in Fig. 1.2 is not a Bravais lattice. Indeed the lattice looks the same when it is viewed from points A and C, but not when it is viewed from point B: in this case the lattice appears rotated by 180◦. We note that for a given Bravais lattice, the choice of the primitive vectors is not unique.
Figure 1.1: (a) Basis composed by two different atoms; (b) bidimensional crystal and (c) corresponding lattice.
Figure 1.2: A two-dimensional honeycomb lattice is not a Bravais lattice.
A lattice can be constructed by infinite repetitions, by translations, of a single cell without any overlapping. This cell can be primitive or non-primitive (or conventional). Primitive and conventional cells are not uniquely determined, as clearly illustrated in Fig. 1.3 . The primitive cell has the minimum possible volume, given by:
and contains exactly one lattice point. Therefore in Fig. 1.3 , which refers to a two-dimensional case, the cells 1, 2 and 3 are all primitive cells: they have the same area and contains 4 · ¹4 = 1 lattice point. The primitive cell contains the minimum possible number of atoms and there is always a lattice point per primitive cell. Cell 4 is not primitive: its area is twice the area of the primitive cell and contains 4 · ¹4 + 2 · ¹2 = 2 lattice points. Not all points in the lattice are linear combinations of a1(3) and a2(3) with integral coefficients. In some cases it is more convenient to consider conventional unit cells, with larger volumes (integer multiple of that of the primitive cell) but characterized by the same symmetry of the lattice.
Figure 1.3: Two-dimensional lattice: cells 1, 2 and 3 are primitive cells. Cell 4 is not primitive.
Without entering into any detail about group theory, we can say that all the possible lattice structures are determined by the symmetry group which describes their properties. A lattice structure can be transformed into itself not only by the translations 1.1 , which define the translational group, but also by many other symmetry operations. The symmetry operations transforming a lattice into itself keeping at least one point fixed form a group called the point group. In the case of three-dimensional structures, the point symmetry gives rise to 14 types of lattices, which can be classified depending on the relationships between the amplitudes of the vectors ai and the angles α, β and γ between them. As reported in Table 1.1, the 14 types of lattices can be grouped in one triclinic, two monoclinic, four orthorhombic, two tetragonal, three cubic, one trigonal and one hexagonal lattices.
Table 1.1: The 14 lattice systems in three dimensions (the last column shows the amplitudes ai and the angles between vectors ai of the unit cell).
Many semiconductors are characterized by a cubic lattice or by an hexagonal lattice. There are three types of cubic lattices: simple cubic (sc), body-centered cubic (bcc) and face-centered cubic (fcc), whose unit cells are shown in Fig. 1.4 . Note that
Figure 1.4: Cubic lattices: (sc) simple cubic; (bcc) body-centered cubic; (fcc) face-centered cubic.
only the simple cubic is a primitive cell, with volume a³ and one lattice point per cell (8 × ¹8). The body-centered cubic lattice can be obtained from the simple cubic by placing a lattice point at the center of the cube. The conventional cell is the cube with edge a. It has 2 lattice points per unit cell (8 × ¹8 + 1). In terms of cube edge a, a set of primitive vectors can be written as:
where ux, uy and uz are the unit vectors of the x, y and z axis, as shown in Fig. 1.5. The corresponding primitive cell, with volume a³/2, contains by definition only one lattice point. This primitive cell does not have an obvious relation with the point symmetry (cubic) of the lattice. For this reason it is useful to consider a unit cell larger than the primitive cell and with the same symmetry of the crystal. For a bcc lattice the unit cell is a cube with edge a, with a volume which is twice the volume of the primitive cell. The bcc lattice can be also considered as a sc lattice with a two-point basis 0, (a/2)(ux + uy + uz).
Figure 1.5: Set of primitive vectors for a body-centered cubic lattice.
The face-centered cubic Bravais lattice can be obtained from the sc lattice by adding a point in the center of each face. The fcc structure has lattice points on the faces of the cube, so that they are shared between two cells: the total number of lattice points in the cell is 4 (8 × ¹8 + 6 × ¹2). A particular set of primitive vectors is (see Fig. 1.6):
The primitive cell has a volume of a³/4 and contains one lattice point. Also in this case the unit cell is generally assumed as a cube of edge a, with a volume which is four times the volume of a primitive cell. The fcc can be described as a sc lattice with a four-point basis 0, (a/2)(ux + uy), (a/2)(uy + uz), (a/2)(uz + ux). We recall that the numbers giving the size of the unit cell (for example, the number a in the case of a cubic crystal) are called lattice constants.
Many important semiconductors, for example silicon and germanium have a diamond structure, which is the lattice formed by the carbon atoms in a diamond crystal. This structure consists of identical atoms, which occupy the lattice points of two inter-penetrating fcc lattices, which are displaced from each other along the body diagonal of the cubic cell by one quarter the length of the diagonal, as shown in Fig. 1.7 . It can be seen as a fcc lattice with a two-point basis 0, (a/4)(ux + uy + uz). The four nearest neighbours of each point are on the vertices of a regular tetrahedron. Note that the diamond lattice is not a Bravais lattice, since it does not look exactly the same when it is viewed from two nearest neighbour points. Since the unit cell of a fcc structure contains 4 lattice points, the unit cell of the diamond structure contains 8 lattice points. In this case it is not possible to choose a primitive cell in such a way that the basis of diamond contains only one atom. When the atoms which occupy one of the two fcc structures are different from the atoms occupying the other, the structure is called zincblende structure. Several semiconductors are characterized by this structure, such as GaAs, AlAs and many others.
Figure 1.6: Primitive vectors and primitive cell of the face-centered cubic lattice.
Figure 1.7: (a) Crystal structure of diamond; the solid lines shows the tetrahedral bond geometry; (b) atomic position in the cubic cell projected on a cube face. The fractions correspond to the height above the cube face in unit of cube edge a; the two colors correspond to the two inter-penetrating fcc lattices, which generate the diamond structure.
Largely used semiconductors such as GaN, AlN, BN, SiC, have a hexagonal closepacket (hcp) structure (wurtzite structure). Also the hcp lattice is not a Bravais lattice. This lattice can be seen as two inter-penetrating simple hexagonal Bravais lattices, displaced vertically by c/2 in the direction of the common c–axis, and displaced in the horizontal plane in such a way that the points of one simple hexagonal lattice are placed above the centres of the triangles formed by the points of the other simple hexagonal lattice. Figure 1.9 shows a simple hexagonal Bravais lattice, which is obtained by stacking two-dimensional triangular Bravais lattices one exactly above the other along a direction perpendicular to each two-dimensional lattice. This stacking direction is usually called crystallographic c–axis. The primitive vectors can be written as:
Figure 1.8: Atomic position in hcp lattice.
Figure 1.9: Simple hexagonal Bravais lattice.
1.1.1 The Wigner-Seitz cell
The primitive cell can be also chosen in such a way that it presents the full symmetry of the Bravais lattice. This can be achieved by considering the Wigner-Seitz cell. The mathematical definition is the following: the Wigner-Seitz cell around a given lattice point is the spatial region that is closest to that particular lattice point than to any of other lattice points. It can also be demonstrated that the Wigner-Seitz cell is a primitive cell. The above definition does not refer to any particular choice of primitive vectors, for this reason the Wigner-Seitz cell is as symmetrical as the Bravais lattice. The procedure for the construction of a Wigner-Seitz cell can be illustrated in the simple case of a two-dimensional rectangular lattice, as shown in Fig. 1.10 (a). To determine the Wigner-Seitz cell about the lattice point P we have first to draw the lines from P to all of its nearest neighbours (Fig. 1.10 (b)) and then the bisectors to each of these lines (Fig. 1.10 (c)). The Wigner-Seitz cell is the innermost region bounded by the perpendicular bisectors, as shown by the shaded region in Fig. 1.10 (c). The same procedure can be applied in the case of a generic three-dimensional lattice.
Figure 1.10: Construction of the Wigner-Seitz cell of a two-dimensional rectangular lattice.
1.2 The reciprocal lattice
In order to develop an analytic study of a crystalline solid, it is often useful to introduce the concept of reciprocal lattice, which basically represents the Fourier transform of the Bravais lattice. Assuming a Bravais lattice defined by the primitive translation vectors (a , a , a ), the reciprocal lattice can be defined by introducing its primitive translation vectors (b1, b2, b3) in analogy with the lattice in real space. The axis vectors of the reciprocal space can be written as:
where Vu is the volume of the unit cell given by:
The reciprocal lattice can be mapped by using the general translation vector G given by:
where m1, m2 and m3 are integers. Vector G is called reciprocal lattice vector. Note that each vector given by 1.6 is orthogonal to two axis vectors of the crystal lattice, so that:
where δij is the Kronecker delta symbol: δij = 0 for i ≠ j and δij = 1 for i = j. Moreover:
so that:
Any function f(r) with the periodicity of the crystal lattice, i.e., f(r + R) = f(r), can be expanded as:
where:
Indeed:
where we have used Eq. 1.11.
While vectors in the direct lattice have the dimension of length, the vectors in the reciprocal lattice have the dimension of [length]–1. The reciprocal space is therefore the most convenient space for wave vectors k. Since each point in the reciprocal space can be reached by the translation vector G, it is evident that we can restrict our analysis to a unit cell defined by the vectors bi. The Wigner-Seitz cell of the reciprocal lattice is called the first Brillouin
Recensioni
Recensioni
Cosa pensano gli utenti di Semiconductor photonics. Principles and Applications
00 valutazioni / 0 recensioni