Introduction

Aggressive transistor scaling with the aim of boosting on-die capabilities has led to an adverse effect on power dissipation. Due to increasing insistence on power dissipation, low voltage requirements, energy conservation, and maximising on-die capabilities has led the devices having the prospect of acquiring sub-threshold swing (SS) surpassing the thermionic limit of 60 mV/dec has earned extensive attention1,2,3. Tunnel field-effect transistors (TFETs) have unique carrier injection mechanisms and inversion layer formation when compared to conventional MOSFETs. These lead to substantial improvement of SS, ION/IOFF ratio, and, as such, TFET devices have been one of the prominent contenders for replacing conventional MOSFETs4,5. However, TFETs have encountered severe ambipolar leakage current (IAMB), low ON-state current (ION), and gradual shifting between ON and OFF states6. With continuous scaling of devices, the reduction of power dissipation remains a primary requisite along with the enhancement of device performance as far as system-on-chip based CMOS integrated circuits are concerned. In addition, the gate-all-around (GAA) architecture has demonstrated enhanced electrostatic control, tunneling efficiency, off-state leakage current, and current drivability, thereby, resulting in the improvement of analog and RF performances7,8,9.

The steps involved in the fabrication of the vertical GAA structure involve several etching, deposition, and growing processes, which generally form an uneven radius rather than the ideal circular GAA structure10,11. Because of this uneven radius, predicting the performance of fabricated GAA device structures is difficult12,13, and only a few reports for its acceptable variations14,15,16,17 are available. Most of the reports only address one type of geometry and are of mostly MOSFET devices, as the report12 shows the impact of hydrogen annealing on the device diameter of GAA MOSFET devices, which led to the elliptical circumference having an ON current of 825 µA/µm, report13 shows the fabricated GAA MOSFET device having an elliptical shape (i.e., EOEC geometry) having ON current of 976 µA/µm, report15 demonstrate the impact of varying channel length on fabricated elliptical GAA MOSFETs, report14 demonstrates numerical analysis on three different types of variations such as tsi fluctuation, elliptical shape fluctuations (i.e., EOEC fluctuations), and corner rounding fluctuations where the acceptable tolerance for RF IC design is analyzed as Δtsi < 1 nm and r/R > 75%, and the report16 mathematically analyzed the short-channel-effects of elliptical GAA MOSFET (i.e., EOEC structure) by varying its effective radius. The experimental18,19 and numerical analyses20 suggest that incorporating III-V broken-bandgap semiconductor material in TFET devices shows improved tunnelling efficiency, leading to enhanced device performance21,22. The ternary compound material of type-II heterostructure such as GaSb/InGaAs gained importance due to its tunable band alignment, direct bandgap, and effective bandgap by varying the mole fraction23,24,25. There are some challenges while fabricating III–V semiconductor TFETs, such as a lack of dielectric material, which causes thermodynamically unstable interface states, thereby leading to fermi level pinning at the interface. Moreover, uncontrolled oxidation on III–V material results in a higher density of bandgap traps at the interface. However, these challenges are encountered by using the interface control (i.e., passivation) layer at the interface26.

In this report, the issues related to fabricating vertical GAA TFET are considered where the most unlikely uneven radius is considered to be elliptical in shape for all possible GAA fabrication defects. In addition, a broken-gap type-II heterojunction of GaSb/InGaAs material has been introduced23, which further boosts the device performance. The objective of this paper is to design and optimize III–V elliptical gate-all-around TFET (III–V eGAA TFET) by varying the minor diameter (MD) and gate dielectric thickness (TOX) of the channel region to get the optimized MD and TOX of the elliptical structure. The optimization process is carried out by considering the higher ION, lower IAMB, low parasitic capacitance, minimum gate delay (τm), and higher mobility. Finally, the optimal structure is investigated for analog and RF performance.

Device structure and simulation methodology

The 3D and cross-sectional schematics of the simulated III–V eGAA TFET structure for elliptical oxide–elliptical channel (EOEC) is shown in Fig. 1, and its corresponding band diagram for circular oxide–circular channel (COCC) for ambipolar, off, and ON state conditions are depicted in Fig. 2. The device under simulation incorporates type-II heterojunction material such as GaSb/InGaAs23. In addition, the uneven radius of the fabricated device is considered elliptical in nature, having a major diameter of 20 nm, and the optimal value of the minor diameter (MD) is calibrated by varying it from 10 to 20 nm. The device also incorporates the source doping (NA) of 1020 cm−3, intrinsic channel doping (NI) of 1015 cm−3, drain doping (ND) of 5 × 1019 cm−3, gate length (LG) of 30 nm, and maximum gate oxide thickness (TOX) of 2 nm. The inset of our simulation is in agreement with ITRS27 and the targeted value considered in our simulation is tabulated in Table 1, where the work function of the metal and supply voltage (VDD) are set at 5.01 eV (i.e., nickel) and 0.3 V, respectively. The simulations are carried out by the 3D numerical device simulator TCAD Sentaurus28 with an optimized meshing strategy as discussed in29. The non-local band-to-band tunnelling (BTBT) model has been employed in our simulation. Apart from this, carrier scattering due to the presence of charged impurity ions and velocity saturation at high-field is activated by considering concentration-dependent and high-electric field-dependent mobility models. The mobility degradation model at interfaces is also included to compute the transverse field, and the old Slotboom model is incorporated to compute the doping-induced bandgap narrowing of the material. Moreover, the effect of the strain through deformation in the lattice structure, causing shifts in the conduction and valance bands, is captured by the piezo deformation potential model. Finally, the quantum correction model is incorporated to take into account the change in the effective bandgap of the device due to the variation in MD from 10 to 20 nm. The non-local BTBT parameters of Sentaurus TCAD for InGaAs and GaSb are used as described in30.

Figure 1
figure 1

EOEC eGAA TFET structure: (a) 3D view, (b) cross-sectional view.

Figure 2
figure 2

Band diagram of COCC structure: (a) ambipolar-state, (b) off-state, (c) ON-state.

Table 1 ITRS 2013 (predicted LP technology requirements of multi-gate (MG) FET devices for year 2021).

Proposed fabrication steps

The process steps for fabricating III–V eGAA TFETs are proposed in Fig. 3. The sequential process steps which can be employed for the fabrication are: (i) Layers of GaSb (p-type, Be ~ 1020 cm−3), 30 nm of InGaAs (intrinsic, Si ~ 1015 cm−3), and 35 nm of InGaAs (n-type, Si ~ 5 × 1019 cm−3) need to be deposited layer-by-layer as shown in Fig. 3a. Such a deposition can be carried out by molecular beam epitaxy (MBE). Then, a nitride hard mask (NHM) is deposited on top of the surface followed by a spreading polystyrene sphere (PSS) to self-assemble on top of the surface (Fig. 3b). To reduce the PSS diameter, the reactive ion etching (RIE) process can be used, which forms colloidal particle nanodots that are no longer closely packed (i.e., space is created for gate oxide and contact pad deposition) for getting the GAA structure. (ii) Tetramethylammonium hydroxide (TMAH) based wet etch process needs to be carried out to remove the material underneath the NHM patterned by PSS (Fig. 3c). The next step involves etching away the PSS with the dissolution of CHCl3 (Fig. 3d).

Figure 3
figure 3

Schematic of III–V eGAA TFET fabrication steps: (a) III–V material deposition, (b) PSS placement, (c) RIE on PSS to get 20 nm sphere followed by NHM deposition, (d) deep RIE etching process, (e) PSS of 20 nm diameter and NHM etched away, (f) SiO2 deposition, (g) PSS of 24 nm diameter and NHM deposition, (h) DHF etch-back up to the source-channel junction, (i) PolySilicon deposition, (j) SiO2 deposition, (k) contact tip implantation followed by contact formation.

Further, a non-conformal high-density plasma (HDP) oxide is deposited (Fig. 3e), followed by dilute hydrofluoric acid (DHF) etch-back to get HDP up to the source-channel junction (Fig. 3f). (iii) To grow gate oxide high-density plasma chemical vapor deposition (HDPCVD) technique can be used (Fig. 3g), followed by the deposition of poly-silicon (Fig. 3h). (iv) Again, non-conformal HDP oxide is deposited (Fig. 3i) and DHF etch-back to get HDP up to the channel-drain junction (Fig. 3j). Further, etch-back the poly-silicon, gate oxide, and NHM to expose the drain region (Fig. 3k). Finally, non-conformal HDP oxide is deposited again and patterned to form the contact pads (Fig. 3l).

We have considered that the RIE process is involved in reducing the PSS and the grown gate oxide, which accounts for the uneven channel radius and TOX variation induced in the fabricated device. The worst-case scenario of structural variations involved in the fabrication process is considered elliptical in shape as depicted in Fig. 4, where, Fig. 4a shows the ideal circular GAA structure having a circular oxide-circular channel (COCC). Figure 4b to Fig. 3d represent elliptical oxide—elliptical channel (EOEC), elliptical oxide—circular channel (EOCC), and circular oxide—elliptical channel (COEC), respectively.

Figure 4
figure 4

Schematic representation of possible cross-sectional variations of GAA structure due to fabrication imperfections: (a) circular oxide—circular channel (COCC), (b) elliptical oxide—elliptical channel (EOEC), (c) elliptical oxide—circular channel (EOCC), (d) circular oxide—elliptical channel (COEC).

Structural optimization

Device optimization for EOEC and COEC structure is carried out by evaluating optimal channel MD by varying it from 10 to 20 nm. For the EOCC structure, optimization is done by varying TOX from 1 to 2 nm along the minor axis (MA).

For optimization of the EOEC structure, the major diameter of the channel region is fixed at 20 nm, while the MD is decreased from 20 nm (which represents a circular GAA TFET) to 10 nm (eGAA TFET), simultaneously maintaining the same gate oxide thickness of 2 nm throughout the entire elliptical periphery. The results of our simulation are depicted in Fig. 5a. Starting from 20 nm, as the channel, MD is reduced, the band bending between the drain-channel and source-channel region gradually increases as depicted in Fig. 6a. The former increases the inflow of carriers from drain to channel, increasing IAMB of the device; the latter increases the inflow of carriers from source to channel, increasing JDS of the device. As the channel MD is reduced, the area of the device decreases gradually, which consequently decreases Cpar followed by a decrease in its τm. For EOEC structure optimization, the point at which the structure has maximum JDS and µcarr, and the minimum possible IAMB, Cpar, and τm is considered to be the optimized geometry. In this report, an EOEC structure having a MD of 14 nm is considered to be the optimized device.

Figure 5
figure 5

Normalized values for different eGAA TFET along: (a) MD for EOEC optimization, (b) MA of EOCC structure for gate oxide thickness optimization, (c) MD for COEC optimization.

Figure 6
figure 6

Band diagram for different eGAA TFET along: (a) MD for EOEC optimization, (b) MA of EOCC structure for gate oxide thickness optimization, (c) MD for COEC optimization.

For EOCC structural optimization, the channel region is fixed at 20 nm in diameter from all sides, while the TOX is reduced from 2 nm (which represents circular GAA TFET) to 1 nm (eGAA TFET) along MA, while keeping a fixed TOX of 2 nm along the major axis. The results of our simulation are depicted in Fig. 5b. Starting from 2 nm, as the TOX along MA is reduced, the vertical electric field penetrates more into the substrate along MA, which consequently increases JDS, and the same is verified from the energy band diagram depicted in Fig. 6b where tunneling width becomes narrow with reducing TOX along MA. As the vertical electric field is penetrating more into the substrate causes a narrowing of depletion width at the drain to the channel region. The reduced depletion width at the drain to channel region gives rise to the inflow of carriers from drain to channel, which leads to an increase in IAMB, followed by an increase in its Cpar. However, τm depends on JDS and the total gate capacitance of the device, where JDS shows the dominant factor which leads to a decrease in its τm. For EOCC structure optimization, the point at which the structure has maximum JDS and µcarr, and the minimum possible IAMB, Cpar, and τm is considered to be the optimized geometry. In this report, the EOCC structure having a TOX of 1.4 nm along MA is considered to be the optimized device.

Finally, for COEC structural optimization, the outer diameter (both channel and TOX together) is fixed at 24 nm from all sides, while with a decrease in channel MD from 20 to 10 nm (all are having an outer diameter of 24 nm) lead to virtual increment of TOX from 2 to 7 nm along MD of the device as depicted in Fig. 4d. Result of our simulation depicted in Fig. 5c does not explicitly show the effect of variation of neither channel MD nor TOX characteristics as obtained in Fig. 5a,b. For COEC characteristics, the impact of channel MD is observed up to certain channel MD, thereafter the impact of TOX plays a significant role as depicted in Fig. 5c. Starting from 20 nm, as the channel, MD is reduced, the JDS decreases up to 14 nm considering the same reason stated for Fig. 5a, thereafter JDS increases due to the impact of TOX which becomes more dominant than the impact of channel MD and follows the same reason stated for Fig. 5b31. The variation in current density is also verified by the energy band diagram depicted in Fig. 6c. For COEC structure optimization, the point at which the structure has maximum JDS and µcarr, and the minimum possible IAMB, Cpar, and τm is considered to be the optimized geometry. In this report, a COEC structure having a channel MD of 18 nm is considered to be the optimized device. It is observed that the channel MD of 14 nm and 18 nm for EOEC and COEC and TOX of 1.4 nm for EOCC exhibit optimized device performance.

Further, the impact of mechanical strain generated during the fabrication process, specifically at the junctions due to lattice mismatches, has been considered in the analysis. These lattice mismatches cause variation in the band structure and carrier mobility of the device, which affects the drain current, as shown in Fig. 7a. To encounter these variations, piezo deformation potential (PDP) has been invoked to get closely matched fabricated device characteristics. It is observed from Fig. 7 that, when PDP is included, then the JDS shows an increment of ~ 3.5%, ~ 3%, ~ 0.5%, and ~ 0.4% for COCC, COEC, EOCC, and EOEC, respectively, when compared to the device which does not have PDP involved in its numerical simulation analysis. Finally, the impact of structural deformation on threshold voltage and ON-state drain current of the device is depicted in Fig. 7b. It is observed that when the channel area is large and the oxide width is comparatively thinner for the EOCC structure, then ON-state drain current is high whereas threshold voltage is comparatively low. The impact is clearly visualized when we compare EOEC and COCC structures, where oxide width is constant and channel area reduces, which causes an increase in ON-state drain current and a decrease in the threshold voltage of the device because channel area is the dominating factor which reduces current density with a reduction in the channel area. On the other hand, the impact of oxide is visualized when we compare EOCC and COEC structures, where oxide width along the minor axis is significantly high. This high oxide width causes a decrease in the device's ON-state drain current and a significant increase in the device's threshold voltage, as shown in Fig. 7b. The detailed discussion of different optimized geometries due to varying channel area and oxide width on analog and RF performance is analyzed in the result and discussion section.

Figure 7
figure 7

(a) Comparison of variation of drain current density (JDS) as a function of gate-to-source voltage at VDS = 0.3 V for all possible optimized structures when PDP is invoked, and (b) variation in threshold voltage and ON-state drain current as a function of deformation in device structures.

Results and discussion

This section presents the analog and RF performances of III-V eGAA TFET considering the channel MD of 14 nm and 18 nm for EOEC and COEC structures whereas TOX of 1.4 nm for EOCC structures to get the better device performance at 300 oK.

Effect of process variation on analog performance parameter

The impact of channel MD and TOX on the analog performance parameters such as transconductance (gm), output conductance (gD), transconductance generation factor (gm/IDS), and intrinsic gain (gmRO) for III-V eGAA TFET structure are analyzed in this section. It is observed from Fig. 8a that the COCC structure has an accumulation of band-to-band generated tunneling carriers from all sides of the gate in a circular fashion and are generally present near the semiconductor-dielectric interface. Whereas, due to the vertical electric field of EOEC structure penetrating more at the center of the device as depicted in Fig. 8b, which results in tunneling of the carrier at the center of the device rather than near the semiconductor-dielectric interface32. As a result, the carrier density increases at the center of the device which narrows the tunneling width at the source-channel junction as depicted in Fig. 6b33. Therefore, the band-to-band generated tunneling carriers take the shape of the elliptical channel and the accumulated carriers shift from the semiconductor-dielectric interface to the center of the device34.

Figure 8
figure 8

Band generation rate of carriers for (a) COCC, (b) EOEC, (c) EOCC, and (d) COEC.

For EOCC structure depicted in Fig. 8c shows that the vertical electric field penetrated more into the region where TOX is minimum causing carriers to accumulate at the region where TOX is minimum. Finally, for COEC structure depicted in Fig. 8d, band-to-band generated carriers follow the same phenomenon stated for Fig. 8b,c considering the dominancy of either channel MD of EOEC structure or TOX of EOCC structures. However, the device having channel MD of 18 nm of COEC structure shows that the carriers are accumulating below the major axis where TOX is fixed at 2 nm17,32. The vertical electric field penetrates more with decreasing channel MD of EOEC structure resulting in an increase in its carrier density33 as depicted in Fig. 9a. As electron density increases cause more carriers to tunnel from the valence band of the source to the conduction band of the channel, which enhances the tunneling probability of carrier at the source to channel junction35. The insertion in Fig. 9a shows the increment in drain current density of optimized EOEC from ~ 3 nA/(nm)2 to ~ 6 nA/(nm)2 from its circular counterpart.

Figure 9
figure 9

Comparison of variation of drain current density (JDS) as a function of gate-to-source voltage at VDS = 0.3 V of COCC structure with (a) EOEC, (b) EOCC, and (c) COEC structure.

In the case of varying TOX keeping fixed channel MD of 20 nm of EOCC structure causes faster accumulation of the carrier density near the semiconductor-dielectric interface where TOX is minimum as depicted in Fig. 9b. However, the gradual increase in the electrostatic control of the EOCC structure shows an increment in its drain current density with a decrease in TOX of the device along MA. The insertion in Fig. 9b shows the increase in the drain current density from ~ 3 nA/(nm)2 to ~ 5 nA/(nm)2 when compared to COCC structure.

On the contrary, the COEC structure shows an unpredictable trend even after keeping the outer diameter fixed at 24 nm. It is evident from Fig. 8d that, varying channel MD causes virtual variation in TOX of the device which leads to accumulation of carrier density depending on the dominancy of either channel MD or TOX of the device. Result of our simulation shown in Fig. 9c shows that when channel MD is 10 nm then TOX along the channel MD is maximum and has a drain current density of ~ 2.2 nA/(nm)2 whereas with the gradual increase in channel MD from 10 to 14 nm shows the effect of TOX on the device is relatively higher than the variation of MD of the channel on the device. As a result, drain current density decreases from ~ 2.2 to ~ 1.9 nA/(nm)2 from channel MD of 10 nm to 14 nm. However, when channel MD increases beyond 14 nm, the impact of variation of channel MD is more in the device when compared to the virtual variation of TOX along MA. The enlarged view under Fig. 8c depicts that, the drain current density increases beyond 14 nm whereas the optimized COEC structure shows a drain current density of ~ 2.4 nA/(nm)2.

For the evaluation of the transconductance (gm) and the drain conductance (gD) of the III-V eGAA-TFETs, the following expressions are used,

$$g_{m} = \frac{{\partial I_{DS} }}{{\partial V_{GS} }},$$
(1)
$$g_{D} = \frac{{\partial I_{DS} }}{{\partial V_{DS} }}$$
(2)

The gm and gD expressions depend on the change in IDS with changing applied port biasing of TFET33. However, the dependency on IDS further led to a dependency on the tunneling probability of the carriers at the source-to-channel junction, effective mobility (µeff)\(\left({\upmu }_{\mathrm{eff}}\right)\), tunneling barrier width, or average tunneling thickness (λ)\(\left(\uplambda \right)\) of the device.

For EOEC and EOCC structures, their tunneling barrier width becomes narrower which increases the tunneling probability, thereby, gm and gD show a rapid increase with VGS as shown in Fig. 10 and Fig. 11. On the other hand, for COEC structure, their tunneling barrier width depends on both the variations (channel MD and TOX) which result in a decrease in gm and gD till 14 nm than a gradual increase beyond 14 nm along channel MD of the device as shown in Fig. 12.

Figure 10
figure 10

Comparison of variation of transconductance (gm) and output conductance (gd) as a function of gate voltage at VDS = 0.3 V for EOEC devices: (a) gm, and (b) gd.

Figure 11
figure 11

Comparison of variation of transconductance (gm) and output conductance (gd) as a function of gate voltage at VDS = 0.3 V for EOCC devices: (a) gm, and (b) gd.

Figure 12
figure 12

Comparison of variation of transconductance (gm) and output conductance (gd) as a function of gate voltage at VDS = 0.3 V for COEC devices: (a) gm, and (b) gd.

The performance parameters of available gain per unit value of power dissipation, quality factor (gm/IDS)35, and intrinsic gain (gmRO)30 of the device as a function of VGS at VDS = 0.3 V are shown in Fig. 13. Since, gm and gD increase with increasing VGS, therefore with a decrease in channel MD and TOX of EOEC and EOCC structures show a significant increase in gm/IDS and gmRO characteristics. On the other hand, gm and gD of COEC structure decreases when channel MD decreases from 14 to 10 nm whereas when channel MD decreases from 20 to 14 nm causes an increase in gm and gD of COEC structure therefore, gm/IDS and gmRO first decrease from 20 to 14 nm channel MD then it increases from 14 to 10 nm channel MD for COEC structure.

Figure 13
figure 13

Comparison of variation of transconductance generation factor and intrinsic gain as a function of gate voltage at VDS = 0.3 V of COCC structure with: (a) EOEC, (b) EOCC, and (c) COEC structures.

For gm/IDS in the subthreshold region, the transconductance and the drain currents for both variations are negligible. However, the peak of the gm depicts the threshold voltage of the device. From Eq. (2), the drain conductance of the device depends on the tunneling current of TFET33. Therefore, increasing the tunneling carrier at the source-to-channel junction causes gD of the device to increase, which consequently leads to the reduction of output resistance (RO) of the device. It is observed from Fig. 13a that the optimized EOEC structure shows an increment in gm/IDS from ~ 86 V−1 to ~ 117 V−1 and an increment in gmRo from ~ 33 to ~ 90 respectively when compared to COCC structure. Moreover, Fig. 13b depicts, the increment in gm/IDS of optimized EOCC from ~ 86 to ~ 96 V−1 and increment in gmRo from ~ 33 to ~ 79 respectively when compared to COCC structure. Finally, Fig. 13c depicts the COEC structure where gm/IDS and gmRo first decreases from ~ 86 V−1 and ~ 33 (MD = 20 nm) to ~ 78 V−1 and ~ 21 (MD = 14 nm) then gm/IDS and gmRo increases to ~ 79 V−1 and ~ 24 respectively. The insertion in Fig. 13c shows the optimized COEC structure having gm/IDS and gmRo as ~ 81 V−1 and ~ 27 respectively.

Effect of process variation on RF performance parameter

The comparison of variations on EOEC, EOCC, and COEC from its equivalent COCC structure has substantially boosted the RF performance parameters. The performance variation of CGD, CGS, and CGG, along with cut-off frequency (fT) and gate delay (τm) is analyzed in this section.

For a fixed VDS, the depletion region at the source-to-channel region is formed due to reverse biasing of the source-to-channel junction, which gives rise to CGS36. However, at low VGS, the depletion charge near the drain side decreases with a decrease in the area of the channel, which causes a decrease in CGD. When high VGS is applied, the channel area decreases, causing a negligible change in inversion charge at the source end. As a result, CGS decreases with a decrease in the area of the channel. Whereas, the inversion charge increases with a decrease in the area of the channel at high VGS. Therefore, CGD increases with a decrease in the area of the channel37.

The variation of CGS along with CGD as a function of VGS for EOEC, EOCC, and COEC structures is plotted in Fig. 14. For a fixed VDS, when channel MD of the EOEC structure decreases from 20 to 10 nm, it causes a decrease in the area of the GAA structure, which consequently reduces the depletion charge at the source-to-channel junction at low VGS. Whereas, at high VGS, a decrease in channel MD shows negligible variation in the inversion charge at the source end. As a result, CGS decreases with the decrease in channel MD of the EOEC structure. The optimized EOEC structure shows a decrease in CGS from ~ 4.5 to ~ 3.5 aF when compared to the COCC structure. On the other hand, a decrease in channel MD of the EOEC structure at low VGS causes a negligible change in depletion charge at the drain-to-channel region. Therefore, CGD follows the same trend with a decrease in channel MD of the EOEC structure. However, at high VGS, inversion charge experiences a sudden increase therefore, CGD experiences cross-over and increases with a decrease in channel MD of EOEC structure from ~ 34 to ~ 37 aF when compared to COCC structure as depicted in Fig. 14a.

Figure 14
figure 14

Comparison of variation of gate-to-source capacitances and gate-to-drain capacitances as a function of gate voltage at VDS = 0.3 V of COCC structure with (a) EOEC, (b) EOCC, and (c) COEC structures.

For the EOCC structure, at low VGS, decreasing the TOX along MA shows a negligible change in its inversion charge at the source end. As a result, CGS shows negligible variation at low VGS. On the other hand, at high VGS, decreasing TOX causes a gradual increment in the electric field which subsequently increases the inversion charge near the drain end. Thus, CGD increases with decreasing TOX for EOCC structure which leads to an increase in CGD from ~ 34 to ~ 43 aF when compared to its COCC structure as depicted in Fig. 14b.

On the contrary, unlike the analog performance parameter, the RF performance of COEC follows the predictable trend due to its dependency on the area and TOX of the GAA structure. As the reason stated, the source side depletion charge at low VGS increases with decreasing channel MD of COEC structure, however it experiences slight reduction with virtual increment in TOX where depletion charge at source end is dominated by the variation of channel MD of COEC structure. Furthermore, at high VGS, the inversion charge experiences a slight increment at the source side with decreasing channel MD. Therefore, CGS is dominated and decreases with decreasing channel MD of COEC structure. The optimized COEC structure shows a decrement in CGS from ~ 4.5 to ~ 4.2 aF when compared to COCC structure. As channel MD of COEC is dominated by virtual variation of TOX, therefore, at high VGS, CGD experiences sudden decrement in inversion charge when compared to its circular structure with a decrease in CGD from ~ 34 to ~ 27 aF from COCC structure as depicted in Fig. 14c.

The total capacitance CGG = CGS + CGD \(\left({\mathrm{C}}_{\mathrm{GG}}={\mathrm{C}}_{\mathrm{GS}}+{\mathrm{C}}_{\mathrm{GD}}\right)\) is plotted in Fig. 15 which follows the variation in CGD as depicted in Fig. 14. It is evident from Fig. 15 that, CGG follows the CGD graph which is the dominant contributor to total gate capacitance for any variation in COCC, EOEC, EOCC, and COEC structures. At high VGS, CGG is almost the same as CGD whereas, at low VGS, CGG shows the same trend as CGS which is higher in magnitude at low VGS.

Figure 15
figure 15

Comparison of variation of total gate capacitance as a function of gate voltage at VDS = 0.3 V of COCC structure with: (a) EOEC, (b) EOCC, and (c) COEC structures.

The cutoff frequency (fT) and the gate delay (τm) are other parameters for the analysis of RF performances. The cutoff frequency and the gate delay are defined by

$$f_{T} = \frac{{g_{m} }}{{2\pi (C_{GS} + C_{GD} )}}$$
(3)
$$\tau_{m} = \frac{{C_{GG} V_{DD} }}{{I_{ON} }}$$
(4)

From the fT expression (3), it is evident that it depends on the transconductance (gm) and the total capacitance CGG = CGS + CGD values, which is a function of VGS38. The characteristics of fT are explained explicitly for the subthreshold and superthreshold regions for all device structures. It is observed from Fig. 15 that in the weak inversion region, for each variation, the value of the intrinsic capacitance is low. However, the value of gm increases rapidly in the weak inversion region for variation in any device structure. Therefore, the value of fT at a low VGS value is dominated by the value of gm as shown in Fig. 16. Whereas, in the strong inversion region, the intrinsic capacitance value shows a rapid increase with VGS. However, the gm of the device is almost invariable at the strong inversion region. Thus, in strong inversion, the fT parameter is dominated by the value of intrinsic capacitance. Results of the simulation depicted in Fig. 16a that, with the decrease in the channel MD of the EOEC structure, the fT of the device increases where the optimized EOEC structure shows increment from ~ 31 to ~ 43 GHz cutoff frequency and increment from ~ 31 to ~ 40 GHz cutoff frequency on EOCC structure when compared to COCC structure. On the other hand, for the COEC structure, intrinsic capacitance is relatively higher. Therefore, the value of fT at a low VGS value has an impact on both the intrinsic capacitance and the gm of the device. As a result, the decrease in MD of the COEC structure causes a decrease in fT of the device where the optimized COEC structure shows a decrement from ~ 31 to ~ 26 GHz when compared to its circular structure as shown in Fig. 16c.

Figure 16
figure 16

Comparison of variation of cut-off frequency and gate delay as a function of gate voltage at VDS = 0.3 V of COCC structure with: (a) EOEC, (b) EOCC, and (c) COEC structures.

The variation of gate delay (τm) as a function of VGS is also plotted in Fig. 16. The gate delay expression (4) stated earlier has supply voltage (VDD), gate capacitance (CGG), and ON-state current (ION) dependence.

It is observed from Fig. 15a that, with a decrease in channel MD of EOEC structure causes CGG and drain current to increase following the reason stated in analog performance. The rate of increase of drain current is dominated by the increase in CGG of the device, which consequently reduces the τm. The insertion in Fig. 16a depicts the decrease in the τm of optimized EOEC structure from ~ 8.5 to ~ 5.5 psec when compared to COCC structure. For EOCC structure, shows an increment in both CGG as well as drain current with decreasing TOX of the device. Likewise, considering the case of EOEC, the drain current of the EOCC structure shows dominating parameter when compared to the CGG of the device followed by a reduction in τm as depicted in Fig. 16b. The enlarged view under Fig. 16b shows the reduction of optimized gate delay from ~ 8.5 to ~ 6.3 psec when compared to its circular structure. Finally, for the COEC structure, variation in channel MD causes virtual variation in TOX which in turn reduces CGG and drain current of the device. Likewise, the drain current shows dominating nature when compared to CGG which further led to an increase in τm of the device as depicted in Fig. 16c. The insertion in Fig. 16c shows the optimized COEC gate delay from ~ 8.5 to ~ 10 psec when compared to the COCC structure.

Benchmarking of proposed EOEC and COCC structures against several GAA MOSFETs and TFETs reports

This section presents a status map of numerous reports available on elliptical or circular GAA MOSFETs and TFETs against the analyzed EOEC and COCC structures as tabulated in Table 2. The status map shows that there is a drastic decrease in ION for GAA TFET when compared to its GAA MOSFET counterpart. This drastic decrease is due to different carrier injection mechanisms as well as comparatively lower VDS applied to the device. On the other hand, the device under consideration also has a lower ION when compared to its TFET counterpart. This lower ION is due to the lower VDS applied to the device under consideration. However, the device shows significantly reduced gate capacitance, which is the crucial parameter for any device’s performance. Besides the lower ION, the device under consideration shows a drastic increase in the ION/IOFF ratio. This high ION/IOFF determines the switching speed of the device.”

Table 2 Details of proposed EOEC and COCC structures against several GAA MOSFETs and TFETs.

Conclusion

The optimization of the structure and the comparative analysis of analog and RF FOMs for the III-V eGAA TFET with its EOEC, EOCC, and COEC structures are performed. Further optimization and comparative analysis of the analog and RF FOMs of the device is done for various EOEC, EOCC, and COEC structures. The device shows comparatively higher mobility, lower parasitic capacitances, and a minimum possible gate delay of the optimized geometries. When compared to their circular structures, the optimised EOEC and EOCC devices show a 64% and 59% increase in intrinsic gain, respectively, and a 28% and 23% increase in cut-off frequency. On the other hand, the optimized COEC structure shows a reduction in its intrinsic gain and its cut-off frequency. However, the physics behind the extracted output characteristics depends upon the dominancy of TOX and MD of the COEC structure. Owing to the higher gain and cut-off frequency of the EOEC and EOCC structures, the proposed optimized device can be an alternative structure that meets the need for higher performance for analog RF CMOS circuit applications. To summarize, the advantage of the undesired fabrication issue up to a certain level (MD up to 30% for EOEC, TOX up to 30% for EOCC, and MD up to 10% for COEC structures), which enhances the analog and RF FOMs, can be passed on to the circuit design where high gain, higher drive current, and higher gain bandwidth product are desired, and thus, the unintentional elliptical device structure proves to be the promising choice for analog/RF application.