Open access peer-reviewed chapter - ONLINE FIRST

New Analytical Potential, Charge, and Current Distributions in JL MOSFETs: Differences between Accurate and Simplified Models

Written By

Pedro Pereyra

Submitted: 28 February 2024 Reviewed: 29 February 2024 Published: 09 July 2024

DOI: 10.5772/intechopen.1005356

MOSFET - Developments and Trends IntechOpen
MOSFET - Developments and Trends Edited by Yuxiang Tu

From the Edited Volume

MOSFET - Developments and Trends [Working Title]

Dr. Yuxiang Tu, Dr. Raed Abd-Alhameed and Dr. Ashwain Rayit

Chapter metrics overview

27 Chapter Downloads

View Full Metrics

Abstract

The operation of metal-oxide-semiconductor-field-effect transistors (MOSFETs) cannot be conceived without the prior formation of inversion layers. Charge, potential, and current distributions in this layer are important quantities for all types of MOSFET configurations. The mathematical difficulty to solve analytically the nonlinear Poisson equation in Kingston-Neustadter (KN) model, where the charge density includes electrons and holes, led to Hauser and Littlejohn (HL) to consider a simplified but solvable model, of only electrons or only holes, and to numerous HL-like compact analytical models. Recently, a new and simple method that overcomes the mathematical difficulty to solve the nonlinear Poisson equation in the inversion layer of a MOS has been introduced, and the more accurate Kingston-Neustadter model was successfully solved. We summarize here the new method and given the analytical solutions for the KN and HL models, we compare the potential and charge distribution predictions and show that they may differ by orders of magnitude. We also briefly outline the analytical results for the inversion layer width, the effective ionized atoms concentration, and the drift-diffusion currents in single and double gate junctionless (JL) MOSFETs. Specific calculations of these quantities, as functions of impurity concentration, gate potential, and oxide layer width, will be shown and compared with the predictions of the most exemplary simplified model, the Hauser-Littlejohn model.

Keywords

  • potential distribution in the inversion layer
  • the inversion-layer width
  • the electric potential in a MOS
  • effective impurity ionization
  • single and double gate JL MOSFET currents

1. Introduction

The calculation of the electric potential in the inversion layer of a MOS requires the solution of the Poisson equation, whose complexity depends on how the charge density is modeled. The charge density of the inversion layer emerges as soon as the gate potential VG becomes larger than the threshold potential VGt and the conduction band edge crosses the Fermi level. Among the first attempts to calculate the electric potential in the inversion layer of a MOS, stand out the Shockley theoretical work on semiconductors physics [1] and the Kingston and Neustadter [2] approach of 1955, where a charge density of electrons and holes is assumed, in general. Kingston and Neustadter were able to analytically perform the first integration of the nonlinear Poisson equation, but not the second integration. Since then, the analytical efforts to obtain the potential distribution for this accurate model were virtually abandoned and alternative methods, related to the calculation of MOS charges, capacitance, and currents, were devised to account for the analysis of the physical and phenomenological properties of the MOS and the field effect transistors. [3, 4, 5, 6, 7] Some years later, in 1968, Hauser and Littlejohn [8] considered a simplified but solvable model in which the charge density contains only electrons or only holes. This is also the characteristic of the numerous analytical compact models introduced since then [9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23]. The literature on the operation of the MOS is overwhelming. A good reference for the standard approaches is Tisividis book [24].

Recently, the original problem of analytically solving the nonlinear Poisson equation in the inversion layer was successfully addressed [25]. In Section 2, we outline the inversion layer problem and the Kingston-Neustadter and Hauser-Littlejohn models. In Section 3, we introduce the bandstructure, relevant parameters, and basic relations to characterize a polarized metal oxide semiconductor. In Section 4, we present a summary of the new method that, based on displacement currents and fields arguments, made possible to replace the nonlinear Poisson equation of the inversion layer by a solvable second-order nonlinear differential equation, the solution of which solves also the original Poisson equation. In Section 5, we present the analytical formulas for the width of the inversion layer and the effective concentration of ionized impurities. In Section 6, we present the potential distributions for the KN and HL models, in a unified representation [25, 26]. We show that although these potentials are qualitatively similar, the predicted inversion layer widths are different. The same occurs with the charge distributions for the KN and HL models, studied in Section 7, as a function of the impurity concentration and the gate potential [26]. In Section 8, we present the analytical formula for the drift-diffusion current for a single-gate JL MOSFET and, in Section 9, the drift-diffusion current for a double-gate JL MOSFET [26].

Advertisement

2. The Kingston-Neustadter and the Hauser-Littlejohn models

In the original semiconductor theories of W. Shockley [1] it was assumed that in the most general case, there were electrons, holes as well as ionized donors and acceptors, and the charge density could be written as [1, 27]

ρyz=qpyzn(yz)NA+ND+,E1

where q is the electric charge, NA and ND+ are the ionized acceptor and donor concentrations, and

nyz=npeyz/kBT,andpyz=ppeyz/kBT.E2

where kB is the Boltzmann constant, T is the temperature, ϕyz is the local electric potential distribution of electrons and holes, and np and pp are the minority and majority charge concentrations, which in terms of the intrinsic concentration ni and of the Fermi and thermal potentials, ϕF and ϕT=kBT/e, can be written as

np=nieϕF/ϕT,andpp=nieϕF/ϕTE3

At high temperatures NA and ND+ approximately equal to NA and ND, respectively. In a metal-oxide-semiconductor, the Kingston-Neustadter model assumption that the semiconductor has only one type of impurity, say type p, is accurate and there is no loss in generality when it is assumed that the charge of the ionized impurity concentration is NA, which because of charge neutrality, deep in the bulk is NA=ppnp. Thus, in the KN model, the density can be written as [24]

ρyz=eppeϕyz/ϕTnpeϕyz/ϕTpp+np.E4

Using the mass action law np=ni2/pp, the relation ni/pp=eϕF/ϕT and the fact that ppNA, the charge density of electrons and holes for a type p semiconductor is written as [24]

ρz=eNAeϕyz/ϕT1e2ϕF/ϕTeϕyz/ϕT1.E5

Before a source-drain potential is applied, the 1D approximation is good enough and the density, in the KN model, can be written as

ρz=eNAeϕz/ϕT1e2ϕF/ϕTeϕz/ϕT1.E6

A common procedure to perform the first integration of the Poisson equation

d2ϕdz2=ρzεs,E7

is to use the identity

ddzdz2=2dzd2ϕdz2,E8

and transform the Poisson equation into

dzddz=ρzεs.E9

A first integration of this equation can easily be performed. Kingston and Neustadter obtained the electric field in the inversion layer as

12zdz2=2eNAεseϕF/ϕTϕTcoshϕzϕFϕT+ϕzsinhϕFϕTϕTcoshϕFϕTGzϕz.E10

Here eϕF is the intrinsic Fermi energy. This first-order nonlinear differential equation could not be integrated once more. Since then, alternative, approximate, and numerical approaches have been introduced. Our purpose here is to show, in the next sections, that this problem has been overcome.

Some years later, in 1968, Hauser and Littlejohn considered a simplified but solvable charge density model where the difference p-n is replaced by only p or n, and the impurity concentration is NA or ND, depending on whether the space-charge region is p-type or n-type. If it is type p, the charge density in the inversion layer is written as

ρHz=eNAeϕHz/ϕT,E11

and the first integration, using also the identity (8), gives the electric field

12dϕHzdz2=eNAϕTεseϕHz/ϕT1+12Eb2,E12

with Eb the electric field at z=zb, the boundary between the inversion and depletion layer, where the HL model assumes that the potential vanishes, that is, ϕHzb=0. This first-order differential equation can be integrated and the electrostatic potential becomes [8].

ϕHz=ϕTlnbasech2zzbb2ϕT+tanh1ab+1,E13

with

a=2eN0ϕTεsandb=Eb2a.E14

Notice that ϕHzb=2qN0VGu/εsEb. Here VGu=VGt+Eoxttox, with VGt the threshold gate potential, Eoxt the electric field in the oxide layer at threshold, and tox the oxide layer width, see Figure 1. In the following, we will refer to these models as the KN and HL models and one of our purposes is to compare their predictions. It is worth mentioning that besides these analytical attempts, specific analytical solutions for double-gate MOSFETs, using basically the HL model, were obtained by, among others, Y. Taur and by A. Ortiz-Conde et al. We will not address these works nor the numerous analytical compact models, nor the numerical approaches and simulation codes.

Figure 1.

A band structure of a metal-oxide-semiconductor with a type-p semiconductor and a positive gate potential VG larger than the threshold potential. Several parameters that characterize the system are shown.

In the last 20 years, new geometries and multi-gate MOSFETs were introduced. The inversion layer phenomenon that occurs in the oxide-semiconductor interface of MOS in a single-gate (SG) MOSFET occurs also in each oxide-semiconductor interface of multi-gate MOSFETs. Hence, studying the inversion layer in a single MOS structure is essential. After obtaining the potential and charge distributions, as explicit functions of the relevant parameters, we will consider the analytical expression for the current distribution in a single gate JL MOSFET, and based on this theory, we will discuss the calculation of the analytical formulas for the potential distribution and current in a double gate JL MOSFET, as a simple extension of the SG MOSFET theory.

Advertisement

3. Parameters and basic relations in biased MOS structures

To fix the boundary conditions, the notation and some of the essential parameters used to characterize the MOS structure, we recall here some well-established relations and definitions. Most of the of these relations can be found in textbooks, and some good references are [1, 6, 24, 27, 28, 29, 30, 31]. We will assume that the potentials are measured from the flat-band condition and that the electric potential at zp (and beyond this point, i.e., in the bulk (see Figure 1)), is ϕzp=0. For gate potentials larger than the threshold potential, we will distinguish three characteristic layers between the z=tox and zp. Between tox and z=0 the oxide-layer, where the electric potential is linear and given by:

ϕoxz=VGz+toxEoxfortox<z<0,E15

and the capacitance per unit area is

Cox=εoxtox,E16

with εox the electric permittivity in the oxide layer. From z=0 to zi we have the inversion layer, and from zi to zp, the depletion layer. We will assume that the charge density of the ionized impurity atoms in the depletion layer is homogeneous and will be denoted as Nd. We will see below that NdNA, being NA the concentration of acceptor impurities. The precise concentration of the ionized impurity atoms will be determined by the continuity requirements. A closed formula will be obtained for this dynamic quantity.

The boundary conditions at the oxide-semiconductor interface, that is, at z=0, and at the edge of the depletion layer, say at zp, are

ϕox0=VGEoxtox,E17

and

ϕdzp=0,dϕdzp/dz=0,E18

respectively. Under these conditions, the electric potential distribution in the depletion layer is given by

ϕdz=eNd2εszpz2+ϕzpzi<z<zp,E19

with zp=zi+wp, and wp=2εsVGu/eNA the depletion layer width. The electric field, as a function of the electric potential, can be written as

Edzϕdz=2eNdϕdzεszi<z<zp.E20

Here and in the following, we write εs instead of εrε0.

As is well known, for the band-edge bending to reach the threshold potential, as shown in Figure 1, the gate potential has to be greater than the threshold potential VGt, defined by

VGt=EgEFs+Eoxttox/e.E21

Here, Eg is the gap energy, EFs the semiconductor Fermi energy, Eoxt the electric field in the oxide layer at threshold, and tox the oxide layer width. Once the inversion regime is established and VG is increased, the highly charged inversion layer emerges whose thickness, the inversion-layer width zi, depends on the gate potential and on the impurity concentration. In terms of the parameters defined above, the condition for determining the inversion layer width is

VGt=Eoxttox+ϕzi,E22

where

ϕzi=EcEFseVGu,E23

is the surface potential threshold denoted here as VGu, with u for umbral (from latin umbra—shadow, and liminaris—limit). After solving Poisson’s equation, we will present a closed formula for the inversion-layer width zi.

The positive charge on the metallic side creates an electric field Ez that penetrates on the semiconductor side, attracts electrons, and repels holes and, at the end, establishes an electric potential ϕz that redefines the valence and conduction band edges as Evz and Ecz, respectively. The inversion population process modifies the nature and the phenomenology of the physical system on the semiconductor side, whose precise description depends on whether the Poisson equation that governs the intertwined relation, between gate potential, charge distribution, and electric potential, can be fully or partially solved. In the next section, we summarize the main argument to transform the Poisson equation into a solvable nonlinear second-order differential equation.

Advertisement

4. Charge and field displacement in the inversion layer

When the gate potential VG becomes greater than the threshold potential VGt, the charge distribution ρi, regardless of how it is modeled, undergoes an inversion-population process (IPP). If the semiconductor is linear and the charge concentration NizϕT in the growing inversion layer is highly localized, the balance between the time rate of the thermal energy kBTNizϕT and the time rate of the displacement field energy uEϕT=tD, can be written as

ϕTρiϕtTρiϕTdρiϕT=12uEϕtTuEϕTduE.E24

where ρiϕT=eNizϕT is the charge distribution in the inversion layer, and ϕt the electric potential at threshold. Assuming that E=z/dzẑ, we have

εs2zdz2εs2Et2=ϕTρiϕρiϕt.E25

It was shown that when the charge density is, for example, the charge density of the KN model, the electric field obtained from this equation is equal to the electric field in Eq. (10). Taking into account the Poisson equation and the electric field in Eq. (10), we obtain the nonlinear second-order differential equation

ϕTd2ϕzdz2=12zdz212ξ2,E26

where ξ is a model and threshold-dependent parameter, given by

ξ2=Et2+2ϕTρϕtεs.E27

These are the main and most consequential results of the new method. In the next sections, we will apply this method for the HL and the KN models. The charge densities in these models are different; hence, the parameter ξ is also different, while the differential equations are formally similar. Eq. (26) shows an important step toward the solution of the Poisson equation in the inversion layer. Eq. (26) is solvable and, as will be seen below, its solution is also a solution of the original Poisson equation. It has been shown in Ref. [25], that the analytical solution of (26) is

ϕz=ϕ02ϕTlncoshξzz02ϕTf0ξsinhξzz02ϕT,E28

with ϕ0 the potential and f0 the electric field at z=z0. Given this solution, and before we refer to the electric potential in the HL and KN models, it is convenient to consider the inversion layer width in general.

Advertisement

5. Inversion layer width and depletion layer concentration

When the continuity conditions are imposed at zi, one obtains both the inversion layer width

zi=2ϕTξcosh1eθ/2ϕTξ2+Es4Es2ξ21eθ/ϕtEs2ξ2,E29

where θ=ϕsVGu, and the concentration of the ionized impurity atoms in the inversion layer regime, given by

Ndl=εs2ziξ2+eϕsϕzi/ϕtEs2ξ2,E30

when θ0, and by

Nds=εsEs22eϕs,E31

when ϕs<ϕzi. Replacing the corresponding parameter ξ one has the inversion layer width predicted in the corresponding model. See the next section.

Advertisement

6. Potential distributions

As mentioned above, the difference between one model and another lies in the charge density and therefore in the parameter ξ. Now we will be more specific.

6.1 Potential distribution in Hauser-Littlejohn model

In the solvable Hauser-Littlejohn model, assuming a type p semiconductor, the charge density and the parameter ξ are

ρHz=eN0eeϕHz/kBTandξH2=Eb22eN0ϕTεs.E32

For the boundary conditions considered by Hauser and Littlejohn, that is, for z0=zb, and at this point ϕ0=ϕb=0 and f0=Eb, the potential distribution becomes

ϕHz=2ϕTlncoshξHzzb2ϕT+EbξHsinhξHzzb2ϕT,E33

with Eb=2eNdVGu/εs. It is not difficult to show that the HL solution in (13), obtained by direct integration, reduces rigorously to ϕHz (see also Appendix of Ref. [25]). In this model, zb=zi defines the boundary between the inversion and depletion layer, and Hauser and Littlejohn assume, that the electric potential at this point vanishes, and we can write also the potential distribution for the HL model in terms of the surface parameters at the oxide-semiconductor interface, that is, at z0=0, where ϕ0=ϕs and f0=EHs. In this case, the potential distribution is written as

ϕLz=ϕs2ϕTlncoshξHz2ϕT+EHsξHsinhξHz2ϕT.E34

Here the surface electric field EHs is

EHsz=eNAϕTeeϕsVGu/ϕT+ξH21/2.E35

Notice that the assumption ϕHzb = 0, implies that ϕHz=ϕLzVGu. In terms of ϕLz, the charge density in the Hauser-Littlejohn model is

ρHz=eNAeeϕLzVGu/kBT.E36

In Figure 2 we plot the potential energy eϕLz as function of z, for different values of the impurity concentration, and for different surface potentials, as well. In the next subsection, by comparing with the more accurate KN model, it will be seen that the HL model predicts much smaller inversion layer widths, and grows faster near the oxide-semiconductor interface. The potential distribution ϕHz is well known and has been widely used as an approximate result for double-gate devices [15, 16].

Figure 2.

The potential energy in the HL model as a function of z for different values of the impurity concentration NA and different surface potentials ϕs. The small circles indicate the limit (at zi) between the inversion and the depletion layers. The dashed black lines define the threshold energies eVGu, in eVs.

6.2 Potential distribution in the Kingston-Neustadter model

In the important Kingston-Neustadter model, the charge density can be written as

ρKz=2eNAeϕF/ϕTsinhϕFϕKzϕTsinhϕFϕT.E37

The parameter ξ is

ξK2=8eNAεseϕF/ϕTϕFϕTsinhϕFϕT,E38

and the potential distribution, for ϕ0=ϕs, f0=Es and z0=0, becomes

ϕKz=ϕs2ϕTlncoshξKz2ϕT+EsξKsinhξKz2ϕT.E39

Given this potential distribution, an alternative representation for the electric field in the inversion layer is

EKz=ξKsinhξKz2ϕT+EsξKcoshξKz2ϕTcoshξKz2ϕT+EsξKsinhξKz2ϕT.E40

Here the surface electric field is

Es=2eNAεseϕF/ϕTϕTcoshϕsϕFϕT+ϕssinhϕFϕTϕTcoshϕFϕT1/2.E41

It was shown in Ref. [25]), by substitution and numerical evaluation, that the function ϕKz in (39) solves also the original Poisson equation and the electric field in (40), matches with the electric field 2Gzϕz in (10), that was obtained by Kingston and Neustadter after the first integration of the original Poisson equation J. Reiter has shown also [32], by numerically solving the differential equation, that the difference between the numerical and analytical solutions is less than 0.5%, in the entire parametric space.

In Figure 3 we plot the potential energy eϕKz as functions of z, for different values of the impurity concentrations NA, as well as for different values of the surface potential ϕs. In this figure, we indicate the Fermi levels (dashed black lines) and the borders between the inversion and depletion layers. These points, where the potential energies cross the Fermi levels, define the inversion layer widths zi. The behavior of ϕKz is qualitatively similar to that of ϕLz in Figure 2. It grows also rapidly for gate potentials VG larger than the threshold potential VGt, or surface potentials ψs larger than the umbral potential VGu.

Figure 3.

The potential energy in the KN model as a function of z for different values of the impurity concentration NA and different surface potentials ϕs. The small circles indicate the limit (at zi) between the inversion and the depletion layers. The dashed black lines define the threshold energies eVGu (or Fermi levels).

It is now possible to compare the electric potentials and the inversion layer widths predicted by the Kingston-Neustadter and the Hauser-Littlejohn models. In Figure 4, we plot the potential energies, predicted in both models, as functions of z, and for different values of the impurity concentration NA and different values of the surface potential ϕs. The potential distributions are qualitatively similar but quantitatively different. In the HL model, the inversion layer widths are smaller than those predicted in the KN model, by almost a factor of 2, the electric potentials are also smaller but near the oxide-semiconductor interface, the electric potentials in the HL model grow faster than in the KN model. In general, the potential and the layer widths are underestimated in the HL model. Because of the smaller inversion layer width and the rapid grow of ϕHz, the charge distribution in the HL model is much more localized and closer to the 2D assumption.

Figure 4.

The electric potential energy in the KN (continuous) and HL (dashed) model as a function of z for different values of the impurity concentration NA and different surface potentials ϕs. The small circles indicate the limit (at zi) between the inversion and the depletion layers.

In Figure 5 we plot the inversion layer widths in the KN model as functions of the gate potential VG=ϕs+κtoxEs, for different values of the impurity concentration. For these plots, the silicon parameter κ=εs/εox=3 and oxide-layer width tox=2nm were considered.

Figure 5.

The inversion-layer width zi in the KN model as function of VG, for different impurity concentrations. In the upper axes, the threshold potentials is written as VGt=VGu+Δox, with Δox=toxEoxt=κtoxEst, the potential drop in the oxide layer.

Given the explicit functions ϕL, ϕK, ξH and ξK, we will in the next section evaluate the charge density distributions.

Advertisement

7. Charge density distributions

An important consequence of the inversion population phenomenon is the rapid increase of the charge density in the inversion layer. Replacing the electric potentials ϕLz and ϕKz in the corresponding density, Eqs. (36) and (37) we have, for the HL model, the charge density

ρHz=eNHdeϕsVGu/ϕTcoshξHz2ϕT+EHsξHsinhξHz2ϕT2,E42

and, for the Kingston-Littlejohn model, the charge density

ρKz=eNKd{eϕs/ϕTcoshξHz2ϕT+EsξHsinhξHz2ϕT21e2ϕF/ϕT1eϕs/ϕTcoshξHz2ϕT+EsξHsinhξHz2ϕT2.E43

To visualize the similarities and differences between these charge densities, we plot them side by side in Figure 6, and the surface densities in Figure 7. The densities in Figure 6 are plotted as functions of the distance z to the oxide-semiconductor interface, for different values of the impurity concentration and for different values of the surface potential ϕs. The qualitative behavior is similar, with rapid growth in the inversion layer and almost constant densities in the depletion layers. In these layer, we take into account the effective impurity ionization Nd. The effect of this concentration is observed in Figure 6a where the charge density varies continuously at zi and vanishes at the border of the depletion layer at zp=zi+wp. The densities in the HL model behave as shown in Figure 6b with a practically constant concentration in the depletion layers. As observed before, the HL predicts thinner inversion layers. Near the inversion layer edge, around zi, the densities are quantitatively similar, but as one approaches the oxide interface, the differences grow. To visualize this difference we plot in Figure 7, the surface densities

Figure 6.

Charge density in the KN (left) and HL (right) models as a function of z, and for different values of NA. The small circles indicate the limit (at zi) between the inversion and the depletion layers.

Figure 7.

The charge density at z=0 in the KN (left) and the HL (right) model, as functions of the surface potential ϕs, for two values of the impurity concentration NA=1023m3 and NA=1024m3.

ρHsz=eNHdeϕsVGu/ϕT,E44

and

ρKsz=2eNKdeϕF/ϕTsinhϕsϕFϕT+sinhϕFϕTE45

From Figure 7, it is clear that the difference between the charge densities in the HL and KN models grows with ϕs by orders of magnitude. We also see that due to the exponential function in (44), the density at the surface, in the HL model, is practically insensitive to the impurity concentration, while in the KN model, the surface density in (45) is sensitive to the impurity concentration, as could be expected.

Advertisement

8. Drift current in a JL MOSFET channel

When the biased MOS system is part of a single or multiple gate MOSFETs, the charge and potential distributions obtained before play an important role in the nature and magnitude of the drift-diffusion currents moving in the MOSFET channel(s), parallel to the oxide-semiconductor interface(s). To simplify the analysis, we will have the so-called junctionless MOSFETs. As is well known and will be seen here, the charge carriers in the inversion layer play the most important role.

A well-established formula for the calculation of the source-drain current is Pao-Sah’s formula [5]

ID=eDnWxL0zpnyzdz0Ldyydy.E46

Here L is the channel length that we will assume alongside the y direction, Wx the channel depth, Dn the diffusion coefficient, and zp=zi+wp the inversion plus depletion layers widths. The semiconductor layer width Wz may be greater or smaller than zp. When the drift and diffusion currents are taken into account, the distribution nyz becomes

nyz=nAeϕzηyϕFs/ϕT,E47

where ϕz is the electrical potential due to the gate potential, ηy is the quasi-Fermi level modified by the drain-source potential UD along the channel, and ϕFs=EcEFs/e is the equilibrium Fermi potential in the bulk, that we will assume to be measured from the conduction band edge. Usually dz is replaced by /Ezϕ, where Ezϕ=2Gzϕ. We do not need to make this change because we do have the electric potential as an explicit function of Z. We keep the integration upon z, but we will change the integration upon y by the integration upon η. We also split the integration in z in two parts, one for the inversion region and one for the depletion layer. In this case,

ID=eDnwxniL0zieϕiz/ϕTdz+zizi+wpeϕdz/ϕTdz0UDeηyϕFs/ϕT,E48

with the function ϕi given by the electric potential defined in (39), and the function ϕd given above for the electric potentials in the depletion layer, Eqs. (30) and (31). Similarly, the current in the HL model is straightforwardly calculated when the electric potential defined in (34) is replaced for ϕi. In this section we will present results only for the KN model and we will drop the subscript K. The integrations are direct and we obtain, in the KN model, the current

ID=eDnwxniLeϕsϕFs/ϕT2ϕTsinhηz2ϕTηcoshηz2ϕT+Esηsinhηz2ϕT0zi+πεsϕTeNdeϕF/2ϕTErfeNdwp2εsϕT1eUD/ϕT,E49

which reduces to

ID=eDnwxniL2ϕTξeϕsϕFs/ϕTeVGuϕs/2ϕTsinhξzi2ϕT+πεsϕTeNdeϕF/2ϕTErfeNdwp2εsϕT1eUD/ϕT.E50

Since VGu=ϕFs, this becomes

ID=eDnwxniL2ϕTξeϕsVGu/2ϕTsinhξzi2ϕT+πεsϕTeNdeϕF/2ϕTErfeNdwp2εsϕT1eUD/ϕTIDi+IDd.E51

As mentioned before, the concentration Nd should be replaced by Nds, when ϕs<VGu, and by Ndl when ϕs>VGu. We shall present now some graphs to visualize the behavior of these currents when both the gate and source-drain potentials are applied. For all graphs shown in this section, we will assume the mobility μ=400cm2/Vs in the depletion and inversion layers. These quantities can of course be changed. The impurity concentrations NA will be indicated in the graphs.

We will plot now the depletion and inversion layer currents and the total current, separately, as functions of the source-drain potential UD and of the surface potential ϕs. Since VG=ϕs+κtoxEs, plotting as a function of ϕs is almost as plotting as function of VG, not exactly as will be seen below. In all cases, of this section, we will assume that the mobility μ=400cm2/Vm is the same in the depletion and inversion layers, L=0.4μm and Wx=1.0μm. In Figure 8 we plot the depletion, inversion, and total currents as functions of UD and of ϕs. In (a) the depletion current below and above the threshold, in (b) the inversión layer current and in (c) the sum of both currents. Above the threshold, the current IDd grows with UD but is independent of VG. Both currents IDd and IDi grow monotonously with UD and tend to a saturation value at higher values. For this graph, we assumed also that Wx=1μm and L=Wx/2.5.

Figure 8.

Depletion, inversion, and total current as functions of the source-drain potential UD and of the surface potential ϕs. In (a) the depletion current IDd below and above the threshold, in (b) the inversión layer current IDi and in (c) the total current ID=IDd+IDi. The depletion current IDd as a function of UD grow monotonously toward a saturation value, as a function of ϕs the depletion current grows steadily until the gate potential reaches the threshold value, where IDd experiences a sharp change. Above this threshold, the depletion current becomes independent of the gate potential. At variance with IDd, the inversión layer current IDi, as a function of the ϕs (i.e. of the gate potential), starts growing at the threshold potential, and for a fixed value of the gate potential, the inversion layer current also grows with UD and tends to a saturation value at higher values. For this graph μ=400cm2/Vs,Wx=1μm,L=Wx/2.5.

Due to the nonlinear relationship between the gate and the surface potential and the nonlinear relationship between the surface electric field Es and ϕs, the growths of the gate current inversion layer as a function of ϕs and as a function of VG should look different. In fact, this difference can be seen in Figure 9 (a) and (b), where the same inversion current is plotted, in one case as a function of ϕs, and as a function of VG in the other case. In Figure 9 (a) the current IDi grows exponentially as a function of ϕs while the currents IDi in Figure 9 (b) grow almost linearly as function of VG. Since the depletion current remains almost constant above the threshold behavior the linear behavior of IDi occurs also in the total current ID.

Figure 9.

The inversion-layer current IDi as function of the surface potential ϕs (panel (a)) and the gate potential VG (panel (b)). In (a) the current is plotted as a function of (equidistant values) ofϕs, and in (b), the same current but as a function of (equidistant values) of VG. In the last case, it is evident that at larger values of VG the current becomes a linear function of VG.

In Figure 10, we plot the total drift-diffusion current ID as a function of the gate potential VG, for different values of the impurity concentration. As mentioned before the inversion layer current and the total current behave, at larger values of the gate potential, as linear functions of this variable, with lower slopes at higher impurity concentrations. It can be seen also in Figure 10, that the threshold potential for higher impurity concentration is larger. These threshold potentials and the so-called operational threshold potentials, determined at the intersection of the linear current with the abscissa, are indicated with arrows. The operational threshold potentials are larger than the gate potential thresholds. In the particular case considered here, the differences are of the order of 0.3 V. Given the device parameters, the operational potentials can also be predicted. To complete the comparison of the KN and HL model predictions, we plot in Figure 11 the drift-diffusion currents in the KN and HL models assuming the same device parameters. The predict currents are also different. The currents in the simplified HL model are smaller than the currents in the most accurate KN model. For these graphs, we assumed that the impurity concentration is NA=1023m3.

Figure 10.

The total source-drain current ID=IDd+IDi in silicon MOSFETs as functions of the gate potential, for three different values of the impurity concentration. Here, we assume that the mobility, in the depletion and inversion layers, is μ=400cm2/Vm, UD=0.001V, and channel length L=0.5μm. The gate potential thresholds VGt are given on the graph. These thresholds and the so-called operational threshold potentials, defined generally at the intersection of the abscissa with the extension of the linear current (see dashed lines), are indicated with arrows.

Figure 11.

The drift-diffusion currents in the KN and HL models as functions of the gate potential. The HL model predicts smaller currents than the KN model.

The theory presented in this section, and before, can be applied almost straightforwardly to more complex devices, among them, to the double-gate (DG) MOSFET studied in the next section, and to multi-gate MOSFETs.

Advertisement

9. The double gate JL MOSFET

Explicit calculations for other MOSFET configurations can be performed almost as a trivial extension of the results presented in the last sections. We will refer only and briefly to the double-gate MOSFET. Consider a double-gate device like the one sketched in Figure 12, where the oxide layers are at a distance tSi and the front and back gate voltages VGf and VGb are assumed to have opposite polarizations. In this case, holes accumulate near the back oxide-semiconductor interface and electrons near the front oxide-semiconductor. Before the source-drain potential UD is applied, we can write the electric potential in the inversion layers in the 1D approximation as

Figure 12.

A double gate structure as a generalization of the single gate MOS of Figure 1. We assume here that the back gate potential is negative and attracts holes. The potential distribution ϕDGz (red curve) is schematically shown.

ϕiDGz=ϕifz+ϕibz.E52

Assuming that the charge densities are described by the Kingston-Neustadter-like distributions, the potential distributions in the front and back inversion layers, respectively, can be written as

ϕifz=ϕsf2ϕTlncoshξKz2ϕT+EKfsξKsinhξKz2ϕT,E53

and

ϕibz=ϕsb+2ϕTlncoshξKz2ϕT+EKbsξKsinhξKz2ϕT.E54

Here, z=tSiz, ϕsf and ϕsb are the front and back surface potentials, and EKfs and EKbs are the front and back surface electric fields. In the symmetric case with toxf=tosb and VGf=VGb, the potential function ϕDGz is asymmetric, as sketched in Figure 12 (red curve), and the potential energy qϕiDGz is symmetric, as shown in Figure 13 (a), where a silicon layer tSi40nm is assumed, which for an impurity concentration NA=1×1018cm3 is 1/2wpd+wnd+zif+zib, where wpd and wnd are the depletion layers widths and zif and zib the inversion layer widths.

Figure 13.

Electric potential of drift-diffusion current in a symmetric double-gate JL MOSFET. In (a) the electric potential for a DG MOS where the silicon thickness tSi is smaller than the expected depletion plus inversion layers width wp+wn+zif+zib. Some parameters are assumed equal to those in Ref. [23], i.e., L=11μm, Wx=10μm. However, nowhere in Ref. [23] are the impurity concentration and mobilities given. In this reference, the silicon layer width tSi is assumed equal to 8 nm, slightly larger than the sum of the inversion layer widths for NA=1×1018cm3. Here, we assume that tSi40nm.

Given the potential distribution ϕiDGz, we can, using Pao-Sah’s [5] formula, obtain the analytical expression for the drift-diffusion current in the DG JL MOSFET channel as a function of the relevant parameters. In the symmetric case, the total current in the DG JL MOSFET channels can be written as

IDG=eDnWxniL0zifeϕifz/ϕTdz+ziftSi/2eϕpdz/ϕTdz+tSi/2tSizibeϕndz/ϕTdz+tSizibtSieϕibz/ϕTdz×0UDeηyϕFs/ϕT.E55

All integrals in this equation can be performed analytically, and for zif+zib<tSi<wp+wn+zif+zib, and zif=zib, the current becomes

IDG=eDnwxniL1eUD/ϕT4ϕTξeϕsVGu/2ϕTsinhξzi2ϕT+πεsϕTeNdeϕF/2ϕTErfeNdwp2εsϕT+ErfeNdtSiwp2zif2εsϕT=IDGi+IDGd.E56

It can easily be seen that the current in the DG JL MOSFET is practically twice the charge and current in the single gate MOSFET channel. [25] In fact, it is exactly twice when tSi=2wp+2zif.

In Figure 13(b) we plot the drift-diffusion current of Eq. (56) as a function of the gate potential, for three different values of the source-drain voltage UD. For this graph we assume that the impurity concentration is NA=1×1018cm3, the channel length is L=11μm, the gate width is Wx=10μm, and the mobility is assumed equal to 40cm2/Vs. Some of these parameters are the same as those used in Figure 7 of Ref. [23], where a simplified charge distribution is assumed; however, we cannot compare with that reference because some parameters, such as the impurities concentration and mobilities, are not given.

It is clear that the analytical expression for the current in Eq. (51) can also be used for asymmetric DG JL MOSFETs. The general formulas for the inversion layer width, surface electric field, and effective concentration of ionized impurity atoms, introduced for the single gate MOS in Section 5, are also valid with the appropriate changes for the double gate MOSFET. The ability to evaluate these quantities, particularly the widths of the inversion layers, is, for a given impurity concentration, an important input in the design of MOSFETs and in the definition of some device parameters, such as the silicon layer width tSi.

Advertisement

10. Conclusions

We presented here an outline of the new theoretical approach for the analytical calculation of potential distributions and drift-diffusion currents in MOS and JL MOSFETs. We presented a summary of the novel method that makes it possible to solve analytically the nonlinear Poisson equation in the inversion layer of MOSs and MOSFETs. We apply this method to the more precise Kingston-Neustadter charge density model, unsolved until now, where the presence of electrons and holes in the inversion layer is assumed, and also to the most representative of the simplified compact analytical models, the Hauser-Littlejohn model, which assumes that only electrons or only holes are present. The predictions of both models for the potential and charge distributions, as well as for the drift-diffusion current, were compared and we have shown that they generally differ. In the HL model, the widths of the inversion layer are much smaller, and the charge distributions are almost 2D, with a higher surface density and practically insensitive to impurity concentration. The drift-diffusion currents predicted by the HL model are also different and smaller than the currents predicted by the KN model. Independent of these comparisons with the standard approaches, we have presented here several analytical formulas for potential and charge distributions and new analytical formulas for the drift-diffusion current in single and double-gate MOSFETs.

Acknowledgments

I acknowledge valuable comments by Herbert P. Simanjuntak, Jürgen Reiter, and A. Robledo-Martinez.

Conflict of interest

The author declares no conflict of interest.

References

  1. 1. Shockley W. The theory of p-n junctions in semiconductors and p-n junction transistors. Bell System Technical Journal. 1949;28:435
  2. 2. Kingston RH, Neustadter SF. Calculation of the space charge, electric field, and free carrier concentration at the surface of a semiconductor. Journal of Applied Physics. 1955;26:718-720
  3. 3. Kahng D, Atalla MM. Silicon-silicon dioxide field induced devices. Solid State Device Research Conference. Pittsburgh. US3204160A. 1960
  4. 4. Sah CT. Characteristics of the metal-oxide-semiconductor transistors. IEEE Transactions on Electron Devices. 1964;ED-11:324-345
  5. 5. Pao HC, Sah CT. Effects of diffusion current on characteristics of metal-oxide (insulator)-semiconductor transistors. Solid State Electonics. 1966;9:927-937
  6. 6. Grove AS. Physics and Technology of Semiconductor Devices. New York: John Wiley; 1967. ISBN: 978-0-471-32998-5
  7. 7. Bassett RK. MOS technology, 1963-1974: A dozen crucial years. The Electrochemical Society Interface. Fall 2007. 1970:46-50
  8. 8. Hauser JR, Littlejohn MA. Approximations for accumulation and inversion space-charge layers in semiconductors. Solid State Electonics. 1968;11:667-674
  9. 9. Demoulin E, van De Wiele F. Inversion layer at the interface of Schottky diodes. Solid State Electonics. 1974;17:825-833
  10. 10. Temple VAK, ShewChun J. The exact low-frequency capacitance-voltage characteristics of metal-oxide-semiconductor (MOS) and semiconductor-insulator-semiconductor (SIS) structures. IEEE Transactions on Electron Devices. 1974;18:235-242
  11. 11. Brews JR. A charge-sheet model of the MOSFET. Solid State Electonics. 1978;21:345-355
  12. 12. Flandre D, van de Wiele F. A new analytical model for the two-terminal MOS-capacitor on SO1 substrate. IEEE Electron Device Letters. 1988;9:296-299. DOI: 10.1109/55.722
  13. 13. Schubert M, Höfflinger B, Zingg RP. An analytical model for strongly inverted and accumulated silicon films. Solid State Electonics. 1990;33:1553-1567
  14. 14. Taur Y. An analytical solution to a double-gate MOSFET with undoped body. IEEE Electron Device Letters. 2000;21:245-247
  15. 15. Taur Y. Analytic solutions of charge and capacitance in symmetric and asymmetric double-gate MOSFETs. IEEE Transactions on Electron Devices. 2001;48:2861-2869
  16. 16. Ortiz-Conde A, García FJ, Muci SJ. Rigorous analytical solution for the drain current undoped symmetric dual-gate MOSFETs. Solid State Electonics. 2005;49:640-647
  17. 17. Taur Y, Liang X, Wang W, Lu H. A continuous, analytic drain-current model for DG MOSFETs. IEEE Transactions on Electron Devices. 2004;25:07
  18. 18. Roy AS, Sallese JM, Enz CC. A closed-form charge-based expression for drain current in symmetric and asymmetric double gate MOSFET. Solid State Electronics. 2006;50:687
  19. 19. Moreno-Pérez E et al. An inversion-charge analytical model for square gate-all-around MOSFETs. IEEE Transactions on Electron Devices. 2011;58:2854
  20. 20. Hu M-C, Jang S-L. An analytical fully-depleted SOI MOSFET model considering the effects of self-heating and source/drain resistance. IEEE Transactions on Electron Devices. 1998;45:797
  21. 21. Reyboz M, Rozeau O, Poiroux T, Martin P, Jomaah J. An explicit analytical charge-based model of undoped independent double gate MOSFET. Solid State Electronics. 2006;50:1276
  22. 22. Oproglidis TA et al. Analytical drain current compact model in the depletion operation region of Short-Channel triple-gate Junctionless transistors. IEEE Transactions on Electron Devices. 2017;64:66
  23. 23. Khandelwal S et al. BSIM-IMG: A compact model for ultrathin-body SOI MOSFETs with Back-gate control. IEEE Transactions on Electron Devices. 2019;59:2012
  24. 24. Tsividis Y. Operation and modelling of MOS transistor. In: The Oxford Series in Electrical and Computer Engineering. 2nd ed. Oxford, UK: Oxford University Press; 2003. See Appendix F, and references therein
  25. 25. Pereyra P. Submitted to Semiconductor Science and Technology
  26. 26. Pereyra P. DOI: 10.22541/au.170864077.78820151/v1
  27. 27. Sze SM. Physics of Semiconductor Devices. Singapore: Wiley Interscience; 1981
  28. 28. Shockley W. Electrons and Holes in Semiconductors. New York: D. Van Nostrand; 1950
  29. 29. Shalimova KV. Física de Semiconductores. Moscou: MIR; 1975
  30. 30. Hook JR, Hall HE. Solid State Physics. Chichester West Sussex, England: J. Wiley & Sons; 1991. ISBN: 0-471-92805-4
  31. 31. Seeger K. Semiconductor Physics, an Introduction. 6th ed. Heidelberg, Germany: Springer; 1997. ISBN:3-540-61507-5
  32. 32. Reiter J. Private communication, work in progress

Written By

Pedro Pereyra

Submitted: 28 February 2024 Reviewed: 29 February 2024 Published: 09 July 2024