Space charge

Last updated

Space charge is an interpretation of a collection of electric charges in which excess electric charge is treated as a continuum of charge distributed over a region of space (either a volume or an area) rather than distinct point-like charges. This model typically applies when charge carriers have been emitted from some region of a solid—the cloud of emitted carriers can form a space charge region if they are sufficiently spread out, or the charged atoms or molecules left behind in the solid can form a space charge region.

Contents

Space charge effects are most pronounced in dielectric media (including vacuum); in highly conductive media, the charge tends to be rapidly neutralized or screened. The sign of the space charge can be either negative or positive. This situation is perhaps most familiar in the area near a metal object when it is heated to incandescence in a vacuum. This effect was first observed by Thomas Edison in light bulb filaments, where it is sometimes called the Edison effect. Space charge is a significant phenomenon in many vacuum and solid-state electronic devices.

Cause

Physical explanation

When a metal object is placed in a vacuum and is heated to incandescence, the energy is sufficient to cause electrons to "boil" away from the surface atoms and surround the metal object in a cloud of free electrons. This is called thermionic emission. The resulting cloud is negatively charged, and can be attracted to any nearby positively charged object, thus producing an electric current which passes through the vacuum.

Space charge can result from a range of phenomena, but the most important are:

  1. Combination of the current density and spatially inhomogeneous resistance
  2. Ionization of species within the dielectric to form heterocharge
  3. Charge injection from electrodes and from a stress enhancement
  4. Polarization in structures such as water trees. "Water tree" is a name given to a tree-like figure appearing in a water-impregnated polymer insulating cable. [1] [2]

It has been suggested that in alternating current (AC) most carriers injected at electrodes during a half cycle are ejected during the next half cycle, so the net balance of charge on a cycle is practically zero. However, a small fraction of the carriers can be trapped at levels deep enough to retain them when the field is inverted. The amount of charge in AC should increase slower than in direct current (DC) and become observable after longer periods of time.

Hetero and homo charge

Hetero charge means that the polarity of the space charge is opposite to that of neighboring electrode, and homo charge is the reverse situation. Under high voltage application, a hetero charge near the electrode is expected to reduce the breakdown voltage, whereas a homo charge will increase it. After polarity reversal under ac conditions, the homo charge is converted to hetero space charge.

Mathematical explanation

If the near "vacuum" has a pressure of 10−6 mmHg or less, the main vehicle of conduction is electrons. The emission current density (J) from the cathode, as a function of its thermodynamic temperature T, in the absence of space-charge, is given by Richardson's law:

where

The reflection coefficient can be as low as 0.105 but is usually near 0.5. For tungsten, (1 − ř)A0 = (0.6 to 1.0)×106 A⋅m−2⋅K−2, and φ = 4.52 eV. At 2500 °C, the emission is 28207 A/m2.

The emission current as given above is many times greater than that normally collected by the electrodes, except in some pulsed valves such as the cavity magnetron. Most of the electrons emitted by the cathode are driven back to it by the repulsion of the cloud of electrons in its neighborhood. This is called the space charge effect. In the limit of large current densities, J is given by the Child–Langmuir equation below, rather than by the thermionic emission equation above.

Occurrence

Space charge is an inherent property of all vacuum tubes. This has at times made life harder or easier for electrical engineers who used tubes in their designs. For example, space charge significantly limited the practical application of triode amplifiers which led to further innovations such as the vacuum tube tetrode.

On the other hand, space charge was useful in some tube applications because it generates a negative EMF within the tube's envelope, which could be used to create a negative bias on the tube's grid. Grid bias could also be achieved by using an applied grid voltage in addition to the control voltage. This could improve the engineer's control and fidelity of amplification. It allowed the construction of space charge tubes for car radios that required only 6 or 12 volts anode voltage (typical examples were the 6DR8/EBF83, 6GM8/ECC86, 6DS8/ECH83, 6ES6/EF97 and 6ET6/EF98).

Space charges can also occur within dielectrics. For example, when gas near a high voltage electrode begins to undergo dielectric breakdown, electrical charges are injected into the region near the electrode, forming space charge regions in the surrounding gas. Space charges can also occur within solid or liquid dielectrics that are stressed by high electric fields. Trapped space charges within solid dielectrics are often a contributing factor leading to dielectric failure within high voltage power cables and capacitors.

In semiconductor physics, space charge layers that are depleted of charge carriers are used as a model to explain the rectifying behaviour of p–n junctions and the buildup of a voltage in photovoltaic cells.

Space-charge-limited current

In vacuum (Child's law)

Graph showing Child-Langmuir law. S and d are constant and equal to 1. ChildLangmuirLawGraph1.png
Graph showing Child–Langmuir law. S and d are constant and equal to 1.

First proposed by Clement D. Child in 1911, Child's law states that the space-charge-limited current (SCLC) in a plane-parallel vacuum diode varies directly as the three-halves power of the anode voltage and inversely as the square of the distance d separating the cathode and the anode. [3]

For electrons, the current density J (amperes per meter squared) is written:

where is the anode current and S the surface area of the anode receiving the current; is the magnitude of the charge of the electron and is its mass. The equation is also known as the "three-halves-power law" or the Child–Langmuir law. Child originally derived this equation for the case of atomic ions, which have much smaller ratios of their charge to their mass. Irving Langmuir published the application to electron currents in 1913, and extended it to the case of cylindrical cathodes and anodes. [4]

The equation's validity is subject to the following assumptions:

  1. Electrons travel ballistically between electrodes (i.e., no scattering).
  2. In the interelectrode region, the space charge of any ions is negligible.
  3. The electrons have zero velocity at the cathode surface.

The assumption of no scattering (ballistic transport) is what makes the predictions of Child–Langmuir law different from those of Mott–Gurney law. The latter assumes steady-state drift transport and therefore strong scattering.

Child's law was further generalized by Buford R. Conley in 1995 for the case of non-zero velocity at the cathode surface with the following equation: [5]

where is the initial velocity of the particle. This equation reduces to Child's Law for the special case of equal to zero.

In recent years, various models of SCLC current have been revised as reported in two review papers. [6] [7]

In semiconductors

In semiconductors and insulating materials, an electric field causes charged particles, electrons, to reach a specific drift velocity that is parallel to the direction of the field. This is different from the behavior of the free charged particles in a vacuum, in which a field accelerates the particle. The proportionality factor between the magnitudes of the drift velocity, , and the electric field, , is called the mobility, :

Drift regime (Mott–Gurney law)

The Child's law behavior of a space-charge-limited current that applies in a vacuum diode doesn't generally apply to a semiconductor/insulator in a single-carrier device, and is replaced by the Mott–Gurney law. For a thin slab of material of thickness , sandwiched between two selective Ohmic contacts, the electric current density, , flowing through the slab is given by: [8] [9]

where is the voltage that has been applied across the slab and is the permittivity of the solid. The Mott–Gurney law offers some crucial insight into charge-transport across an intrinsic semiconductor, namely that one should not expect the drift current to increase linearly with the applied voltage, i.e., from Ohm's law, as one would expect from charge-transport across a metal or highly doped semiconductor. Since the only unknown quantity in the Mott–Gurney law is the charge-carrier mobility, , the equation is commonly used to characterize charge transport in intrinsic semiconductors. Using the Mott–Gurney law for characterizing amorphous semiconductors, along with semiconductors containing defects and/or non-Ohmic contacts, should however be approached with caution as significant deviations both in the magnitude of the current and the power law dependence with respect to the voltage will be observed. In those cases the Mott–Gurney law can not be readily used for characterization, and other equations which can account for defects and/or non-ideal injection should be used instead.

During the derivation of the Mott–Gurney law, one has to make the following assumptions:

  1. There is only one type of charge carrier present, i.e., only electrons or holes.
  2. The material has no intrinsic conductivity, but charges are injected into it from one electrode and captured by the other.
  3. The carrier mobility, , and the permittivity, , are constant throughout the sample.
  4. The current flow is not limited by traps or energetic disorder.
  5. The current is not predominantly due to doping.
  6. The electric field at the charge-injecting electrode is zero, meaning that the current is governed by drift only.

Derivation

Consider a crystal of thickness carrying a current . Let be the electric field at a distance from the surface, and the number of electrons per unit volume. Then the current is given has two contributions, one due to drift and the other due to diffusion:

When is the electrons mobility and the diffusion coefficient. Laplace's equation gives for the field:

Hence, eliminating , we have:

After integrating, making use of the Einstein relation and neglecting the term we obtain for the electric field:

where is a constant. We may neglect the term because we are supposing that and .

Since, at , , we have:

It follows that the potential drop across the crystal is:

Making use of ( ) and ( ⁎⁎ ) we can write in terms of . For small , is small and , so that:

Thus the current increases as the square of . For large , and we obtain:

As an application example, the steady-state space-charge-limited current across a piece of intrinsic silicon with a charge-carrier mobility of 1500 cm2/V-s, a relative dielectric constant of 11.9, an area of 10−8 cm2 and a thickness of 10−4 cm can be calculated by an online calculator to be 126.4 μA at 3 V. Note that in order for this calculation to be accurate, one must assume all the points listed above.

In the case where the electron/hole transport is limited by trap states in the form of exponential tails extending from the conduction/valence band edges,

the drift current density is given by the Mark-Helfrich equation, [10]

where is the elementary charge, with being the thermal energy, is the effective density of states of the charge carrier type in the semiconductor, i.e., either or , and is the trap density.

Low voltage regime

In the case where a very small applied bias is applied across the single-carrier device, the current is given by: [11] [12] [13]

Note that the equation describing the current in the low voltage regime follows the same thickness scaling as the Mott–Gurney law, , but increases linearly with the applied voltage.

Saturation regimes

When a very large voltage is applied across the semiconductor, the current can transition into a saturation regime.

In the velocity-saturation regime, this equation takes the following form

Note the different dependence of on between the Mott–Gurney law and the equation describing the current in the velocity-saturation regime. In the ballistic case (assuming no collisions), the Mott–Gurney equation takes the form of the more familiar Child–Langmuir law.

In the charge-carrier saturation regime, the current through the sample is given by,

where is the effective density of states of the charge carrier type in the semiconductor.

Shot noise

Space charge tends to reduce shot noise. [14] Shot noise results from the random arrivals of discrete charge; the statistical variation in the arrivals produces shot noise. [15] A space charge develops a potential that slows the carriers down. For example, an electron approaching a cloud of other electrons will slow down due to the repulsive force. The slowing carriers also increases the space charge density and resulting potential. In addition, the potential developed by the space charge can reduce the number of carriers emitted. [16] When the space charge limits the current, the random arrivals of the carriers are smoothed out; the reduced variation results in less shot noise. [15]

See also

Related Research Articles

<span class="mw-page-title-main">Scanning tunneling microscope</span> Instrument able to image surfaces at the atomic level by exploiting quantum tunneling effects

A scanning tunneling microscope (STM) is a type of microscope used for imaging surfaces at the atomic level. Its development in 1981 earned its inventors, Gerd Binnig and Heinrich Rohrer, then at IBM Zürich, the Nobel Prize in Physics in 1986. STM senses the surface by using an extremely sharp conducting tip that can distinguish features smaller than 0.1 nm with a 0.01 nm (10 pm) depth resolution. This means that individual atoms can routinely be imaged and manipulated. Most scanning tunneling microscopes are built for use in ultra-high vacuum at temperatures approaching absolute zero, but variants exist for studies in air, water and other environments, and for temperatures over 1000 °C.

<span class="mw-page-title-main">Permittivity</span> Measure of the electric polarizability of a dielectric

In electromagnetism, the absolute permittivity, often simply called permittivity and denoted by the Greek letter ε (epsilon), is a measure of the electric polarizability of a dielectric. A material with high permittivity polarizes more in response to an applied electric field than a material with low permittivity, thereby storing more energy in the material. In electrostatics, the permittivity plays an important role in determining the capacitance of a capacitor.

<span class="mw-page-title-main">Capacitance</span> Ability of a body to store an electrical charge

Capacitance is the capability of a material object or device to store electric charge. It is measured by the charge in response to a difference in electric potential, expressed as the ratio of those quantities. Commonly recognized are two closely related notions of capacitance: self capacitance and mutual capacitance. An object that can be electrically charged exhibits self capacitance, for which the electric potential is measured between the object and ground. Mutual capacitance is measured between two components, and is particularly important in the operation of the capacitor, an elementary linear electronic component designed to add capacitance to an electric circuit.

<span class="mw-page-title-main">Noether's theorem</span> Statement relating differentiable symmetries to conserved quantities

Noether's theorem or Noether's first theorem states that every differentiable symmetry of the action of a physical system with conservative forces has a corresponding conservation law. The theorem was proven by mathematician Emmy Noether in 1915 and published in 1918. The action of a physical system is the integral over time of a Lagrangian function, from which the system's behavior can be determined by the principle of least action. This theorem only applies to continuous and smooth symmetries over physical space.

<span class="mw-page-title-main">Fermi gas</span> Physical model of gases composed of many non-interacting identical fermions

A Fermi gas is an idealized model, an ensemble of many non-interacting fermions. Fermions are particles that obey Fermi–Dirac statistics, like electrons, protons, and neutrons, and, in general, particles with half-integer spin. These statistics determine the energy distribution of fermions in a Fermi gas in thermal equilibrium, and is characterized by their number density, temperature, and the set of available energy states. The model is named after the Italian physicist Enrico Fermi.

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

<span class="mw-page-title-main">Ampère's circuital law</span> Concept in classical electromagnetism

In classical electromagnetism, Ampère's circuital law relates the circulation of a magnetic field around a closed loop to the electric current passing through the loop.

In the calculus of variations and classical mechanics, the Euler–Lagrange equations are a system of second-order ordinary differential equations whose solutions are stationary points of the given action functional. The equations were discovered in the 1750s by Swiss mathematician Leonhard Euler and Italian mathematician Joseph-Louis Lagrange.

<span class="mw-page-title-main">Displacement current</span> Physical quantity in electromagnetism

In electromagnetism, displacement current density is the quantity D/∂t appearing in Maxwell's equations that is defined in terms of the rate of change of D, the electric displacement field. Displacement current density has the same units as electric current density, and it is a source of the magnetic field just as actual current is. However it is not an electric current of moving charges, but a time-varying electric field. In physical materials, there is also a contribution from the slight motion of charges bound in atoms, called dielectric polarization.

p–n junction Semiconductor–semiconductor junction

A p–n junction is a boundary or interface between two types of semiconductor materials, p-type and n-type, inside a single crystal of semiconductor. The "p" (positive) side contains an excess of holes, while the "n" (negative) side contains an excess of electrons in the outer shells of the electrically neutral atoms there. This allows electric current to pass through the junction only in one direction. The p- and n-type regions creating the junction are made by doping the semiconductor, for example by ion implantation, diffusion of dopants, or by epitaxy.

In solid-state physics, the electron mobility characterises how quickly an electron can move through a metal or semiconductor when pulled by an electric field. There is an analogous quantity for holes, called hole mobility. The term carrier mobility refers in general to both electron and hole mobility.

In plasmas and electrolytes, the Debye length, is a measure of a charge carrier's net electrostatic effect in a solution and how far its electrostatic effect persists. With each Debye length the charges are increasingly electrically screened and the electric potential decreases in magnitude by 1/e. A Debye sphere is a volume whose radius is the Debye length. Debye length is an important parameter in plasma physics, electrolytes, and colloids. The corresponding Debye screening wave vector for particles of density , charge at a temperature is given by in Gaussian units. Expressions in MKS units will be given below. The analogous quantities at very low temperatures are known as the Thomas–Fermi length and the Thomas–Fermi wave vector. They are of interest in describing the behaviour of electrons in metals at room temperature.

In semiconductor physics, the Haynes–Shockley experiment was an experiment that demonstrated that diffusion of minority carriers in a semiconductor could result in a current. The experiment was reported in a short paper by Haynes and Shockley in 1948, with a more detailed version published by Shockley, Pearson, and Haynes in 1949. The experiment can be used to measure carrier mobility, carrier lifetime, and diffusion coefficient.

In electrical engineering, dielectric loss quantifies a dielectric material's inherent dissipation of electromagnetic energy. It can be parameterized in terms of either the loss angleδ or the corresponding loss tangenttan(δ). Both refer to the phasor in the complex plane whose real and imaginary parts are the resistive (lossy) component of an electromagnetic field and its reactive (lossless) counterpart.

The word electricity refers generally to the movement of electrons through a conductor in the presence of a potential difference or an electric field. The speed of this flow has multiple meanings. In everyday electrical and electronic devices, the signals travel as electromagnetic waves typically at 50%–99% of the speed of light in vacuum, while the electrons themselves move much more slowly; see drift velocity and electron mobility.

<span class="mw-page-title-main">Mathematical descriptions of the electromagnetic field</span> Formulations of electromagnetism

There are various mathematical descriptions of the electromagnetic field that are used in the study of electromagnetism, one of the four fundamental interactions of nature. In this article, several approaches are discussed, although the equations are in terms of electric and magnetic fields, potentials, and charges with currents, generally speaking.

When an electromagnetic wave travels through a medium in which it gets attenuated, it undergoes exponential decay as described by the Beer–Lambert law. However, there are many possible ways to characterize the wave and how quickly it is attenuated. This article describes the mathematical relationships among:

Diffusion current is a current in a semiconductor caused by the diffusion of charge carriers. This is the current which is due to the transport of charges occurring because of non-uniform concentration of charged particles in a semiconductor. The drift current, by contrast, is due to the motion of charge carriers due to the force exerted on them by an electric field. Diffusion current can be in the same or opposite direction of a drift current. The diffusion current and drift current together are described by the drift–diffusion equation.

In quantum field theory, a Ward–Takahashi identity is an identity between correlation functions that follows from the global or gauge symmetries of the theory, and which remains valid after renormalization.

The Wannier equation describes a quantum mechanical eigenvalue problem in solids where an electron in a conduction band and an electronic vacancy within a valence band attract each other via the Coulomb interaction. For one electron and one hole, this problem is analogous to the Schrödinger equation of the hydrogen atom; and the bound-state solutions are called excitons. When an exciton's radius extends over several unit cells, it is referred to as a Wannier exciton in contrast to Frenkel excitons whose size is comparable with the unit cell. An excited solid typically contains many electrons and holes; this modifies the Wannier equation considerably. The resulting generalized Wannier equation can be determined from the homogeneous part of the semiconductor Bloch equations or the semiconductor luminescence equations.

References

  1. Moreau, E.; Mayoux, C.; Laurent, C.; Boudet, A. (February 1993), "The Structural Characteristics of Water Trees in Power Cables and Laboratory Specimens", IEEE Transactions on Electrical Insulation, IEEE, 28 (1): 54–64, doi:10.1109/14.192240, ISSN   0018-9367
  2. Hennuy, Blandine; Marginet, Joachim; François, Alain; Platbrood, Gérard; Tits, Yvan; De Clerck, Quentin (June 2009). Water Trees in Medium Voltage XLPE Cables: Very Short Time Accelerated Ageing Tests (PDF). 20th International Conference on Electricity Distribution (CIRED2009). Prague. Paper 1060.
  3. Child, C. D. (1 May 1911). "Discharge From Hot CaO". Physical Review. Series I. 32 (5): 492–511. Bibcode:1911PhRvI..32..492C. doi:10.1103/PhysRevSeriesI.32.492.
  4. Langmuir, Irving (1913). "The Effect of Space Charge and Residual Gases on Thermionic Currents in High Vacuum". Physical Review. 2 (6): 450–486. Bibcode:1913PhRv....2..450L. doi:10.1103/PhysRev.2.450.
  5. Conley, Buford Ray (May 1995). "Utilization of Ambient Gas as a Propellant for Low Earth Orbit Electric Propulsion" (PDF). Masters Thesis, Massachusetts Institute of Technology, Cambridge, MA: Page 24, equation 3.43 via dspace.mit.edu.
  6. P. Zhang, A. Valfells, L. K. Ang, J. W. Luginsland and Y. Y. Lau (2017). "100 years of the physics of diodes". Applied Physics Reviews . 4 (1): 011304. Bibcode:2017ApPRv...4a1304Z. doi: 10.1063/1.4978231 .{{cite journal}}: CS1 maint: multiple names: authors list (link)
  7. P Zhang, Y. S. Ang, A. L. Garner, A. Valfells, J. L. Luginsland, and L. K. Ang (2021). "Space–charge limited current in nanodiodes: Ballistic, collisional, and dynamical effects". Journal of Applied Physics . 129 (10): 100902. Bibcode:2021JAP...129j0902Z. doi:10.1063/5.0042355. hdl: 20.500.11815/2643 . S2CID   233643434.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  8. Mott, Nevill F.; Gurney, R. W. (1940). Electronic Processes in Ionic Crystals, 1st ed. Oxford University Press.
  9. Murgatroyd, P. N. J. (1970). "Theory of space-charge-limited current enhanced by Frenkel effect". J. Phys. D. 3 (2): 151. Bibcode:1970JPhD....3..151M. doi:10.1088/0022-3727/3/2/308. S2CID   250765910.
  10. Mark, P.; Helfrich, W. (1962). "Space-Charge-Limited Currents in Organic Crystals". Journal of Applied Physics. 33 (1): 205–215. Bibcode:1962JAP....33..205M. doi:10.1063/1.1728487.
  11. de Levie, R.; Seidah, N. G.; Moreira, H. (1972). "Transport of Ions of One Kind through Thin Membranes". J. Membrane Biol. 10 (2): 171–92. doi:10.1007/BF01867852. PMID   4669446. S2CID   33548484.
  12. van Mensfoort, S.; Coehoorn, R (2008). "Effect of Gaussian Disorder on the Voltage Dependence of the Current Density in Sandwich-Type Devices based on Organic Semiconductors". Physical Review B. 78 (8): 085207(16). Bibcode:2008PhRvB..78h5207V. doi:10.1103/PhysRevB.78.085207.
  13. Röhr, J. A.; Kirchartz, T.; Nelson, J. (2017). "On the Correct Interpretation of the Low Voltage Regime in Intrinsic Single-Carrier Devices". Journal of Physics: Condensed Matter. 29 (20): 205901. Bibcode:2017JPCM...29t5901R. doi:10.1088/1361-648X/aa66cc. PMID   28294108. S2CID   46817172.
  14. Terman, Frederick Emmons (1943), Radio Engineers' Handbook (first ed.), New York: McGraw-Hill, pp. 286–294
  15. 1 2 Terman 1943 , pp. 292–293
  16. Terman 1943 , pp. 286–287