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; A_{0}, E=-\nabla A_{0} and \varepsilon are the electric potential, field, and permittivity, respectively, J_{q}, J_{d}=\varepsilon \partial E/ \partial t and J=J_{q}+J_{d} are the conduction, displacement and, in a 'quasi-electrostatic' assumption, total current density, respectively; F=-\nabla \Phi is an arbitrary irrotational vector in an arbitrary volume \Omega enclosed by the surface S, with the constraint that \nabla (\varepsilon F)=0. Now let us integrate over \Omega the scalar product of the vector F by the two members of the above-mentioned current equation. Indeed, by applying the divergence theorem, the vector identity a\cdot\nabla\gamma=\nabla\cdot(\gamma a)-\gamma\nabla\cdot a, the above-mentioned constraint and the fact that \nabla\cdot J=0, we obtain the electrokinematics theorem in the first form

-\int_{S} \Phi J\cdot dS=\int_{\Omega}J_{q}\cdot Fd^{3}r-\int_{S}\varepsilon\frac{\partial A_{0}}{\partial t}F\cdot dS ,

which, taking into account the corpuscular nature of the current J_{q}=\sum_{j=1}^{N(t)} q_{j}\delta(r-r_{j})v_{j}, where \delta(r-r_{j}) is the Dirac delta function and N(t) is the carrier number in \Omega at the time t, becomes

-\int_{S} \Phi J\cdot dS=\sum_{j=1}^{N(t)} q_{j}v_{j}\cdot F(r_{j})-\int_{S}\varepsilon\frac{\partial A_{0}}{\partial t}F\cdot dS .

A component A_{Vk}[r,V_{k}(t)]=V_{k}(t)\psi_{k}(r) of the total electric potential A_{0}=A_{Vk}+A_{qj} is due to the voltage V_{k}(t) applied to the kth electrode on S, on which \psi_{k}(r)=1 (and with the other boundary conditions \psi_{k}(r)=\psi_{k}(\infty)=0 on the other electrodes and for r \to \infty), and each component A_{qj}[r,r_{j}(t)] is due to the jth charge carrier qj , being A_{qj}[r,r_{j}(t)]=0 for r and r_j(t) over any electrode and for r \to \infty. Moreover, let the surface S enclosing the volume \Omega consist of a part S_{E}=\sum_{k=1}^{n}S_{k} covered by n electrodes and an uncovered part S_R.

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

-\int_{S_{E}} \Phi J_{q}\cdot dS=\sum_{j=1}^{N(t)} q_{j}v_{j}\cdot F(r_{j})+\sum_{j=1}^{M(t)}\int_{S_{R}}\varepsilon(\Phi\frac{\partial E_{j}}{\partial t}-\frac{\partial A_{qj}}{\partial t}F)\cdot dS ,
-\int_{S_{E}} \Phi J_{V}\cdot dS=\sum_{k=1}^n \int_{S_{R}}\varepsilon \Phi \frac{\partial E_{k}}{\partial t} \cdot dS-\sum_{k=1}^n \int_{S}\varepsilon \frac{\partial A_{Vk}}{\partial t} F\cdot dS,

relative to the carriers and to the electrode voltages, respectively, M(t) being the total number of carriers in the space, inside and outside \Omega, at time t, E_{j}=-\nabla A_{qj} and E_{k}=-\nabla A_{Vk}. The integrals of the above equations account for the displacement current, in particular across S_R.

Current and capacitance

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

i_{h}\equiv-\int_{S_{h}}J\cdot dS=i_{qh}+i_{Vh} ,

through an hth electrode of interest corresponding to the surface S_{h}, i_{qh} and i_{Vh} 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., S_{R} \ne 0, and let us choose Dirichlet boundary conditions \Phi=\Phi_{h}=1 on the hth electrode of interest and of \Phi_{h}=0 on the other electrodes so that, from the above equations we have

i_{qh}=\sum_{j=1}^{N(t)}q_{j}v_{j}\cdot F_{h}(r_{j})+\sum_{j=1}^{M(t)}\int_{S_{R}}\varepsilon(\Phi_{h}\frac{\partial E_{j}}{\partial t}-\frac{\partial A_{qj}}{\partial t}F_{h})\cdot dS=
i_{dh}=\sum_{k=1}^{n}C_{hk}\frac{dV_{k}}{dt} ,

where F=F_{h}(r_{j}) is relative to the above-mentioned boundary conditions and C_{hk} is a capacitive coefficient of the hth electrode given by

C_{hk}=-(\int_{S_{k}}\varepsilon F_{h}\cdot dS+\int_{S_{R}}\varepsilon(\Phi_{h}\nabla\psi_{k}+\psi_{k}F_{h})\cdot dS) .

V_{h} 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 \Phi_h and for any other choice of boundary conditions on S_R.

A second case can be that in which \Phi_{h}=0 also on S_R so that such equations reduce to

i_{qh}=\sum_{j=1}^{N(t)}q_{j}v_{j}\cdot F_h(r_{j})-\sum_{j=1}^{M(t)}\int_{S_{R}}\varepsilon\frac{\partial A_{0j}}{\partial t}F_{h}\cdot dS ,
C_{hk}=-(\int_{S_{k}}\varepsilon F_{h}\cdot dS+\int_{S_{R}}\varepsilon\Psi_{k}F_{h}\cdot dS) .

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

i_{qh}=\sum_{j=1}^{N(t)}q_{j}v_{j}\cdot F_{h}(r_{j})-\sum_{j=1}^{M(t)}\int_{S_{R}}\varepsilon\Phi_{h}\frac{\partial E_{j}}{\partial t}\cdot dS ,
C_{hk}=-(\int_{S_{k}}\varepsilon F_{h}\cdot dS+\int_{S_{R}}\varepsilon\nabla\Psi_{h}\cdot dS) .

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

As a fourth application let us assume \Phi=1 in the whole the volume \Omega, i.e., F=0 in it, so that from the first equation of Section 1 we have

\sum_{h=1}^{n}i_{h}-\int_{S_{R}}\varepsilon(\sum_{j=1}^{M(t)}\frac{\partial E_{j}}{\partial t}+\sum_{k=1}^{n}\frac{\partial E_{k}}{\partial t})\cdot dS=0 ,

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

Enclosed devices

A fifth case, historically significant, is that of electrodes that completely enclose the volume \Omega of the device, i.e. S_{R}=0 . Indeed, choosing again the Dirichlet boundary conditions of \Phi_{h}=1 on S_h and \Phi_{h}=0 on the other electrodes, from the equations for the open device we get the relationships

i_{h}=\sum_{j=1}^{N(t)}q_{j}v_{j}\cdot F_h(r_{j})+\sum_{k=1}^{n}C_{hk}\frac{dV_h}{dt} ,

with

C_{hk}=-\int_{S_k}\varepsilon F_{h}\cdot dS ,

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 F(t) depends on time, we can have a sixty application if we select as F=F_{V}=-\sum_{k=1}^{n}V_{k}(t)\nabla\psi_{k}(r) the electric field generated by the electrode voltages when there is no charge in \Omega. Indeed, as the first equation can be written in the form

-\int_{S}\Phi J\cdot dS=\int_{\Omega}J\cdot Fd^{3}r ,

from which we have

\sum_{h=1}^{n}V_{h}i_{h}=\int_{\Omega}J\cdot F_{V}d^{3}r\equiv W,

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

\int_{\Omega}(E\cdot J_{q}+E\cdot \frac{\varepsilon \partial E}{\partial t})d^{3}r=\int_{\Omega}E\cdot Jd^{3}r\equiv \frac{d\Xi}{dt} ,

gives the increment of the internal energy \Xi in \Omega in a unit of time, E=F_{V}+E_{q} being the total electric field of which F_{V} is due the electrodes and E_{q}=-\nabla \psi_{q}(r,t) is due to the whole charge density in \Omega with \psi_{q}(r,t)=0 over S. Then it is \int_{\Omega}E_{q}\cdot Jd^{3}r=0, so that, according to such equations, we also verify the energy balance W=d\Xi/dt 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 S_R.

Fluctuations

A meaningful application of the above results is also the computation of the fluctuations of the current i_{h}=i_{qh} 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 f=\omega/(2\pi)\ll1/(2\pi t_{j}), (t_{j} 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  \int_{0}^{t_{j}}exp(-j\omega t)(\partial Q/\partial t)dt\approx Q(t_{j})-Q(0) , in which Q(t_{j})=Q(0)=0. Therefore, for the PSD computation we can exploit the relationships

i_{qh}=\sum_{j=1}^{N(t)}q_jv_j\cdot F_h(r_j)=-\sum_{j=1}^{N(t)}q_{j}\frac{d\Phi_{h}[r_{j}(t)]}{dt}=\int_{\Omega}J_{q}\cdot Fd^{3}r

Moreover, as it can be shown,[6] this happens also for f\gg 1/(2\pi t_{j}), for instance when the jth carrier is stored for a long time \tau_{j} in a trap if the screening length due to the other carriers is small in comparison to \Omega size. All the above considerations hold true for any size of \Omega, including nanodevices. In particular we have a meaningful case when the device is a right parallelepiped or cylinder with S_R as lateral surface and u as the unit vector along its axis, with the bases S_{E1} and S_{E2} located at a distance L as electrodes, and with S_{E1}\rightarrow u\rightarrow S_{E2}. Indeed, choosing F_{h}=F=-u/L, from the above equation we finally obtain the current i=i_{1}=i_{q1}=-i_{2},

i=\frac{1}{L}\sum_{j=1}^{N(t)}q_{j}v_{ju}=\frac{1}{L}\int_{\Omega}J_{qu}d^{3}r ,

where v_{ju} and J_{qu} are the components of v and J_{q} along u. 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 S_S of the shot noise of the current i=i_{qh} for short circuited device terminals, i.e. when the V_{h}'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

D(\omega_{l})\equiv\frac{1}{T^'}\int_{-T^'/2}^{T^'/2}\Delta i(t)exp(-j\omega_{l}t)dt

and the relationship

S_{S}(\omega_{l})\equiv\lim_{\Delta f\to 0}\frac{\left \langle D(\omega_l)D^*(\omega_l)\right \rangle}{\Delta f}=\lim_{T^' \to \infty}(2T^'\left \langle D(\omega_l)D^*(\omega_l)\right \rangle)

where \omega_l=l(2\pi /T^'), l=...,-2,-1,1,2,... in the second term and l=1,2,... in the third. If we define with t_{bj} and (t_{bj}+t_j) the beginning and the end of the jth carrier motion inside \Omega, we have either \Phi_h[r_j(t_{bj})]=1 and \Phi_h[r_j(t_{bj}+t_j)]=0 or vice versa (the case of \Phi_h[r_j(t_{bj})]=\Phi_h[r_j(t_{bj}+t_j)] give no contribution), so that from the first equations of the above and of this Section, we get

D(\omega_l)\equiv \frac{q}{T^'}(\Delta N^+-\Delta N^-) ,

where N^+ (N^-) is the number of the carriers (with equal charge q) that start from (arrive on) the electrode of interest during the time interval -T^'/2, T^'/2. Finally for \tau_c\ll t_{jmin}, \tau_c being the correlation time, and for carriers with a motion that is statistically independent and a Poisson process we have \left\langle\Delta N^+\Delta N^-\right \rangle=0, \left\langle\Delta N^+\Delta N^+\right\rangle=\left\langle N^+\right\rangle and \left\langle\Delta N^-\Delta N^-\right\rangle=\left\langle N^-\right\rangle so that we obtain

S_S=2q(I^++I^-) ,

where I^+ (I^-) 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 I^+=I_0exp(qv/k_BT), I^-=I_0, where k_B is the Boltzmann constant, T the absolute temperature, v the voltage and I=I^+-I^- the total current. In particular, for v=0 the conductance becomes g=(dI/dv)=qI_0/(k_BT) and the above equation gives

S_S=4k_BTg ,

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 I^+\gg I^-)

S_S=F_a(2qI) ,

where F_a 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 \left \langle i(t)i(t+\theta)\right\rangle of i(t) by means of the second term of the second equation of section Fluctuations, that for the short circuit condition V_1=V_2=0 (i.e., at thermal equilibrium) which implies N(t)=\overline{N}, becomes

\left \langle i(t)i(t+\theta)\right\rangle=\frac{q^2}{L^2}\sum_{j=1}^{\overline{N}}\left \langle v_{ju}^2(t)\right \rangle_t exp(-\left |\theta\right |/\tau_c)=\frac{q^2\overline{N}k_BT}{L^2m}exp(-\left |\theta\right |/\tau_c) ,

where m is the carrier effective mass and \tau_c\ll\tau_{jmin}. As \mu=q\tau_c/[m(1+j\omega)] and G=q\mu\overline{N}/L^2 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

S_T=4k_BTRe\{G(j\omega)\} ,

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 J_{qu}=q\mu n_qE, (E\equiv E_u), that is by neglecting the above velocity fluctuations of thermal origin, from the mentioned equation we get

i=\frac{1}{L}\int_{\Omega}q\mu n_qEd^3r ,

in which n_q is the carrier density, and its steady state value is \overline{i}\equiv I=q\mu n_qEA, A 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

\frac{\Delta i}{I}=\frac{1}{\Omega}(\frac{1}{n_q}\int_{\Omega}\Delta n_qd^3r+\frac{1}{E}\int_{\Omega}\Delta Ed^3r+\frac{1}{\mu}\int_{\Omega}\Delta \mu d^3r) ,

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 \Delta i through the other two terms via the fluctuation of the electron number \chi=0,1 in the energy level \varepsilon_t of a single trap in the channel or in its neighborhood. Indeed, the charge fluctuation q\Delta\chi in the trap generates variations of n_q and of E. However, the variation \Delta E does not contribute to \Delta i because it is odd in the u direction, so that we get

\frac{\Delta i}{I}=\frac{1}{\Omega n_q}\int_{\Omega}\Delta n_q d^3r ,

from which we obtain

\frac{\Delta i}{I}=\frac{1}{\Omega n_q}\int_{\delta\Omega}\Delta n_q ad^3r=-\frac{1}{\Omega n_q}\Delta\chi ,

where the reduction of the integration volume from \Omega to the much smaller one \delta\Omega around the defect is justified by the fact that the effects of \Delta n_q and \Delta E 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 \int_{\delta\Omega}\Delta n_qd^3r=-\Delta\chi and the r.h.s. of the equation. In it the variation \Delta\chi occurs around the average value \overline{\chi} given by the Fermi-Dirac factor \overline{\chi}\equiv\phi=\{[1+exp[(\varepsilon_t-\varepsilon_f)/k_B T]\}^{-1}, \varepsilon_f being the Fermi level. The PSD S_t of the fluctuation \Delta i due to a single trap then becomes S_t/I^2=[1/(\Omega n_q)]^2S_{\chi}, where S_{\chi}=4\phi(1-\phi)\tau/[1+(\omega\tau)^2] is the Lorentzian PSD of the random telegraph signal \chi[13] and \tau is the trap relaxation time. Therefore, for a density n_t of equal and uncorrelated defects we have a total PSD S_{gr} of the g-r noise given by

S_{gr}=\frac{4I^2n_t\phi(1-\phi)\tau}{\Omega n^2_{q}[1+(\omega\tau)^2]} .

Flicker noise

When the defects are not equal, for any distribution of \tau (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 \tau, the total PSD S_f of i, corresponding to the sum of the PSD S_t of all the n_t\Omega (statistically independent) traps of the device, becomes[14]

S_f=\frac{n_tB}{\Omega n_q^2}\frac{I^2}{f^\gamma} ,

where 0.85<\gamma<1.15 down to the frequency 1/2\pi\tau_M, \tau_M being the largest \tau and B(\varepsilon_f/k_B T) a proper coefficient. In particular, for unipolar conducting materials (e.g., for electrons as carriers) it can be n_q\propto exp(\varepsilon_f/k_B T) and, for trap energy levels \varepsilon_t>\varepsilon_f, from S_{\chi}\propto \phi=exp(\varepsilon_f/k_B T) we also have B(\varepsilon_f/k_B T)\propto exp(\varepsilon_f/k_B T), so that from the above equation we obtain,[6]

S_f=\frac{\alpha I^2}{N_qf^{\gamma}} ,

where N_q is the total number of the carriers and \alpha 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 \omega 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 S_R 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 F(r,t)=-\nabla\Phi 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), "1/f^{\gamma} 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 1/f^\gamma noise", Microelectronics Reliability 40: 1775-1780.

See also

This article is issued from Wikipedia - version of the Monday, November 16, 2015. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.