Metal-induced gap states

Last updated

In bulk semiconductor band structure calculations, it is assumed that the crystal lattice (which features a periodic potential due to the atomic structure) of the material is infinite. When the finite size of a crystal is taken into account, the wavefunctions of electrons are altered and states that are forbidden within the bulk semiconductor gap are allowed at the surface. Similarly, when a metal is deposited onto a semiconductor (by thermal evaporation, for example), the wavefunction of an electron in the semiconductor must match that of an electron in the metal at the interface. Since the Fermi levels of the two materials must match at the interface, there exists gap states that decay deeper into the semiconductor.

Contents

Band-bending at the metal-semiconductor interface

Band diagram of the band-bending at the interface of (a) a low work function metal and n-type semiconductor, (b) a low work function metal and a p-type semi conductor, (c) a high work function metal and an n-type semi conductor, (d) a high work function metal and a p-type semi conductor. (Figure adapted from H. Luth's Solid Surfaces, Interfaces, and Thin Films, p. 384. ) Diagram of band-bending interfaces between two different metals and two different semiconductors.jpg
Band diagram of the band-bending at the interface of (a) a low work function metal and n-type semiconductor, (b) a low work function metal and a p-type semi conductor, (c) a high work function metal and an n-type semi conductor, (d) a high work function metal and a p-type semi conductor. (Figure adapted from H. Luth's Solid Surfaces, Interfaces, and Thin Films, p. 384. )

As mentioned above, when a metal is deposited onto a semiconductor, even when the metal film as small as a single atomic layer, the Fermi levels of the metal and semiconductor must match. This pins the Fermi level in the semiconductor to a position in the bulk gap. Shown to the right is a diagram of band-bending interfaces between two different metals (high and low work functions) and two different semiconductors (n-type and p-type).

Volker Heine was one of the first to estimate the length of the tail end of metal electron states extending into the semiconductor's energy gap. He calculated the variation in surface state energy by matching wavefunctions of a free-electron metal to gapped states in an undoped semiconductor, showing that in most cases the position of the surface state energy is quite stable regardless of the metal used. [2]

Branching point

It is somewhat crude to suggest that the metal-induced gap states (MIGS) are tail ends of metal states that leak into the semiconductor. Since the mid-gap states do exist within some depth of the semiconductor, they must be a mixture (a Fourier series) of valence and conduction band states from the bulk. The resulting positions of these states, as calculated by C. Tejedor, F. Flores and E. Louis, [3] and J. Tersoff, [4] [5] must be closer to either the valence- or conduction- band thus acting as acceptor or donor dopants, respectively. The point that divides these two types of MIGS is called the branching point, E_B. Tersoff argued

, where is the spin orbit splitting of at the point.
is the indirect conduction band minimum.

Metal–semiconductor contact point barrier height

Band diagram of the contact point potential barrier at the interface of a metal and semiconductor. Shown are
e
Ph
b
h
{\displaystyle e\Phi _{bh}}
, the energy of the barrier, and
e
V
i
f
{\displaystyle eV_{if}}
, the maximum band bending in the semiconductor. (Figure adapted from H. Luth's Solid Surfaces, Interfaces, and Thin Films, p. 408 (see Refs.) Migsbarrier.JPG
Band diagram of the contact point potential barrier at the interface of a metal and semiconductor. Shown are , the energy of the barrier, and , the maximum band bending in the semiconductor. (Figure adapted from H. Luth's Solid Surfaces, Interfaces, and Thin Films, p. 408 (see Refs.)

In order for the Fermi levels to match at the interface, there must be charge transfer between the metal and semiconductor. The amount of charge transfer was formulated by Linus Pauling [6] and later revised [7] to be:

where and are the electronegativities of the metal and semiconductor, respectively. The charge transfer produces a dipole at the interface and thus a potential barrier called the Schottky barrier height. In the same derivation of the branching point mentioned above, Tersoff derives the barrier height to be:

where is a parameter adjustable for the specific metal, dependent mostly on its electronegativity, . Tersoff showed that the experimentally measured fits his theoretical model for Au in contact with 10 common semiconductors, including Si, Ge, GaP, and GaAs.

Another derivation of the contact barrier height in terms of experimentally measurable parameters was worked out by Federico Garcia-Moliner and Fernando Flores who considered the density of states and dipole contributions more rigorously. [8]

is dependent on the charge densities of the both materials
density of surface states
work function of metal
sum of dipole contributions considering dipole corrections to the jellium model
semiconductor gap
Ef – Ev in semiconductor

Thus can be calculated by theoretically deriving or experimentally measuring each parameter. Garcia-Moliner and Flores also discuss two limits

(The Bardeen Limit), where the high density of interface states pins the Fermi level at that of the semiconductor regardless of .
(The Schottky Limit) where varies with strongly with the characteristics of the metal, including the particular lattice structure as accounted for in .

Applications

When a bias voltage is applied across the interface of an n-type semiconductor and a metal, the Fermi level in the semiconductor is shifted with respect to the metal's and the band bending decreases. In effect, the capacitance across the depletion layer in the semiconductor is bias voltage dependent and goes as . This makes the metal/semiconductor junction useful in varactor devices used frequently in electronics.

Related Research Articles

<span class="mw-page-title-main">Feynman diagram</span> Pictorial representation of the behavior of subatomic particles

In theoretical physics, a Feynman diagram is a pictorial representation of the mathematical expressions describing the behavior and interaction of subatomic particles. The scheme is named after American physicist Richard Feynman, who introduced the diagrams in 1948. The interaction of subatomic particles can be complex and difficult to understand; Feynman diagrams give a simple visualization of what would otherwise be an arcane and abstract formula. According to David Kaiser, "Since the middle of the 20th century, theoretical physicists have increasingly turned to this tool to help them undertake critical calculations. Feynman diagrams have revolutionized nearly every aspect of theoretical physics." While the diagrams are applied primarily to quantum field theory, they can also be used in other areas of physics, such as solid-state theory. Frank Wilczek wrote that the calculations that won him the 2004 Nobel Prize in Physics "would have been literally unthinkable without Feynman diagrams, as would [Wilczek's] calculations that established a route to production and observation of the Higgs particle."

In solid-state physics, the work function is the minimum thermodynamic work needed to remove an electron from a solid to a point in the vacuum immediately outside the solid surface. Here "immediately" means that the final electron position is far from the surface on the atomic scale, but still too close to the solid to be influenced by ambient electric fields in the vacuum. The work function is not a characteristic of a bulk material, but rather a property of the surface of the material.

In physics, screening is the damping of electric fields caused by the presence of mobile charge carriers. It is an important part of the behavior of charge-carrying fluids, such as ionized gases, electrolytes, and charge carriers in electronic conductors . In a fluid, with a given permittivity ε, composed of electrically charged constituent particles, each pair of particles interact through the Coulomb force as

In the calculus of variations, a field of mathematical analysis, the functional derivative relates a change in a functional to a change in a function on which the functional depends.

<span class="mw-page-title-main">Schottky barrier</span> Potential energy barrier in metal–semiconductor junctions

A Schottky barrier, named after Walter H. Schottky, is a potential energy barrier for electrons formed at a metal–semiconductor junction. Schottky barriers have rectifying characteristics, suitable for use as a diode. One of the primary characteristics of a Schottky barrier is the Schottky barrier height, denoted by ΦB. The value of ΦB depends on the combination of metal and semiconductor.

<span class="mw-page-title-main">Density of states</span> Number of available physical states per energy unit

In solid-state physics and condensed matter physics, the density of states (DOS) of a system describes the number of modes per unit frequency range. The density of states is defined as , where is the number of states in the system of volume whose energies lie in the range from to . It is mathematically represented as a distribution by a probability density function, and it is generally an average over the space and time domains of the various states occupied by the system. The density of states is directly related to the dispersion relations of the properties of the system. High DOS at a specific energy level means that many states are available for occupation.

<span class="mw-page-title-main">Hyperfine structure</span> Small shifts and splittings in the energy levels of atoms, molecules and ions

In atomic physics, hyperfine structure is defined by small shifts in otherwise degenerate energy levels and the resulting splittings in those energy levels of atoms, molecules, and ions, due to electromagnetic multipole interaction between the nucleus and electron clouds.

A heterojunction is an interface between two layers or regions of dissimilar semiconductors. These semiconducting materials have unequal band gaps as opposed to a homojunction. It is often advantageous to engineer the electronic energy bands in many solid-state device applications, including semiconductor lasers, solar cells and transistors. The combination of multiple heterojunctions together in a device is called a heterostructure, although the two terms are commonly used interchangeably. The requirement that each material be a semiconductor with unequal band gaps is somewhat loose, especially on small length scales, where electronic properties depend on spatial properties. A more modern definition of heterojunction is the interface between any two solid-state materials, including crystalline and amorphous structures of metallic, insulating, fast ion conductor and semiconducting materials.

<span class="mw-page-title-main">Fermi's interaction</span> Mechanism of beta decay proposed in 1933

In particle physics, Fermi's interaction is an explanation of the beta decay, proposed by Enrico Fermi in 1933. The theory posits four fermions directly interacting with one another. This interaction explains beta decay of a neutron by direct coupling of a neutron with an electron, a neutrino and a proton.

In quantum physics, the spin–orbit interaction is a relativistic interaction of a particle's spin with its motion inside a potential. A key example of this phenomenon is the spin–orbit interaction leading to shifts in an electron's atomic energy levels, due to electromagnetic interaction between the electron's magnetic dipole, its orbital motion, and the electrostatic field of the positively charged nucleus. This phenomenon is detectable as a splitting of spectral lines, which can be thought of as a Zeeman effect product of two relativistic effects: the apparent magnetic field seen from the electron perspective and the magnetic moment of the electron associated with its intrinsic spin. A similar effect, due to the relationship between angular momentum and the strong nuclear force, occurs for protons and neutrons moving inside the nucleus, leading to a shift in their energy levels in the nucleus shell model. In the field of spintronics, spin–orbit effects for electrons in semiconductors and other materials are explored for technological applications. The spin–orbit interaction is at the origin of magnetocrystalline anisotropy and the spin Hall effect.

<span class="mw-page-title-main">Pseudopotential</span>

In physics, a pseudopotential or effective potential is used as an approximation for the simplified description of complex systems. Applications include atomic physics and neutron scattering. The pseudopotential approximation was first introduced by Hans Hellmann in 1934.

<span class="mw-page-title-main">Anderson's rule</span>

Anderson's rule is used for the construction of energy band diagrams of the heterojunction between two semiconductor materials. Anderson's rule states that when constructing an energy band diagram, the vacuum levels of the two semiconductors on either side of the heterojunction should be aligned.

<span class="mw-page-title-main">Angle-resolved photoemission spectroscopy</span> Experimental technique to determine the distribution of electrons in solids

Angle-resolved photoemission spectroscopy (ARPES) is an experimental technique used in condensed matter physics to probe the allowed energies and momenta of the electrons in a material, usually a crystalline solid. It is based on the photoelectric effect, in which an incoming photon of sufficient energy ejects an electron from the surface of a material. By directly measuring the kinetic energy and emission angle distributions of the emitted photoelectrons, the technique can map the electronic band structure and Fermi surfaces. ARPES is best suited for the study of one- or two-dimensional materials. It has been used by physicists to investigate high-temperature superconductors, graphene, topological materials, quantum well states, and materials exhibiting charge density waves.

In solid-state physics, a metal–semiconductor (M–S) junction is a type of electrical junction in which a metal comes in close contact with a semiconductor material. It is the oldest practical semiconductor device. M–S junctions can either be rectifying or non-rectifying. The rectifying metal–semiconductor junction forms a Schottky barrier, making a device known as a Schottky diode, while the non-rectifying junction is called an ohmic contact.

In solid-state physics, the Poole–Frenkel effect is a model describing the mechanism of trap-assisted electron transport in an electrical insulator. It is named after Yakov Frenkel, who published on it in 1938, extending the theory previously developed by H. H. Poole.

The Mattis–Bardeen theory is a theory that describes the electrodynamic properties of superconductivity. It is commonly applied in the research field of optical spectroscopy on superconductors.

An electric dipole transition is the dominant effect of an interaction of an electron in an atom with the electromagnetic field.

The Elliott formula describes analytically, or with few adjustable parameters such as the dephasing constant, the light absorption or emission spectra of solids. It was originally derived by Roger James Elliott to describe linear absorption based on properties of a single electron–hole pair. The analysis can be extended to a many-body investigation with full predictive powers when all parameters are computed microscopically using, e.g., the semiconductor Bloch equations or the semiconductor luminescence equations.

<span class="mw-page-title-main">Phonovoltaic</span>

A phonovoltaic (pV) cell converts vibrational (phonons) energy into a direct current much like the photovoltaic effect in a photovoltaic (PV) cell converts light (photon) into power. That is, it uses a p-n junction to separate the electrons and holes generated as valence electrons absorb optical phonons more energetic than the band gap, and then collects them in the metallic contacts for use in a circuit. The pV cell is an application of heat transfer physics and competes with other thermal energy harvesting devices like the thermoelectric generator.

In solid-state physics, band bending refers to the process in which the electronic band structure in a material curves up or down near a junction or interface. It does not involve any physical (spatial) bending. When the electrochemical potential of the free charge carriers around an interface of a semiconductor is dissimilar, charge carriers are transferred between the two materials until an equilibrium state is reached whereby the potential difference vanishes. The band bending concept was first developed in 1938 when Mott, Davidov and Schottky all published theories of the rectifying effect of metal-semiconductor contacts. The use of semiconductor junctions sparked the computer revolution in 1990. Devices such as the diode, the transistor, the photocell and many more still play an important role in technology.

References

  1. H. Luth, Solid Surfaces, Interfaces, and Films, Springer-Verlag Berlin Heidelberg, New York, NY, 2001.
  2. Heine, Volker (1965-06-14). "Theory of Surface States". Physical Review. American Physical Society (APS). 138 (6A): A1689–A1696. Bibcode:1965PhRv..138.1689H. doi:10.1103/physrev.138.a1689. ISSN   0031-899X.
  3. Tejedor, C; Flores, F; Louis, E (1977-06-28). "The metal-semiconductor interface: Si (111) and zincblende (110) junctions". Journal of Physics C: Solid State Physics. IOP Publishing. 10 (12): 2163–2177. Bibcode:1977JPhC...10.2163T. doi:10.1088/0022-3719/10/12/022. ISSN   0022-3719.
  4. Tersoff, J. (1984-10-15). "Theory of semiconductor heterojunctions: The role of quantum dipoles". Physical Review B. American Physical Society (APS). 30 (8): 4874–4877. Bibcode:1984PhRvB..30.4874T. doi:10.1103/physrevb.30.4874. ISSN   0163-1829.
  5. Tersoff, J. (1985-11-15). "Schottky barriers and semiconductor band structures". Physical Review B. American Physical Society (APS). 32 (10): 6968–6971. Bibcode:1985PhRvB..32.6968T. doi:10.1103/physrevb.32.6968. ISSN   0163-1829. PMID   9936825.
  6. L. Pauling, The Nature of the Chemical Bond. Cornell University Press, Ithaca, 1960.
  7. Hannay, N. Bruce; Smyth, Charles P. (1946). "The Dipole Moment of Hydrogen Fluoride and the Ionic Character of Bonds". Journal of the American Chemical Society. American Chemical Society (ACS). 68 (2): 171–173. doi:10.1021/ja01206a003. ISSN   0002-7863.
  8. Garcia-Moliner, Federico and Flores, Fernando, Introduction to the theory of solid surfaces, Cambridge University Press, Cambridge, London, 1979.