Electrokinematics theorem

The electrokinematics theorem[1][2][3] connects the velocity and the charge of carriers moving inside an arbitrary volume to the currents, voltages and power on its surface through an arbitrary irrotational vector. Since it contains, as a particular application, the Ramo-Shockley theorem,[4][5] the electrokinematics theorem is also known as Ramo-Shockly-Pellegrini theorem.

Statement

To introduce the electrokinematics theorem let us first list a few definitions: qj, rj and vj are the electric charge, position and velocity, respectively, at the time t of the jth charge carrier; , and are the electric potential, field, and permittivity, respectively, , and are the conduction, displacement and, in a 'quasi-electrostatic' assumption, total current density, respectively; is an arbitrary irrotational vector in an arbitrary volume enclosed by the surface S, with the constraint that . Now let us integrate over the scalar product of the vector by the two members of the above-mentioned current equation. Indeed, by applying the divergence theorem, the vector identity , the above-mentioned constraint and the fact that , we obtain the electrokinematics theorem in the first form

,

which, taking into account the corpuscular nature of the current , where is the Dirac delta function and N(t) is the carrier number in at the time t, becomes

.

A component of the total electric potential is due to the voltage applied to the kth electrode on S, on which (and with the other boundary conditions on the other electrodes and for ), and each component is due to the jth charge carrier qj , being for and over any electrode and for . Moreover, let the surface S enclosing the volume consist of a part covered by n electrodes and an uncovered part .

According to the above definitions and boundary conditions, and to the superposition theorem, the second equation can be split into the contributions

,
,

relative to the carriers and to the electrode voltages, respectively, being the total number of carriers in the space, inside and outside , at time t, and . The integrals of the above equations account for the displacement current, in particular across .

Current and capacitance

One of the more meaningful application of the above equations is to compute the current

,

through an hth electrode of interest corresponding to the surface , and being the current due to the carriers and to the electrode voltages, to be computed through third and fourth equations, respectively.

Open devices

Let us consider as a first example, the case of a surface S that is not completely covered by electrodes, i.e., , and let us choose Dirichlet boundary conditions on the hth electrode of interest and of on the other electrodes so that, from the above equations we have

,

where is relative to the above-mentioned boundary conditions and is a capacitive coefficient of the hth electrode given by

.

is the voltage difference between the hth electrode and an electrode held to a constant voltage (DC), for instance, directly connected to ground or through a DC voltage source. The above equations hold true for the above Dirichlet conditions for and for any other choice of boundary conditions on .

A second case can be that in which also on so that such equations reduce to

,
.

As a third case, exploiting also to the arbitrariness of , we can choose a Neumann boundary condition of tangent to in any point. Then the equations become

,
.

In particular, this case is useful when the device is a right parallelepiped, being and the lateral surface and the bases, respectively.

As a fourth application let us assume in the whole the volume , i.e., in it, so that from the first equation of Section 1 we have

,

which recover the Kirchhoff law with the inclusion the displacement current across the surface that is not covered by electrodes.

Enclosed devices

A fifth case, historically significant, is that of electrodes that completely enclose the volume of the device, i.e. . Indeed, choosing again the Dirichlet boundary conditions of on and on the other electrodes, from the equations for the open device we get the relationships

,

with

,

thus obtaining the Ramo-Shockly theorem as a particular application of the electrokinematics theorem, extended from the vacuum devices to any electrical component and material.

As the above relationships hold true also when depends on time, we can have a sixty application if we select as the electric field generated by the electrode voltages when there is no charge in . Indeed, as the first equation can be written in the form

,

from which we have

,

where corresponds to the power entering the device across the electrodes (enclosing it). On the other side

,

gives the increment of the internal energy in in a unit of time, being the total electric field of which is due the electrodes and is due to the whole charge density in with over S. Then it is , so that, according to such equations, we also verify the energy balance by means of the electrokinematics theorem. With the above relationships the balance can be extended also to the open devices by taking into account the displacement current across .

Fluctuations

A meaningful application of the above results is also the computation of the fluctuations of the current when the electrode voltages is constant, because this is useful for the evaluation of the device noise. To this end, we can exploit the first equation of section Open devices, because it concerns the more general case of an open device and it can be reduced to a more simply relationship. This happens for frequencies , ( being the transit time of the jth carrier across the device) because the in time integral of the above equation of the Fourier transform to be performed to compute the power spectral density (PSD) of the noise, the time derivatives provides no contribution. Indeed, according to the Fourier transform, this result derives from integrals such as , in which . Therefore, for the PSD computation we can exploit the relationships

Moreover, as it can be shown,[6] this happens also for , for instance when the jth carrier is stored for a long time in a trap if the screening length due to the other carriers is small in comparison to size. All the above considerations hold true for any size of , including nanodevices. In particular we have a meaningful case when the device is a right parallelepiped or cylinder with as lateral surface and u as the unit vector along its axis, with the bases and located at a distance L as electrodes, and with . Indeed, choosing , from the above equation we finally obtain the current ,

,

where and are the components of and along . The above equations in their corpuscular form are particularly suitable for the investigation of transport and noise phenomena from the microscopic point of view, with the application of both the analytical approaches and numerical statistical methods, such as the Monte Carlo techniques. On the other side, in their collective form of the last terms, they are useful to connect, with a general and new method, the local variations of continuous quantities to the current fluctuation at the device terminals. This will be shown in the next sections.

Noise

Shot noise

Let us first evaluate the PSD of the shot noise of the current for short circuited device terminals, i.e. when the 's are constant, by applying the third member of the first equation of the above Section. To this end, let us exploit the Fourier coefficient

and the relationship

where , in the second term and in the third. If we define with and the beginning and the end of the jth carrier motion inside , we have either and or vice versa (the case of give no contribution), so that from the first equations of the above and of this Section, we get

,

where is the number of the carriers (with equal charge q) that start from (arrive on) the electrode of interest during the time interval . Finally for , being the correlation time, and for carriers with a motion that is statistically independent and a Poisson process we have , and so that we obtain

,

where is the average current due to the carriers leaving (reaching) the electrode. Therefore, we recover and extend the Schottky's theorem[7] on shot noise. For instance for an ideal pn junction, or Schottky barrier diode, it is , , where is the Boltzmann constant, T the absolute temperature, v the voltage and the total current. In particular, for the conductance becomes and the above equation gives

,

that is the thermal noise at thermal equilibrium given by the Nyquist theorem.[8] If the carrier motions are correlated, the above equation has to be changed to the form (for )

,

where is the so-called Fano factor that can be both less than 1 (for instance in the case of carrier generation-recombination in nonideal pn junctions[9]), and greater than 1 (as in the negative resistance region of resonant-tunneling diode, as a result of the electron-electron interaction being enhanced by the particular shape of the density of states in the well.[2][10])

Thermal noise

Once again from the corpuscular point of view, let us evaluate the thermal noise with the autocorrelation function of by means of the second term of the second equation of section Fluctuations, that for the short circuit condition (i.e., at thermal equilibrium) which implies , becomes

,

where m is the carrier effective mass and . As and are the carrier mobility and the conductance of the device, from the above equation and the Wiener-Khintchine theorem[11][12] we recover the result

,

obtained by Nyquist from the second principle of the thermodynamics, i.e. by means of a macroscopic approach.[8]

Generation-recombination (g-r) noise

A significant example of application of the macroscopic point of view expressed by the third term of the second equation of section Fluctuations is represented by the g-r noise generated by the carrier trapping-detrapping processes in device defects. In the case of constant voltages and drift current density , that is by neglecting the above velocity fluctuations of thermal origin, from the mentioned equation we get

,

in which is the carrier density, and its steady state value is , being the device cross-section surface; furthermore, we use the same symbols for both the time averaged and the instantaneous quantities. Let us first evaluate the fluctuations of the current i, that from the above equation are

,

where only the fluctuation terms are time dependent. The mobility fluctuations can be due to the motion or to the change of status of defects that we neglect here. Therefore, we ascribe the origin of g-r noise to the trapping-detrapping processes that contribute to through the other two terms via the fluctuation of the electron number in the energy level of a single trap in the channel or in its neighborhood. Indeed, the charge fluctuation in the trap generates variations of and of . However, the variation does not contribute to because it is odd in the u direction, so that we get

,

from which we obtain

,

where the reduction of the integration volume from to the much smaller one around the defect is justified by the fact that the effects of and fade within a few multiples of a screening length, which can be small (of the order of nanometers[7] in graphene[11]); from Gauss's theorem, we obtain also and the r.h.s. of the equation. In it the variation occurs around the average value given by the Fermi-Dirac factor , being the Fermi level. The PSD of the fluctuation due to a single trap then becomes , where is the Lorentzian PSD of the random telegraph signal [13] and is the trap relaxation time. Therefore, for a density of equal and uncorrelated defects we have a total PSD of the g-r noise given by

.

Flicker noise

When the defects are not equal, for any distribution of (except a sharply peaked one, as in the above case of g-r noise), and even for a very small number of traps with large , the total PSD of i, corresponding to the sum of the PSD of all the (statistically independent) traps of the device, becomes[14]

,

where down to the frequency , being the largest and a proper coefficient. In particular, for unipolar conducting materials (e.g., for electrons as carriers) it can be and, for trap energy levels , from we also have , so that from the above equation we obtain,[6]

,

where is the total number of the carriers and is a parameters that depends on the material, structure and technology of the device.

Extensions

Electromagnetic field

The shown electrokinetics theorem holds true in the 'quasi electrostatic' condition, that is when the vector potential can be neglected or, in other terms, when the squared maximum size of is much smaller than the squared minimum wavelength of the electromagnetic field in the device. However it can be extended to the electromagnetic field in a general form.[2] In this general case, by means of the displacement current across the surface it is possible, for instance, to evaluate the electromagnetic field radiation from an antenna. It holds true also when the electric permittivity and the magnetic permeability depend on the frequency. Moreover, the field other than the electric field in 'quasi electrostatic' conditions, can be any other physical irrotational field.

Quantum mechanics

Finally, the electrokinetics theorem holds true in the classical mechanics limit, because it requires the simultaneous knowledge of the position and velocity of the carrier, that is, as a result of the uncertainty principle, when its wave function is essentially non null in a volume smaller than that of device. Such a limit can however be overcome computing the current density according to the quantum mechanical expression.[2][3]

Notes

References

  1. Pellegrini, B. (1986), "Electric charge motion, induced current, energy balance, and noise", Phys. Rev. B 34: 5921-5924.
  2. 1 2 3 4 Pellegrini, B.(1993), "Extension of the electrokinematics theorem to the electromagnetic field and quantum mechanics", Il Nuovo Cimento 15 D: 855–879.
  3. 1 2 Pellegrini, B.(1993), "Elementary application of quaantum-electrokinematics theorem to the electromagnetic field and quantum mechanics", Il Nuovo Cimento 15 D: 881-896.
  4. Ramo, S.(1939), "Currents induced by electron motion", Proc. IRE 27: 584–585.
  5. Shockley, W. (1938), Currents to conductors induced by a moving point charge, J. App. Phys. 9: 635-636.
  6. 1 2 Pellegrini, B. (2013), " noise in graphene", The European Physical Journal B:373-385.
  7. 1 2 Schottky, W. (1918). "Über spontane stromschwankungen in verschiedenen elektrizitätsleitern". Annalen der Physik 57: 541–567.
  8. 1 2 Nyquist,H. (1928). Thermal Agitation of Electric Charge in Conductors. Phys. Rev. 32: 110–113.
  9. Maione, I. A. Pellegrini, B., Fiori, G. Macucci, M. , Guidi, L. and Basso, G.(2011). Shot noise suppression in p-n junctions due to carrier generation-recombination. Phys. Rev. B 83, 155309 –155317.
  10. Iannaccone, G., Lombardi, G., Macucci, M. and Pellegrini B. (1998), Enhanced shot noise in resonant tunneling: theory and experiment. Phys. Rev. Lett. 80: 1054-1058.
  11. 1 2 Wiener, N. (1930), "Generalized Harmonic Analysis", Acta mathematica:118-242.
  12. Khintchine, A.(1934), "Korrelationstheorie der stationären stochastischen Prozesse", Mathematische Annalen 109: 604–615.
  13. Machlup, S. (1954), "Noise in semiconductor: spectrum of two parameter random signal", J. Appl. Phys. 25: 341-343.
  14. Pellegrini, B. (2000), "A general model of noise", Microelectronics Reliability 40: 1775-1780.

See also

This article is issued from Wikipedia - version of the 9/21/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.