Sie sind auf Seite 1von 9

Electronic band structure From Wikipedia, the free encyclopedia (Redirected from Band theory of solids) Jump to:

navigation, search Electronic structure methods Valence bond theory Generalized valence bond Modern valence bond Molecular orbital theory Hartree Fock method Mller Plesset perturbation theory Configuration interaction Coupled cluster Multi-configurational self-consistent field Quantum chemistry composite methods Quantum Monte Carlo Linear combination of atomic orbitals Electronic band structure Nearly free electron model Tight binding Muffin-tin approximation Density functional theory kp perturbation theory Empty Lattice Approximation Wikipedia book Book v t e In solid-state physics, the electronic band structure (or simply band structure) of a solid describes those ranges of energy, called energy bands, that an elect ron within the solid may have ("allowed bands"), and ranges of energy called ban d gaps ("forbidden bands"), which it may not have. Band theory models the behavi or of electrons in solids by postulating the existence of energy bands. It succe ssfully uses a material's band structure to explain many physical properties of solids, such as electrical resistivity and optical absorption. Bands may also be viewed as the large-scale limit of molecular orbital theory. A solid creates a large number of closely spaced molecular orbitals, which appear as a band. Band structure derives from the dynamical theory of diffraction of the quantum mechan ical electron waves in a periodic crystal lattice with a specific crystal system and Bravais lattice. Contents 1 Why bands occur in materials 2 Basic concepts 2.1 Symmetry 2.2 Band structures in different types of solids 2.3 Solid state properties and the Pauli principle 2.4 Density of states 2.5 Filling of bands 3 Theory of band structures in crystals 3.1 Nearly free electron approximation 3.2 Tight binding model 3.3 KKR model 3.4 Density-functional theory 3.5 Green's function methods and the ab initio GW approximation 3.6 Mott insulators 3.7 Others

4 5 6 7 8

References Bibliography Further reading External links See also

Why bands occur in materials See also: Conduction band, Valence band, and Bandgap The electrons of a single isolated atom occupy atomic orbitals, which form a dis crete set of energy levels. If several atoms are brought together into a molecul e, their atomic orbitals split into separate molecular orbitals each with a diff erent energy. This is due to the Pauli exclusion principle, which says that elec trons that are close together must have different sets of quantum numbers (energ y). This produces a number of molecular orbitals proportional to the number of v alence electrons. When a large number of atoms (of order 1020 or more) are brough t together to form a solid, the number of orbitals becomes exceedingly large. Co nsequently, the difference in energy between them becomes very small. Thus, in s olids the levels form continuous bands of energy rather than the discrete energy levels of the atoms in isolation. However, some intervals of energy contain no orbitals, no matter how many atoms are aggregated, forming band gaps. Within an energy band, energy levels can be regarded as a near continuum for two reasons. First, the separation between energy levels in a solid is comparable w ith the energy that electrons constantly exchange with phonons (atomic vibration s). Second, this separation is comparable with the energy uncertainty due to the Heisenberg uncertainty principle, for reasonably long intervals of time. As a r esult, for shorter timescales the separation between energy levels is of no cons equence. Several approaches to finding band structure are discussed below. Basic concepts Animation for the formation of a band in a metal and an insulator Any solid has a large number of bands. In theory, a solid can ny bands (just as an atom has infinitely many energy levels). few of these bands lie at energies so high that any electron energies will escape from the solid. These bands are usually have infinitely ma However, all but a that attains those disregarded.

Bands have different widths, based upon the properties of the atomic orbitals fr om which they arise. Also, allowed bands may overlap, producing (for practical p urposes) a single large band. Figure 1 shows a simplified picture of the bands in a solid that allows the thre e major types of materials to be identified: metals, semiconductors and insulato rs. Metals contain a band that is partly empty and partly filled regardless of tempe rature. Therefore they have very high conductivity. The lowermost band is called the valence band by analogy with the valence electr ons of individual atoms. This band is almost fully occupied in an insulator or s emiconductor. The uppermost, almost unoccupied band is called the conduction ban d because only when electrons are excited to the conduction band can current flo w in these materials. The difference between insulators and semiconductors is on ly that the forbidden band gap between the valence band and conduction band is l arger in an insulator, so that fewer electrons are found there and the electrica l conductivity is lower. Because one of the main mechanisms for electrons to be excited to the conduction band is due to thermal energy, the conductivity of sem iconductors is strongly dependent on the temperature of the material.

This band gap is one of the most useful aspects of the band structure, as it str ongly influences the electrical and optical properties of the material. Electron s can transfer from one band to the other by means of carrier generation and rec ombination processes. The band gap and defect states created in the band gap by doping can be used to create semiconductor devices such as solar cells, diodes, transistors, laser diodes, and others. Symmetry Figure 2: First Brillouin zone of FCC lattice showing symmetry labels Figure 3: Bulk band structure for Si, Ge, GaAs and InAs generated with tight bin ding model. Note that Si and Ge are indirect, while GaAs and InAs are direct ban d gap materials. See also: Symmetry in physics, Crystallographic point group, and Space group A more complete view of the band structure takes into account the periodic natur e of a crystal lattice using the symmetry operations that form a space group. Th e Schrdinger equation is solved for the crystal, which has Bloch waves as solutio ns: \psi_{n\mathbf{k}}(\mathbf{r})=e^{i\mathbf{k}\cdot\mathbf{r}}u_{n\mathbf{k}} (\mathbf{r}), where k is called the wavevector, and is related to the direction of motion of t he electron in the crystal, and n is the band index, which simply numbers the en ergy bands. The wavevector k takes on values within the Brillouin zone (BZ) corr esponding to the crystal lattice, and particular directions/points in the BZ are assigned conventional names like G, ?, ?, S, etc. These directions are shown fo r the face-centered cubic lattice geometry in Figure 2. The available energies for the electron also depend upon k, as shown in Figure 3 for silicon in the more complex energy band diagram at the right. In this diagr am the topmost energy of the valence band is labeled Ev and the bottom energy in the conduction band is labeled Ec. Note that for silicon, the top of the valenc e band is not directly below the bottom of the conduction band (Ev is for an ele ctron traveling in direction G, Ec in direction X), so silicon is called an indi rect gap material. For an electron to be excited from the valence band to the co nduction band within an indirect gap material, it needs something to give it bot Ev and a change in direction/momentum. In other semiconductors (for h energy Ec example III-V materials, such as GaAs) both Ec and Ev are at G, and therefore th ese materials are direct gap materials (no momentum change required). Direct gap materials benefit the operation of semiconductor laser diodes. Anderson's rule is used to align band diagrams between two different semiconduct ors in contact. Band structures in different types of solids Although electronic band structures are usually associated with crystalline mate rials, quasi-crystalline and amorphous solids may also exhibit band structures. However, the periodic nature and symmetrical properties of crystalline materials makes it much easier to examine the band structures of these materials theoreti cally. In addition, the well-defined symmetry axes of crystalline materials make s it possible to determine the dispersion relationship between the momentum (a 3 -dimension vector quantity) and energy of a material. As a result, virtually all of the existing theoretical work on the electronic band structure of solids has focused on crystalline materials. Solid state properties and the Pauli principle In conductors and semi-conductors, free electrons have to share entire bulk spac e. Thus, their energy levels stack up, creating band structure out of each atomi c energy level. In strong conductors (metals) electrons are so degenerate that t

hey can not even contribute much to the thermal capacity of a metal. Many mechan ical, electrical, magnetic, optical and chemical properties of solids are the di rect consequence of the Pauli exclusion principle. Density of states Main article: Density of states See also: Effective mass (solid-state physics) While the density of energy states in a band could be very large for some materi als, it may not be uniform. It approaches zero at the band boundaries, and is ge nerally highest near the middle of a band. The density of states for the free el ectron model in three dimensions is given by, D(\epsilon)= \frac{V}{2\pi^2}\left(\frac {2m}{\hbar^2}\right)^{3/2} \epsilon ^{1/2} Filling of bands Filling of electronic band structure in various types of material at equilibrium . In metals and semimetals the Fermi level EF lies inside at least one band. In insulators and semiconductors the Fermi level is inside a band gap, however in s emiconductors the bands are near enough to the Fermi level to be thermally popul ated with electrons or holes. See also: Fermi statistics Although the number of states in all of the bands is effectively infinite, in an uncharged material the number of electrons is equal only to the number of proto ns in the atoms of the material. Therefore not all of the states are occupied by electrons ("filled") at any time. The likelihood of any particular state being filled at any temperature is given by Fermi-Dirac statistics. The probability is given by the following expression: f(E) = \frac{1}{1 + e^{\frac{E-\mu}{k_B T}}} where: kB is Boltzmann's constant,yhjghjgh T is the temperature, is the chemical potential of electrons (in semiconductor physics, this quant ity is more often called the "Fermi level" and denoted EF). The Fermi level naturally is the level at which the electrons and protons are ba lanced. At T=0, the distribution is a simple step function: f(E) = \begin{cases} 1 & \mbox{if}\ E < E_F \\ 0.5 & \mbox{if}\ E = E_F \\ 0 & \mbox{if}\ E_F < E \end{cases} At nonzero temperatures, the step "smooths out", so that an appreciable number o f states below the Fermi level are empty, and some states above the Fermi level are filled. Theory of band structures in crystals The ansatz is the special case of electron waves in a periodic crystal lattice u sing Bloch waves as treated generally in the dynamical theory of diffraction. Ev ery crystal is a periodic structure which can be characterized by a Bravais latt ice, and for each Bravais lattice we can determine the reciprocal lattice, which encapsulates the periodicity in a set of three reciprocal lattice vectors (b1,b 2,b3). Now, any periodic potential V(r) which shares the same periodicity as the direct lattice can be expanded out as a Fourier series whose only non-vanishing components are those associated with the reciprocal lattice vectors. So the exp

ansion can be written as: V(\mathbf{r}) = \sum_{\mathbf{K}}{V_{\mathbf{K}}e^{i \mathbf{K}\cdot\mathbf{ r}}} where K = m1b1 + m2b2 + m3b3 for any set of integers (m1,m2,m3). From this theory, an attempt can be made to predict the band structure of a part icular material, however most ab initio methods for electronic structure calcula tions fail to predict the observed band gap. Nearly free electron approximation Main articles: Nearly free electron model, Free electron model, and pseudopotent ial In the nearly free electron approximation, interactions between electrons are co mpletely ignored. This approximation allows use of Bloch's Theorem which states that electrons in a periodic potential have wavefunctions and energies which are periodic in wavevector up to a constant phase shift between neighboring recipro cal lattice vectors. The consequences of periodicity are described mathematicall y by the Bloch wavefunction: {\Psi}_{n,\mathbf{k}} (\mathbf{r}) = e^{i \mathbf{k}\cdot\mathbf{r}} u_n(\ma thbf{r}) where the function u_n(\mathbf{r}) is periodic over the crystal lattice, that is , u_n(\mathbf{r}) = u_n(\mathbf{r-R}) . Here index n refers to the n-th energy band, wavevector k is related to the dire ction of motion of the electron, r is the position in the crystal, and R is the location of an atomic site.[1] The NFE model works particularly well in materials like metals where distances b etween neighbouring atoms are small. In such materials the overlap of atomic orb itals and potentials on neighbouring atoms is relatively large. In that case the wave function of the electron can be approximated by a (modified) plane wave. T he band structure of a metal like Aluminum even gets close to the empty lattice approximation. Tight binding model Main article: Tight binding The opposite extreme to the nearly free electron approximation assumes the elect rons in the crystal behave much like an assembly of constituent atoms. This tigh t binding model assumes the solution to the time-independent single electron Sch rdinger equation \Psi is well approximated by a linear combination of atomic orbi tals \psi_n(\mathbf{r}).[2] \Psi(\mathbf{r}) = \sum_{n,\mathbf{R}} b_{n,\mathbf{R}} \psi_n(\mathbf{r-R}) , where the coefficients b_{n,\mathbf{R}} are selected to give the best approximat e solution of this form. Index n refers to an atomic energy level and R refers t o an atomic site. A more accurate approach using this idea employs Wannier funct ions, defined by:[3][4] a_n(\mathbf{r-R}) = \frac{V_{C}}{(2\pi)^{3}} \int_{BZ} d\mathbf{k} e^{-i\mat hbf{k}\cdot(\mathbf{R-r})}u_{n\mathbf{k}}; in which u_{n\mathbf{k}} is the periodic part of the Bloch wave and the integral

is over the Brillouin zone. Here index n refers to the n-th energy band in the crystal. The Wannier functions are localized near atomic sites, like atomic orbi tals, but being defined in terms of Bloch functions they are accurately related to solutions based upon the crystal potential. Wannier functions on different at omic sites R are orthogonal. The Wannier functions can be used to form the Schrdi nger solution for the n-th energy band as: \Psi_{n,\mathbf{k}} (\mathbf{r}) = \sum_{\mathbf{R}} e^{-i\mathbf{k}\cdot(\m athbf{R-r})}a_n(\mathbf{r-R}). The TB model works well in materials with limited overlap between atomic orbital s and potentials on neighbouring atoms. Band structures of materials like Si, Ga As, SiO2 and diamond for instance are well described by TB-Hamiltonians on the b asis of atomic sp3 orbitals. In transition metals a mixed TB-NFE model is used t o describe the broad NFE conduction band and the narrow embedded TB d-bands. The radial functions of the atomic orbital part of the Wannier functions are most e asily calculated by the use of pseudopotential methods. NFE, TB or combined NFETB band structure calculations,[5] sometimes extended with wave function approxi mations based on pseudopotential methods, are often used as an economic starting point for further calculations. KKR model Main article: Muffin-tin approximation The simplest form of this approximation centers non-overlapping spheres (referre d to as muffin tins) on the atomic positions. Within these regions, the potentia l experienced by an electron is approximated to be spherically symmetric about t he given nucleus. In the remaining interstitial region, the screened potential i s approximated as a constant. Continuity of the potential between the atom-cente red spheres and interstitial region is enforced. A variational implementation was suggested by Korringa and by Kohn and Rostocker , and is often referred to as the KKR model.[6][7] Density-functional theory Main article: Density functional theory See also: Kohn-Sham equations In recent physics literature, a large majority of the electronic structures and band plots are calculated using density-functional theory (DFT), which is not a model but rather a theory, i.e., a microscopic first-principles theory of conden sed matter physics that tries to cope with the electron-electron many-body probl em via the introduction of an exchange-correlation term in the functional of the electronic density. DFT-calculated bands are in many cases found to be in agree ment with experimentally measured bands, for example by angle-resolved photoemis sion spectroscopy (ARPES). In particular, the band shape is typically well repro duced by DFT. But there are also systematic errors in DFT bands when compared to experiment results. In particular, DFT seems to systematically underestimate by about 30-40% the band gap in insulators and semiconductors. DFT is, in principl e, an exact theory to reproduce and predict ground state properties (e.g., the t otal energy, the atomic structure, etc.). However, DFT is not a theory to addres s excited state properties, such as the band plot of a solid that represents the excitation energies of electrons injected or removed from the system. What in t he literature is quoted as a DFT band plot is a representation of the DFT Kohn-S ham energies, i.e., the energies of a fictive non-interacting system, the Kohn-S ham system, which has no physical interpretation at all. The Kohn-Sham electroni c structure must not be confused with the real, quasiparticle electronic structu re of a system, and there is no Koopman's theorem holding for Kohn-Sham energies , as there is for Hartree-Fock energies, which can be truly considered as an app roximation for quasiparticle energies. Hence, in principle, Kohn-Sham based DFT is not a band theory, i.e., not a theory suitable for calculating bands and band -plots. In principle time-dependent DFT can be used to calculate the true band s

tructure although in practise this is often difficult. Green's function methods and the ab initio GW approximation Main articles: Green's function (many-body theory) and Green Kubo relations To calculate the bands including electron-electron interaction many-body effects , one can resort to so-called Green's function methods. Indeed, knowledge of the Green's function of a system provides both ground (the total energy) and also e xcited state observables of the system. The poles of the Green's function are th e quasiparticle energies, the bands of a solid. The Green's function can be calc ulated by solving the Dyson equation once the self-energy of the system is known . For real systems like solids, the self-energy is a very complex quantity and u sually approximations are needed to solve the problem. One such approximation is the GW approximation, so called from the mathematical form the self-energy take s as the product S = GW of the Green's function G and the dynamically screened i nteraction W. This approach is more pertinent when addressing the calculation of band plots (and also quantities beyond, such as the spectral function) and can also be formulated in a completely ab initio way. The GW approximation seems to provide band gaps of insulators and semiconductors in agreement with experiment, and hence to correct the systematic DFT underestimation. Mott insulators Main article: Mott insulator Although the nearly free electron approximation is able to describe many propert ies of electron band structures, one consequence of this theory is that it predi cts the same number of electrons in each unit cell. If the number of electrons i s odd, we would then expect that there is an unpaired electron in each unit cell , and thus that the valence band is not fully occupied, making the material a co nductor. However, materials such as CoO that have an odd number of electrons per unit cell are insulators, in direct conflict with this result. This kind of mat erial is known as a Mott insulator, and requires inclusion of detailed electronelectron interactions (treated only as an averaged effect on the crystal potenti al in band theory) to explain the discrepancy. The Hubbard model is an approxima te theory that can include these interactions. It can be treated non-perturbativ ely within the so-called Dynamical Mean Field Theory, which bridges the gap betw een the nearly free electron approximation and the atomic limit. Others Calculating band structures is an important topic in theoretical solid state phy sics. In addition to the models mentioned above, other models include the follow ing: kp perturbation theory is a technique that allows a band structure to be appr oximately described in terms of just a few parameters. The technique is commonly used for semiconductors, and the parameters in the model are often determined b y experiment. The Kronig-Penney Model, a one-dimensional rectangular well model useful for illustration of band formation. While simple, it predicts many important phenom ena, but is not quantitative. Hubbard model The band structure has been generalised to wavevectors that are complex numbers, resulting in what is called a complex band structure, which is of interest at s urfaces and interfaces. Each model describes some types of solids very well, and others poorly. The near ly free electron model works well for metals, but poorly for non-metals. The tig ht binding model is extremely accurate for ionic insulators, such as metal halid e salts (e.g. NaCl). References Wikimedia Commons has media related to: Dispersion relations of electron

s ^ Kittel, p. 179 ^ Kittel, pp. 245-248 ^ Kittel, Eq. 42 p. 267 ^ Daniel Charles Mattis (1994). The Many-Body Problem: Encyclopaedia of Exac tly Solved Models in One Dimension. World Scientific. p. 340. ISBN 981-02-1476-6 . ^ Walter Ashley Harrison (1989). Electronic Structure and the Properties of Solids. Dover Publications. ISBN 0-486-66021-4. ^ Joginder Singh Galsin (2001). Impurity Scattering in Metal Alloys. Springe r. Appendix C. ISBN 0-306-46574-4. ^ Kuon Inoue, Kazuo Ohtaka (2004). Photonic Crystals. Springer. p. 66. ISBN 3-540-20559-4. Bibliography Charles Kittel (1996). Introduction to Solid State Physics (Seventh ed.). Ne w York: Wiley. ISBN 0-471-11181-3. Further reading Microelectronics, by Jacob Millman and Arvin Gabriel, ISBN 0-07-463736-3, Ta ta McGraw-Hill Edition. Solid State Physics, by Neil Ashcroft and N. David Mermin, ISBN 0-03-0839939 Elementary Solid State Physics: Principles and Applications, by M. Ali Omar, ISBN 0-201-60733-6 Electronic and Optoelectronic Properties of Semiconductor Structures - Chapt er 2 and 3 by Jasprit Singh, ISBN 0-521-82379-X Electronic Structure: Basic Theory and Practical Methods by Richard Martin, ISBN 0-521-78285-6 Condensed Matter Physics by Michael P. Marder, ISBN 0-471-17779-2 V. Antonov, ISBN 90-5699-094-2 Elementary Electronic Structure by Walter A. Harrison, ISBN 981-238-708-0 Pseudopotentials in the theory of metals by Walter A. Harrison, W.A. Benjami n (New York) 1966 Tutorial on Bandstructure Methods by Dr. Vasileska (2008) Britney Spears' Guide to Semiconductor Physics. (2000) External links Animation, applications and research about quantum physics and band theory ( Universit Paris Sud) See also Alan Herries Wilson Anderson's rule Band Gap Bloch waves Dynamical Mean Field Theory Dynamical theory of diffraction Effective mass Empty Lattice Approximation Fermi gas Fermi surface Free electron model kp method

Kronig-Penney model Local-density approximation Nearly

Das könnte Ihnen auch gefallen