Physics:Electrokinematics theorem

From HandWiki
Revision as of 06:53, 18 November 2022 by LinXED (talk | contribs) (change)
(diff) ← Older revision | Latest revision (diff) | Newer revision → (diff)

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

[math]\displaystyle{ -\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 }[/math] ,

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

[math]\displaystyle{ -\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 }[/math] .

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

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

[math]\displaystyle{ -\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 }[/math] ,
[math]\displaystyle{ -\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 }[/math],

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

Current and capacitance

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

[math]\displaystyle{ i_{h}\equiv-\int_{S_{h}}J\cdot dS=i_{qh}+i_{Vh} }[/math] ,

through an hth electrode of interest corresponding to the surface [math]\displaystyle{ S_{h} }[/math], [math]\displaystyle{ i_{qh} }[/math] and [math]\displaystyle{ i_{Vh} }[/math] being the current due to the carriers and to the electrode voltages, to be computed through third and fourth equations, respectively.

Open devices

Consider as a first example, the case of a surface S that is not completely covered by electrodes, i.e., [math]\displaystyle{ S_{R} \ne 0 }[/math], and let us choose Dirichlet boundary conditions [math]\displaystyle{ \Phi=\Phi_{h}=1 }[/math] on the hth electrode of interest and of [math]\displaystyle{ \Phi_{h}=0 }[/math] on the other electrodes so that, from the above equations we have

[math]\displaystyle{ 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= }[/math]
[math]\displaystyle{ i_{dh}=\sum_{k=1}^{n}C_{hk}\frac{dV_{k}}{dt} }[/math] ,

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

[math]\displaystyle{ 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) }[/math] .

[math]\displaystyle{ V_{h} }[/math] 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 [math]\displaystyle{ \Phi_h }[/math] and for any other choice of boundary conditions on [math]\displaystyle{ S_R }[/math].

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

[math]\displaystyle{ 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 }[/math] ,
[math]\displaystyle{ C_{hk}=-(\int_{S_{k}}\varepsilon F_{h}\cdot dS+\int_{S_{R}}\varepsilon\Psi_{k}F_{h}\cdot dS) }[/math] .

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

[math]\displaystyle{ 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 }[/math] ,
[math]\displaystyle{ C_{hk}=-(\int_{S_{k}}\varepsilon F_{h}\cdot dS+\int_{S_{R}}\varepsilon\nabla\Psi_{h}\cdot dS) }[/math] .

In particular, this case is useful when the device is a right parallelepiped, being [math]\displaystyle{ S_R }[/math] and [math]\displaystyle{ S_E }[/math] the lateral surface and the bases, respectively.

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

[math]\displaystyle{ \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 }[/math] ,

which recover the Kirchhoff law with the inclusion the displacement current across the surface [math]\displaystyle{ S_R }[/math] that is not covered by electrodes.

Enclosed devices

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

[math]\displaystyle{ 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} }[/math] ,

with

[math]\displaystyle{ C_{hk}=-\int_{S_k}\varepsilon F_{h}\cdot dS }[/math] ,

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

[math]\displaystyle{ -\int_{S}\Phi J\cdot dS=\int_{\Omega}J\cdot Fd^{3}r }[/math] ,

from which we have

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

where [math]\displaystyle{ W }[/math] corresponds to the power entering the device [math]\displaystyle{ \Omega }[/math] across the electrodes (enclosing it). On the other side

[math]\displaystyle{ \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} }[/math] ,

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

Fluctuations

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

[math]\displaystyle{ 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 }[/math]

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

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

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

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

and the relationship

[math]\displaystyle{ 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) }[/math]

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

[math]\displaystyle{ D(\omega_l)\equiv \frac{q}{T^'}(\Delta N^+-\Delta N^-) }[/math] ,

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

[math]\displaystyle{ S_S=2q(I^++I^-) }[/math] ,

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

[math]\displaystyle{ S_S=4k_BTg }[/math] ,

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 [math]\displaystyle{ I^+\gg I^- }[/math])

[math]\displaystyle{ S_S=F_a(2qI) }[/math] ,

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

[math]\displaystyle{ \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) }[/math] ,

where m is the carrier effective mass and [math]\displaystyle{ \tau_c\ll\tau_{jmin} }[/math]. As [math]\displaystyle{ \mu=q\tau_c/[m(1+j\omega)] }[/math] and [math]\displaystyle{ G=q\mu\overline{N}/L^2 }[/math] are the carrier mobility and the conductance of the device, from the above equation and the Wiener–Khinchin theorem[11][12] we recover the result

[math]\displaystyle{ S_T=4k_BT\Re\{G(j\omega)\} }[/math] ,

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

[math]\displaystyle{ i=\frac{1}{L}\int_{\Omega}q\mu n_qEd^3r }[/math] ,

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

[math]\displaystyle{ \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) }[/math] ,

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

[math]\displaystyle{ \frac{\Delta i}{I}=\frac{1}{\Omega n_q}\int_{\Omega}\Delta n_q d^3r }[/math] ,

from which we obtain

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

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

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

Flicker noise

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

[math]\displaystyle{ S_f=\frac{n_tB}{\Omega n_q^2}\frac{I^2}{f^\gamma} }[/math] ,

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

[math]\displaystyle{ S_f=\frac{\alpha I^2}{N_qf^{\gamma}} }[/math] ,

where [math]\displaystyle{ N_q }[/math] is the total number of the carriers and [math]\displaystyle{ \alpha }[/math] 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 [math]\displaystyle{ \omega }[/math] 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 [math]\displaystyle{ S_R }[/math] 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 [math]\displaystyle{ F(r,t)=-\nabla\Phi }[/math] 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

  • Bruno Pellegrini has been the first Electronic Engineering graduate at the University of Pisa, where he is currently Professor Emeritus. He is also author of the cut-insertion theorem, that is at the basis of a novel feedback theory for linear circuits.

References

  1. Pellegrini, B. (1986-10-15). "Electric charge motion, induced current, energy balance, and noise". Physical Review B (American Physical Society (APS)) 34 (8): 5921–5924. doi:10.1103/physrevb.34.5921. ISSN 0163-1829. PMID 9940440. Bibcode1986PhRvB..34.5921P. 
  2. 2.0 2.1 2.2 2.3 Pellegrini, B. (1993). "Extension of the electrokinematics theorem to the electromagnetic field and quantum mechanics". Il Nuovo Cimento D (Springer Science and Business Media LLC) 15 (6): 855–879. doi:10.1007/bf02482462. ISSN 0392-6737. Bibcode1993NCimD..15..855P. 
  3. 3.0 3.1 Pellegrini, B. (1993). "Elementary applications of quantum-electrokinematics theorem". Il Nuovo Cimento D (Springer Science and Business Media LLC) 15 (6): 881–896. doi:10.1007/bf02482463. ISSN 0392-6737. Bibcode1993NCimD..15..881P. 
  4. Ramo, S. (1939). "Currents Induced by Electron Motion". Proceedings of the IRE (Institute of Electrical and Electronics Engineers (IEEE)) 27 (9): 584–585. doi:10.1109/jrproc.1939.228757. ISSN 0096-8390. 
  5. Shockley, W. (1938). "Currents to Conductors Induced by a Moving Point Charge". Journal of Applied Physics (AIP Publishing) 9 (10): 635–636. doi:10.1063/1.1710367. ISSN 0021-8979. Bibcode1938JAP.....9..635S. 
  6. 6.0 6.1 Pellegrini, Bruno (2013). "[math]\displaystyle{ 1/f }[/math] noise in graphene". The European Physical Journal B (Springer Science and Business Media LLC) 86 (9): 373-385. doi:10.1140/epjb/e2013-40571-7. ISSN 1434-6028. 
  7. 7.0 7.1 Schottky, W. (1918). "Über spontane Stromschwankungen in verschiedenen Elektrizitätsleitern" (in de). Annalen der Physik (Wiley) 362 (23): 541–567. doi:10.1002/andp.19183622304. ISSN 0003-3804. Bibcode1918AnP...362..541S. https://zenodo.org/record/1424341. 
  8. 8.0 8.1 Nyquist, H. (1928-07-01). "Thermal Agitation of Electric Charge in Conductors". Physical Review (American Physical Society (APS)) 32 (1): 110–113. doi:10.1103/physrev.32.110. ISSN 0031-899X. Bibcode1928PhRv...32..110N. 
  9. Maione, I. A.; Pellegrini, B.; Fiori, G.; Macucci, M.; Guidi, L.; Basso, G. (2011-04-15). "Shot noise suppression in p−n junctions due to carrier generation-recombination". Physical Review B (American Physical Society (APS)) 83 (15): 155309–155317. doi:10.1103/physrevb.83.155309. ISSN 1098-0121. Bibcode2011PhRvB..83o5309M. 
  10. Iannaccone, G.; Lombardi, G.; Macucci, M.; Pellegrini, B. (1998-02-02). "Enhanced Shot Noise in Resonant Tunneling: Theory and Experiment". Physical Review Letters (American Physical Society (APS)) 80 (5): 1054–1057. doi:10.1103/physrevlett.80.1054. ISSN 0031-9007. Bibcode1998PhRvL..80.1054I. 
  11. 11.0 11.1 Wiener, Norbert (1930). "Generalized harmonic analysis". Acta Mathematica (International Press of Boston) 55: 117–258. doi:10.1007/bf02546511. ISSN 0001-5962. 
  12. Khintchine, A. (1934). "Korrelationstheorie der stationären stochastischen Prozesse" (in de). Mathematische Annalen (Springer Science and Business Media LLC) 109 (1): 604–615. doi:10.1007/bf01449156. ISSN 0025-5831. 
  13. Machlup, Stefan (1954). "Noise in Semiconductors: Spectrum of a Two‐Parameter Random Signal". Journal of Applied Physics (AIP Publishing) 25 (3): 341–343. doi:10.1063/1.1721637. ISSN 0021-8979. Bibcode1954JAP....25..341M. 
  14. Pellegrini, Bruno (2000). "A general model of [math]\displaystyle{ 1/f^\gamma }[/math] noise". Microelectronics Reliability (Elsevier BV) 40 (11): 1775–1780. doi:10.1016/s0026-2714(00)00061-5. ISSN 0026-2714. 

See also