FIG 3 - uploaded by Juan José Sáenz
Content may be subject to copyright.
Energy band diagram for the three possible operation regimes of the MOSFET. 

Energy band diagram for the three possible operation regimes of the MOSFET. 

Source publication
Article
Full-text available
We present a compact model based on the Landauer transmission theory for the silicon quantum wire and quantum well metal-oxide-semiconductor field effect transistor (MOSFET) working in the ballistic limit. This model captures the static current-voltage characteristics in all the operation regimes, below and above threshold voltage. The model provid...

Contexts in source publication

Context 1
... H ( E ) is the Heaviside function ͓ H ( E ) 1 if E 0; H ( E ) ϭ 0 otherwise ͔ . As we will see in the next section, E FS Ϫ E v n ( x max ) depends on V GS and V DS . Equation ͑ 5 ͒ reveals three possible operation regimes depending on the lo- cation of the subbands with respect to the Fermi levels ͑ Fig. 3 ͒ . To discuss these regimes, we define E n ( x max ) as the or- dered set of energies E v n ( x max ), being E 1 ( x max ) the lowest energy in this set. a E FS E 1 ( x max ): At zero Kelvin the current is zero because the population of ϩ k x ( Ϫ k x ) states at source and drain contacts have not enough energy to transmit across the barrier towards the drain ͑ source ͒ . If the temperature is greater than zero, the tails of the Fermi functions extend above the barrier. Moreover, if the device is out of the equilibrium ( V DS Ͼ 0) some leakage current flows because the tail of f S ( E ) above the maximum of the barrier, which is transmitted towards the drain, is larger than the corresponding tail of f D ( E ) that cross the channel ballistically towards the source. This leakage current corresponds to the diffusion current. This regime takes place when V GS Ͻ V th , where V th ͑ threshold voltage ͒ is the gate voltage for the first subband to cross the Fermi level at the source; i.e., the gate voltage for which E 1 ( x max ) ϭ E FS . The situation described corresponds to the subthreshold regime. ͑ b ͒ E FS Ͼ E i ( x max ), where i у 1; and E FD Ͼ E j ( x max ), where j у 1: The forward I ϩ ͑ backward I Ϫ ) current is given by the electrons with energies between E i ( x max ) ͓ E j ( x max ) ͔ and E FS ( E FD ) ͑ at zero Kelvin ͒ . The net current is the difference I ϩ Ϫ I Ϫ , and the slope ͑ G ͒ of the output characteristic is an even multiple of the quantum conductance ( G 0 ): G ϭ dI / dV DS ϭ 2 M G 0 , where G 0 ϭ q 2 / ␲ ប ; the factor 2 arises from the two-fold degeneracy of each set of valleys and M is the number of subbands below E FD . This situation takes place when V GS Ͼ V th and V DS is low enough. It corresponds to the linear regime of the MOSFET operation. ͑ c ͒ E FS Ͼ E i ( x max ), where i у 1; and E FD р E 1 ( x max ): In this case, the forward current I ϩ is given by the electrons with energies between E i ( x max ) and E FS . The backward current I Ϫ is suppresed because the occupied Ϫ k x states at the drain contact cannot transmit across the barrier towards the source. This case corresponds to the saturation regime V GS Ͼ V th and V DS high ͒ , for which the current is independent of the applied source-drain voltage. The onset of the saturation regime occurs at the drain voltage V DS ,sat that makes E ( x ) ϭ E . In the last section we have derived the current equation of the quantum wire MOSFET in terms of the separation between the Fermi level and the bottom of the subbands at the maximum energy point ( x max ). The goal of this section is to find how this separation depends on the applied gate and drain voltages for the DG structure. We will assume that the two-dimensional short-channel effects are small ͑ future extensions of the compact model will deal with two- dimensional electrostatics ͒ . The potential and charge distribution of the intrinsic DG structure has been reported by Taur. 13,14 The electrostatic analysis was only done in the vertical direction, but it serves our purposes if we apply the results at the maximum energy point ( x max ). At this special point, the potential and charge distributions are essentially controlled by the electric field perpendicular to the silicon-insulator interface; the lateral electric field is close to zero, and the gradual channel ap- proximation applies. 15 The potential along the vertical direction (0 р z р t /2) is given by 13,14 ln cos e 0 z , 6 2 kT 2 ␧ Si kT where ␺ 0 ε ␺ ( z ϭ 0) is the potential at the center of the silicon film, ␧ Si is the permittivity of silicon and n i is the intrinsic carrier density ͑ see Appendix A for a detailed derivation of this quantity in a quantum wire ͒ . The surface potential is ␺ s ε ␺ ( z ϭ t Si /2); ␺ s is also related to V GS and t ox through the boundary condition at the silicon-oxide interface ...
Context 2
... E n is the separation of the bottom of the n th- subband from the bottom of the conduction band, for a two- dimensional infinite square-well potential ͑ see Appendix B ͒ ; E 1 ( x max ) is the energy of the lowest subband, and ⌬ E 1 is the separation between E 1 ( x max ) and the bottom of the conduction band. Equation ͑ 14 ͒ can be numerically solved for the only unknown E FS Ϫ E 1 ( x max ) for a given V GS and V DS . We present now an illustrative example using this model. In this section, we show calculations of the transconductance ͑ Fig. 4 ͒ and conductance ͑ Fig. 5 ͒ of a quantum wire DG-MOSFET at different temperatures. The silicon film thickness and width are 2 and 3 nm, respectively. The oxide thickness is 1.5 nm, and the gate electrodes have a midgap gate work-function ͑ 4.61 eV ͒ . In Figs. 4 ͑ a ͒ and 4 ͑ c ͒ we show the bottom of the subbands at the virtual cathode, the surface, and center potentials, and the normalized current ( qI / G 0 ), all of them plotted versus V GS ͑ the Fermi level E FS is taken as a reference ͒ . Figures 4 ͑ b ͒ and 4 ͑ c ͒ show the transconductance at low ͑ 40 K ͒ and high temperatures ͑ 300 K ͒ , respectively. The increase of the gate voltage induces more mobile charge at the virtual cathode, pulling down the maximum of the barrier ͑ respect to the Fermi levels ͒ in order to inject this incremental charge. For the low temperature case, when the gate voltage is below 0.8 V all the subbands are above E FS , and the device operates in the subthreshold region. In this regime, both the current and the transconduc- tance are close to zero. At 0.8 V the threshold voltage , the first subband ( E 1 1 ) intersects the Fermi level, producing the first peak of the transconductance, and the device enters the above threshold saturation region. The threshold voltage is Ϸ 0.7 V at 300 K because the tail of the Fermi function popu- lates subbands above the Fermi level, contributing to the current before the first crossing. By increasing the gate voltage the subbands cross the source and drain Fermi levels producing a structure of peaks and valleys on the transconductance. This is very clear at low temperature, but at room temperature the structure is partially lost because the large extension over the energies of the tail of the Fermi function. In Figs. 5 ͑ a ͒ and 5 ͑ c ͒ we show the bottom of the subbands at the virtual cathode and the normalized current, plotted versus V DS . Note first of all the slight decrease of the bottom of the subbands with V DS . The mobile charge at the virtual cathode is constant, because V GS is fixed. When V DS is raised, the population of Ϫ k x states at the virtual cathode is gradually reduced. To maintain Q constant, the maximum of the barrier must be pulled down in order to inject more ϩ k x states from the source, and balance the Ϫ k x states. This effect occurs until the negative states population is totally suppressed at V DS,sat ϭ 0.55 V. For low V DS , four subbands are below the two Fermi levels, and the conductance is about 8 G 0 ͑ it would be exactly 8 G 0 at zero Kelvin ͒ , each subband contributing with 2 G 0 . Increasing V DS , the fourth, third, second, and first subband cross successively E FD , reducing the conductance to about 2 G 0 each one ͓ Fig. 5 ͑ b ͔͒ . The first subband crossing with E FD determines the onset of the saturation region. At T ϭ 300 K, saturation occurs at V DS,sat ϭ 0.6 V. The conductance structure is partially lost at this temperature because the tail of the Fermi function smooth the conductance ͓ Fig. 5 ͑ d ͔͒ . If the DG-MOSFET has a very large width W , the confinement in the y -direction is lost, and the electron gas movement is restricted only in the z -direction. In this limit case, the compact model provides good results for the voltage- current characteristics. A width W ϭ 500 nm is large enough to remove the confinement in the y -direction. We want to check if the important parameters like the threshold voltage, subthreshold swing, off-current, and on-current are well predicted by our model. For this purpose we have performed a simulation using a DG-MOSFET, with a channel length of 40 nm, a silicon film thickness of 1.5 and 3 nm, an oxide thickness of 1.5 nm, and gate electrodes having a work function of 4.25 eV. The device was simulated at low ͑ 100 K ͒ and room temperature ͑ 300 K ͒ . The results of our model are compared with accurate self-consistent simulations provided by the nanoMOS simulator, 8,9 based on the nonequilibrium Green’s functions. 17,18 This simulator deals with quantum well DG-MOSFETs, and it can consider up to six two- dimensional subbands. It can be seen from Fig. 6 ͑ a ͒ that the subthreshold region is well modeled, which exhibits the ideal subthreshold swing of 60 mV/decade at room temperature ͑ at 100 K this is Ϸ 20 mV/decade). Below the threshold voltage the subbands are all above the Fermi level, but they are populated because the tail of the Fermi functions extend over them ͑ Fig. 3 ͒ . The difference between the transmitted tails of the source and drain Fermi functions explains, in the Landauer approach, the diffusion current, which is the dominant current in the subthreshold region. The linear extrapolated threshold voltage ͑ at low drain bias and T ϭ 300 K), ex- tracted from a linear I Ϫ V GS plot is about 330 mV for t Si ϭ 1.5 nm and 195 mV for t Si ϭ 3 nm. It is well predicted by our model even at very small silicon film thickness, where deviation from classical models is very important. The nanoMOS results are 320 and 185 mV, respectively. The off- current and on-current are also well captured by our model ͓ Fig. 6 ͑ a ͔͒ . For a detailed inspection of the on-current we have plotted in Fig. 6 ͑ b ͒ the I Ϫ V DS characteristic for a DG- MOSFET with a silicon film thickness of 1.5 nm working at room temperature. The departure on the characteristics predicted by the compact model with respect to the calculated ones by the nanoMOS simulator arises mainly from the number of subbands considered in both the compact model and nanoMOS simulator, the former dealing with one- dimensional subbands and the latter with two-dimensional subbands. For the nanoMOS calculations we have considered the maximum number of allowed subbands by the pro- gram; in the compact model a number exceeding one hun- dred. In this section we extend the DG-MOSFET compact model formulated in the above sections to the cylindrical GAA-MOSFET. The following modifications have to be in- corporated: ͑ i ͒ the intrinsic carrier concentration, given by Eq. ͑ A3 ͒ , has to be divided by the area of the circular section ( ␲ t Si 2 /4) instead of the area of the rectangular section ( Wt Si ); ͑ ii ͒ the separation between the bottom of the conduction band and the bottom of the subbands is given by the theory of the cylindrical quantum potential well ͑ see Appendix B ͒ ; ͑ iii ͒ the new oxide capacitance ͑ per unit area ͒ for the cylindrical geometry is given by C ox ox 2 ln ͑ 1 2 t ox / t Si ͒ , 15 ͑ iv ͒ the potential in the channel is a solution of the Poisson’s equation for the cylindrical geometry ͑ see Appendix C ͒ . The surface and center potentials appearing in Eqs. ͑ 7 ͒ and ͑ 9 ͒ must be modified in accordance with it; and ͑ v ͒ for the cylindrical geometry, there are two sets of valleys, the first one is four-fold degenerated with a density of states effective mass m 1 d ϭ m T , and the second set, two-fold degenerated, has a density of states effective mass m 2 d ϭ m L . In order to test our approach we have compared the results given by our compact model with an experiment reported in the literature. 19 The GAA-MOSFET of the experiment has a diameter ( t Si ) of 65 nm, the insulator is silicon oxide with a thickness of 35 nm. The work-function of the gate material has been fixed to 4.43 eV in the compact model to match the experimental threshold voltage ͑ 0.48 V ͒ . To observe one-dimensional subband effects for this size, the operation temperature was 70 mK. In Fig. 7 we compare the measured transconductance ͑ dotted line ͒ with the simulated results by the compact model ͑ solid line ͒ . The position of the five main peaks is well predicted by the compact model. A detailed analysis reveals eight subbands below E FS in the analyzed voltage range. When the gate voltage is large enough the subband bottom energy reaches E FD . Every time one subband crosses E FS ( E FD ) a positive ͑ negative ͒ peak on the transconductance is recorded. The transconductance step is given by the derivative of the subband bottom energy with respect to the gate voltage multiplied by 4 G 0 or 2 G 0 , depending whether the subband belongs to the first or second set of valleys, respectively. The need for compact models of nanoscale transistors is increasingly important for future development of computer- aided-design tools at the circuit level. In this work we have presented a physics-based model for the ballistic quantum wire and quantum well MOSFET, derived from the Landauer transmission theory. The main novelty is that it extends modeling concepts of quantum well MOSFETs, to structures of lower dimensionality ͑ quantum wire ͒ , and is easily adaptable to gate structures as the single-gate, double-gate, or gate-all- around. The proposed model works in all the operation regions, below and above threshold, for low and high temperatures, incorporating effects of multisubband conduction, and taking into account the band structure of silicon. To verify the model we have proposed numerical experiments compar- ing the results with self-consistent simulations. The quantum wire model was used to predict the transconductance structure of a real device, showing good agreement. The intrinsic carrier density is evaluated by integrating the product N 1 D ( E ) • f ( E ) over all the energies in the conduction band and summing for all the subbands and ...

Similar publications

Article
Full-text available
A continuous analytical model for radiation-induced degradation in fully-depleted (FD) silicon on insulator (SOI) n-channel MOSFETs is presented. The combined effects of defect buildup in the buried oxide and band-to-band tunneling (BBT) have been shown to be the primary mechanisms that determine the radiation effects on the electrical characterist...

Citations

... Among many transistors, FETs with ballistic transport stand out as the ones that can reach the terahertz operational frequencies. That is why intense calculations and modeling of devices with minimized electron scattering was intensified in recent years [1][2][3][4][5][6][7][8][9][10][11][12][13][14][15][16]. Although the mean free path of electrons, for example, in Si is 100Å [1][2], the practical design is looking for FETs with channel lengths of a few hundred Ångströms. ...
... However, in our device electrons move through long enough channel of 400Å, where the quantization effects along the channel direction are insignificant, but somewhat longer distance could result in quasi-ballistic electron transport. The largest group of studies is focused on design of ballistic MOSFETs [11][12][13][14][15][16]. The most remarkable in this group of publications are K. Nishiguchi and S. Oda studies [14][15] of manufactured vertical MOSFET structure. ...
... where the effective mass of electron m * is 0. Here, the two sets of conductance fluctuations are related to the sub-band energy levels in the channel (E channel ) intersecting with the Fermi levels of the source (E fS ) and drain Fermi energy level (E fD ) in turn as V g increases. [22] The first sub-band in the channel intersects the source Fermi level (E fS = E channel ) at the position of V th (the conduction band edge), resulting in the current starting to rise rapidly as the first transconductance peak appearing. [23][24][25] The energy level in the channel and the energy barrier between the source and the channel are strongly controlled by the gate voltage. ...
Article
Full-text available
We investigate the influence of source and drain bias voltages ( V DS ) on the quantum sub-band transport spectrum in the 10-nm width N-typed junctionless nanowire transistor at the low temperature of 6 K. We demonstrate that the transverse electric field introduced from V DS has a minor influence on the threshold voltage of the device. The transverse electric field plays the role of amplifying the gate restriction effect of the channel. The one-dimensional (1D)-band dominated transport is demonstrated to be modulated by V DS in the saturation region and the linear region, with the sub-band energy levels in the channel ( E channel ) intersecting with Fermi levels of the source ( E fS ) and the drain ( E fD ) in turn as V g increases. The turning points from the linear region to the saturation region shift to higher gate voltages with V DS increase because the higher Fermi energy levels of the channel required to meet the situation of E fD = E channel . We also find that the bias electric field has the effect to accelerate the thermally activated electrons in the channel, equivalent to the effect of thermal temperature on the increase of electron energy. Our work provides a detailed description of the bias-modulated quantum electronic properties, which will give a more comprehensive understanding of transport behavior in nanoscale devices.
... As the silicon thickness is heavily scaled down to 10 nm, the threshold voltage is strongly affected by quantum confinement. This imposes a need for a closed form expression of the threshold voltage including quantum confinement for accurate circuit simulations [13]. ...
... In this model, the particle system considered is confined in two directions, and one transport direction in order to further extend its application to GAA structures. This is based on modelling a 1D quantum wire formed between the gates in the same manner as a quantum wire transistor [13]. ...
Article
Full-text available
Analytical Verilog-A compatible 2D model including quantum short channel effects and confinement for the potential, threshold voltage and the carrier charge sheet density for symmetrical lightly doped double-gate metal-oxide-semiconductor field effect transistors (MOSFETs) is developed. The proposed models are not only applicable to ultra-scaled devices but they have also been derived from 2D Poisson and 1D Schrödinger equations including 2D electrostatics, in order to incorporate quantum mechanical effects. Electron and hole quasi-Fermi potential effects were considered. The models are continuous and have been verified by comparison with COMSOL and BALMOS numerical simulations for channel lengths down to 7 nm at 1 nm oxide thicknesses; very good agreement within ±5% has been observed for silicon thicknesses down to 3 nm.
... Microscopic modeling of the drain current of such a nanotransistor thus requires the evaluation of the wave functions of the charge carriers [1][2][3][4][5][6][7][8][9][10] or the evaluation of Greens functions [11][12][13][14][15][16][17][18][19][20][21][22]. Because of the considerable effort for such microscopic calculations, there is a need for more flexible compact modeling like BSIM (Berkeley Short-channel IGFET Model) [23][24][25][26][27], the virtual source model [28][29][30][31][32][33] or wave function-based models [1, 5,7,[34][35][36][37][38][39][40]. In this paper, we focus on the latter approach in which the microscopic wave functions of the charge carriers are approximated in a compact form. ...
Article
Full-text available
One major concern of channel engineering in nanotransistors is the coupling of the conduction channel to the source/drain contacts. In a number of previous publications, we have developed a semiempirical quantum model in quantitative agreement with three series of experimental transistors. On the basis of this model, an overlap parameter 0 ≤ C ≤ 1 can be defined as a criterion for the quality of the contact-to-channel coupling: A high level of C means good matching between the wave functions in the source/drain and in the conduction channel associated with a low contact-to-channel reflection. We show that a high level of C leads to a high saturation current in the ON-state and a large slope of the transfer characteristic in the OFF-state. Furthermore, relevant for future device miniaturization, we analyze the contribution of the tunneling current to the total drain current. It is seen for a device with a gate length of 26 nm that for all gate voltages, the share of the tunneling current becomes small for small drain voltages. With increasing drain voltage, the contribution of the tunneling current grows considerably showing Fowler–Nordheim oscillations. In the ON-state, the classically allowed current remains dominant for large drain voltages. In the OFF-state, the tunneling current becomes dominant.
... This is based on modelling a 1D quantum wire formed between the gates in the same manner as a quantum wire transistor. [59] The system is first solved for one dimensional confinement, then solved for two dimensional confinement in order to be certain that it is valid for both cases. The first correction for the lowest confinement energy sub-band can be solved by: This formula leads to extensive calculations since the expression for is large. ...
Thesis
The downscaling of MOSFET devices leads to well-studied short channel effects and more complex quantum mechanical effects. Both quantum and short channel effects not only alter the performance but they also affect the reliability. This continued scaling of the MOS device gate length puts a demand on the reduction of the gate oxide thickness and the substrate doping density. Quantum mechanical effects give rise to the quantization of energy in the conduction band, which consequently creates a larger effective bandgap and brings a displacement of the inversion layer charge out of the Si/SiO2 interface. Such a displacement of charge is equivalent to an increase in the effective oxide layer thickness, a growth in the threshold voltage, and a decrease in the current level. Therefore, using the classical analysis approach without including the quantum effects may lead to perceptible errors in the prognosis of the performance of modern deep submicron devices. In this work, compact Verilog-A compatible 2D models including quantum short channel effects and confinement for the potential, threshold voltage, and the carrier charge sheet density for symmetrical lightly doped double-gate MOSFETs are developed. The proposed models are not only applicable to ultra-scaled devices but they have also been derived from analytical 2D Poisson and 1D Schrödinger equations including 2D electrostatics, in order to incorporate quantum mechanical effects. Electron and hole quasi-Fermi potential effects were considered. The models were further enhanced to include negative bias temperature instability (NBTI) in order to assess the reliability of the device. NBTI effects incorporated into the models constitute interface state generation and hole-trapping. The models are continuous and have been verified by comparison with COMSOL and BALMOS numerical simulations for channel lengths down to 7nm; very good agreement within ±5% has been observed for silicon thicknesses ranging from 3nm to 20nm at 1 GHz operation after 10 years.
... Semiconductor devices based on quasi-one-dimensional (quasi-1D) electron systems are attractive for high speed electronics. Field-effect transistors (FETs) with quasi-1D channels demonstrate near ballistic electron transport due to high drift electron velocities and a long mean free path of electrons [1,2]. Quasi-1D FETs can be realized by means of split-gates that induce confinement from quasi-2D to quasi-1D behavior of the electron transport [3]. ...
... The advantages of the given method are a relatively simple approach to fabrication and the fact that in the obtained structures the formation of Sn-NWs occurs simultaneously with the other layers in a single technological process of molecular-beam epitaxy (MBE). However, the mobility of the current carriers in such structures is strongly limited due to scattering on ion impurities and does not exceed 2000 cm V s 2 1 1 − − , that restricts use of them to the production of high speed devices. ...
... For this purpose T s was reduced to a value which could still ensure smoothness and crystallographic perfection of the growing layers. Nevertheless the grown samples of GaAs with Sn-NWs demonstrated low electron mobility μ = 1960 cm V s 2 1 1 − − . Therefore the materials with a reduced electron effective mass, for instance InGaAs [18], are preferable to GaAs. . ...
Article
Full-text available
The first pseudomorphic HEMT (PHEMT) with a doping profile of Sn nanowires (Sn-NWs) on a vicinal GaAs substrate has been demonstrated. Anisotropy of saturation current at 300 K was obtained when current flows along and across Sn-NWs. In the case of PHEMT orientation for a current flow along Sn-NWs, a higher power gain cut-off frequency was observed in comparison with PHEMT orientation for a current flow perpendicular to Sn-NWs (). We relate the obtained anisotropy of the static and dynamic characteristics of PHEMT to the formation of quasi-one-dimensional conductivity channels in a quantum well.
... is the radical effective mass for the valley concerned. Electron density is the statistics of each subband from Landauer theory [16]. Suppose the doping concentration in the source region is ; equation of charge neutrality is ∑ , , ...
... The transmission probability is either one or zero depending on whether the electron energy is larger or smaller than barrier energy or not. Drain current was then calculated with Landauer's theory [16]: ...
... where FD = FS − DS is Fermi level of drain region. Note that the subband levels in (16) are the most important parameter for determining the drain current. In the flatband condition, the subband levels are characterized by (4). ...
Article
Full-text available
A physically based subthreshold current model for silicon nanowire transistors working in the ballistic regime is developed. Based on the electric potential distribution obtained from a 2D Poisson equation and by performing some perturbation approximations for subband energy levels, an analytical model for the subthreshold drain current is obtained. The model is further used for predicting the subthreshold slopes and threshold voltages of the transistors. Our results agree well with TCAD simulation with different geometries and under different biasing conditions.
... No doubt, the model of the ballistic MOSFET proposed by Natori [40], extended by Lundstrom [44] to include scattering, also manage to model the essential device physics in an elegant manner through simple mathematical formulations. However, these and the derivative works [47,65,66,67] fail in providing a consistent set of charge, potential and drain current equations, that are seamlessly scalable from the ballistic to the long channel diffusive regime. Apart from the initial works, most of the subsequent models cannot even be classified as compact models because they rely on numerical solutions of integrals. ...
Thesis
Full-text available
The scaling of device technologies poses new challenges, not only in circuit design, but also in device modeling, especially because of the short-channel effects and the emergence of novel phenomena like ballistic transport. Nonetheless, it enables the design of ultra low-power analog and Radio Frequency (RF) circuits by allowing to push the operating points intomoderate and eventually weak inversion regions, which are increasingly becoming the preferred regions of operation for such applications. Even though modern compact models have evolved to adequately model the short-channel effects in all regions of operation, there is a lack of simpler models that (a) reliably predict the physics of downscaled devices while (b) remaining continuous through moderate inversion and (c) aid the designer’s intuition through simple designmethodologies. In this work, we extend the EKV charge based model to include the velocity saturation effect for weak inversion operation. Using the simple analytical model hence developed, we propose a design methodology for low-power analog circuit design. Then, we focus our attention on ballistic transport in MOSFETs, that is expected to dominate in the deeply scaled devices. Again, despite the extensive body of work available in the literature, most models remain deeply rooted in physics, consisting of fairly complicated equations, that are of little use for an intuitive understanding and design. In addition, the quasi-ballistic devices, which lie on the continuumbetween the ballistic and the diffusive devices, pose their own modeling challenges: a model for the quasi-ballistic devices would have to remain continuous between the ballistic and diffusive regimes. Most of the published works, based on the carrier flux transport over the source-channel potential barrier approach, seem to ignore the electrostatics in the rest of the channel. The shape of the electrostatic potential in the channel is approximated through polynomial functions, which is adequate for the very short-channel devices but not scalable to long channel quasi-ballistic devices. In this work, we study the role of the gate and the electrostatics in a ballistic channel by drawing on the insights gained from Monte-Carlo simulations on quasi-ballistic and ballistic doublegate MOSFETs. We propose a simple semi-empirical model of the channel charge, using which we develop an analytical model for the channel potential, both of which could be used as precursors to a scalable compact model that would encompass the ballistic, quasi-ballistic and drift-diffusion regimes.
... However, a net flow of forward electrons, with energies lying between E FS and E FD appear because energies above the Fermi level corresponding to the backward electrons are unpopulated (at zero Kelvin operation temperature). The effect of an applied voltage over the gate and drain electrodes on the potential barrier shape is illustrated in Fig. 1d (Jimenez et al. 2003) (the Fermi level is taken as a reference). The separation between the Fermi level and the maximum of the barrier increases with the applied gate-source voltage (V GS ). ...
... Equation (9) was observed before for a Si-quantum wire (see for example Jimenez et al. 2003). The total density of states is then given by the sum of the density of states of all the populated sub-bands, which can be written as: ...
... Since the nanowire channel is isolated from any other source of mobile carrier, source and drain are the sole source of carriers inside the nanowire channel. A detailed description of this generic picture of nanowire MOSFET can be found in existing literature (Jimenez et al. 2003;Wang et al. 2003) and hence is avoided here. In such an electrostatic system, the carrier density for all valleys of any subband (nth) of a nanowire channel can be expressed as ...
Article
Full-text available
We studied the physics insight the GaN (example) quantum wire FET transistors. The model is based on the four $\mathbf{k}{\cdot } \mathbf{p}$ k · p Kane band model. We have introduced closed compact model for the Einstein relation of the diffusivity to mobility ratio (DMR) in quantum wires. The model can be applied for both wide and narrow band gaps of nonparabolic conduction band dispersion. The model is related to the optical matrix elements between conduction and valence bands. We have used 1D electrostatic to model the electron density over the maximum energy point. We have studied the effects of gate-to-source and drain-to-source voltages on the DMR by calculating the electron density using flux theory. We observed that above the threshold the non-parabolic dispersion increases the DMR. Additionally, we have studied the nonparabolic effects on the Fermi level and found that for low doping concentrations, the nonparabolic effect must be considered and an accurate calculation for the optical matrix elements is needed.
... where φ is the electrostatic potential referred to the Fermi level in the source [17], [19], [20], and the electron density is given as ...
... where from [19] and [20] ...
... The solution for the 1-D potential term φ 0 (r) is given by [5], [19], [20] ...
Data
Full-text available
Analytical physically based models for the threshold voltage, subthreshold swing, and drain-induced barrier lowering (DIBL) of undoped cylindrical gate-all-around MOSFETs have been derived based on an analytical solution of 2-D Poisson’s equation (in cylindrical coordinates) in which the mobile charge term has been included. Using the new model, threshold voltage, DIBL and subthreshold swing sensitivities to channel length, and channel thickness have been investigated. The models for DIBL, subthreshold swing, and threshold voltage rolloff parameters have been verified by comparison with 3-D numerical simulations; close agreement with the numerical simulations has been observed.