contact resistances in trigate and finfet devices in …contact resistances in trigate and finfet...

16
Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach eo Bourdet, 1, 2 Jing Li, 1, 2 Johan Pelloux-Prayer, 3, 2 Fran¸coisTriozon, 3, 2 Mika¨ el Cass´ e, 3, 2 Sylvain Barraud, 3, 2 ebastien Martinie, 3, 2 Denis Rideau, 4 and Yann-Michel Niquet 1,2, a) 1) CEA, INAC-MEM, L Sim, Grenoble, France 2) Univ. Grenoble Alpes, Grenoble, France 3) CEA, LETI-MINATEC, Grenoble, France 4) STMicroelectronics, Crolles, France We compute the contact resistances R c in trigate and FinFET devices with widths and heights in the 4 to 24 nm range using a Non-Equilibrium Green’s Functions approach. Electron-phonon, surface roughness and Coulomb scattering are taken into account. We show that R c represents a significant part of the total resistance of devices with sub-30 nm gate lengths. The analysis of the quasi-Fermi level profile reveals that the spacers between the heavily doped source/drain and the gate are major contributors to the contact resistance. The conductance is indeed limited by the poor electrostatic control over the carrier density under the spacers. We then disentangle the ballistic and diffusive components of R c , and analyze the impact of different design parameters (cross section and doping profile in the contacts) on the electrical performances of the devices. The contact resistance and variability rapidly increase when the cross sectional area of the channel goes below 50 nm 2 . We also highlight the role of the charges trapped at the interface between silicon and the spacer material. I. INTRODUCTION The electrical performances of field-effect transistors are increasingly limited by the contact resistances R c as the gate length L decreases. 1,2 At low drain bias V ds , the “apparent” contact resistance can be defined as the linear extrapolation to zero gate length of the total re- sistance R(L) of the device. It embeds, therefore, all contributions that are independent of L, namely: i) the resistance of the metal-semiconductor contact, ii) the re- sistance of the highly doped source/drain and of the lowly doped spacers between the contacts and the gate, and iii) the so-called “ballistic” resistance of the device (residual resistance in the absence of scattering mechanisms). Al- though the latter can be intrinsic to the channel, it is usually considered part of the contact resistance as it is independent on the gate length. There has been a lot of theoretical and experimental works aimed at understanding and improving the trans- port through the channel of the transistor. At low field, the resistance of the channel can be characterized by the carrier mobility μ, which has been extensively measured and calculated in a variety of silicon structures (bulk, 3 films 4–12 and wires 13–27 ). There is much less literature on the contact resistances, 28–39 even though they have be- come a major bottleneck for device performances. Most models for the contact resistances are based on drift- diffusion equations needing carrier mobilities as input, which might not be well characterized in the very inho- mogeneous source/drain extensions. In this work, we compute the contact resistances R c of n-type, [110] oriented nanowire transistors 40 with widths W and heights H in the 4 to 24 nm range. We use a a) Electronic mail: [email protected] Non-Equilibrium Green’s Functions (NEGF) approach, which explicitly accounts for confinement and scatter- ing by phonons, surface roughness, and dopants in the inhomogeneous source and drain. 41 We consider square (W = H) and rectangular (W>H) trigate devices as well as FinFET (H>W ) devices with realistic, bulk-like source and drain contacts. We focus on components ii) and iii) of the contact resistance (source/drain extensions and ballistic resistances), and show that they indeed rep- resent a significant fraction of the total resistance of these devices. The methodology, based on the R(L) data ex- trapolation, is a numerical implementation of the trans- mission line approach 10,42 (section II). It can be supple- mented with a quasi-Fermi level analysis (section III), which highlights where the potential drops are located in the device, and helps to bridge NEGF with Technology Computer Aided Design (TCAD) tools based on drift- diffusion models. The quasi-Fermi level profile shows, in particular, that the contact resistance is dominated by the lowly doped spacers between the source/drain and the channel. We then disentangle the ballistic contribu- tion from the scattering by phonons, surface roughness and impurities (section IV). Simple models for the differ- ent terms confirm that the contact conductance is most often limited by the (poor) electrostatic control over the carrier density in the spacers. We next discuss the im- pact of some design parameters (channel cross section and doping profile) on the carrier mobility in the chan- nel, contact resistance and device variability (section V). We also investigate the effect of charges trapped at the interface between the silicon wire and the spacer material (Si 3 N 4 ). We finally provide experimental support for the main conclusions of this work (section VI). arXiv:1602.07545v1 [cond-mat.mes-hall] 24 Feb 2016

Upload: others

Post on 09-Sep-2020

14 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

Contact resistances in trigate and FinFET devices in a Non-EquilibriumGreen’s Functions approach

Leo Bourdet,1, 2 Jing Li,1, 2 Johan Pelloux-Prayer,3, 2 Francois Triozon,3, 2 Mikael Casse,3, 2 Sylvain Barraud,3, 2

Sebastien Martinie,3, 2 Denis Rideau,4 and Yann-Michel Niquet1, 2, a)

1)CEA, INAC-MEM, L Sim, Grenoble, France2)Univ. Grenoble Alpes, Grenoble, France3)CEA, LETI-MINATEC, Grenoble, France4)STMicroelectronics, Crolles, France

We compute the contact resistances Rc in trigate and FinFET devices with widths and heights in the 4to 24 nm range using a Non-Equilibrium Green’s Functions approach. Electron-phonon, surface roughnessand Coulomb scattering are taken into account. We show that Rc represents a significant part of the totalresistance of devices with sub-30 nm gate lengths. The analysis of the quasi-Fermi level profile reveals that thespacers between the heavily doped source/drain and the gate are major contributors to the contact resistance.The conductance is indeed limited by the poor electrostatic control over the carrier density under the spacers.We then disentangle the ballistic and diffusive components of Rc, and analyze the impact of different designparameters (cross section and doping profile in the contacts) on the electrical performances of the devices.The contact resistance and variability rapidly increase when the cross sectional area of the channel goes below' 50 nm2. We also highlight the role of the charges trapped at the interface between silicon and the spacermaterial.

I. INTRODUCTION

The electrical performances of field-effect transistorsare increasingly limited by the contact resistances Rc asthe gate length L decreases.1,2 At low drain bias Vds,the “apparent” contact resistance can be defined as thelinear extrapolation to zero gate length of the total re-sistance R(L) of the device. It embeds, therefore, allcontributions that are independent of L, namely: i) theresistance of the metal-semiconductor contact, ii) the re-sistance of the highly doped source/drain and of the lowlydoped spacers between the contacts and the gate, and iii)the so-called “ballistic” resistance of the device (residualresistance in the absence of scattering mechanisms). Al-though the latter can be intrinsic to the channel, it isusually considered part of the contact resistance as it isindependent on the gate length.

There has been a lot of theoretical and experimentalworks aimed at understanding and improving the trans-port through the channel of the transistor. At low field,the resistance of the channel can be characterized by thecarrier mobility µ, which has been extensively measuredand calculated in a variety of silicon structures (bulk,3

films4–12 and wires13–27). There is much less literature onthe contact resistances,28–39 even though they have be-come a major bottleneck for device performances. Mostmodels for the contact resistances are based on drift-diffusion equations needing carrier mobilities as input,which might not be well characterized in the very inho-mogeneous source/drain extensions.

In this work, we compute the contact resistances Rc ofn-type, [110] oriented nanowire transistors40 with widthsW and heights H in the 4 to 24 nm range. We use a

a)Electronic mail: [email protected]

Non-Equilibrium Green’s Functions (NEGF) approach,which explicitly accounts for confinement and scatter-ing by phonons, surface roughness, and dopants in theinhomogeneous source and drain.41 We consider square(W = H) and rectangular (W > H) trigate devices aswell as FinFET (H > W ) devices with realistic, bulk-likesource and drain contacts. We focus on components ii)and iii) of the contact resistance (source/drain extensionsand ballistic resistances), and show that they indeed rep-resent a significant fraction of the total resistance of thesedevices. The methodology, based on the R(L) data ex-trapolation, is a numerical implementation of the trans-mission line approach10,42 (section II). It can be supple-mented with a quasi-Fermi level analysis (section III),which highlights where the potential drops are located inthe device, and helps to bridge NEGF with TechnologyComputer Aided Design (TCAD) tools based on drift-diffusion models. The quasi-Fermi level profile shows, inparticular, that the contact resistance is dominated bythe lowly doped spacers between the source/drain andthe channel. We then disentangle the ballistic contribu-tion from the scattering by phonons, surface roughnessand impurities (section IV). Simple models for the differ-ent terms confirm that the contact conductance is mostoften limited by the (poor) electrostatic control over thecarrier density in the spacers. We next discuss the im-pact of some design parameters (channel cross sectionand doping profile) on the carrier mobility in the chan-nel, contact resistance and device variability (section V).We also investigate the effect of charges trapped at theinterface between the silicon wire and the spacer material(Si3N4). We finally provide experimental support for themain conclusions of this work (section VI).

arX

iv:1

602.

0754

5v1

[co

nd-m

at.m

es-h

all]

24

Feb

2016

Page 2: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

2

FIG. 1. A typical n-type trigate device with width W = 10nm, height H = 10 nm, and gate length L = 30 nm. Siliconis in red, SiO2 in green, HfO2 in blue, and the gate in gray(Si3N4 otherwise). The white dots in the source and draincontacts are donor impurities. The silicon substrate is notrepresented.

II. METHODOLOGY AND DEVICES

In this section, we discuss the devices, the NEGF ap-proach, the numerical transmission line method for thecontact resistance, and a specific example.

A. Devices

A typical device is represented in Fig. 1. The channelis a z = [110] oriented rectangular nanowire with widthW and height H, etched in a (001) Silicon-On-Insulator(SOI) layer.26,27 It is lying on a 25 nm thick buried oxide(BOX) and a n-doped Si substrate (donor concentrationNd = 1018 cm−3). The gate stack covers the top (001)and side (110) facets of the nanowire. It is made of a 0.8nm thick layer of SiO2 (dielectric constant ε = 3.9) anda 2.2 nm thick layer of HfO2 (ε = 20). The source anddrain are modeled as films with thickness tSD ≥ H + 7nm. Periodic boundary conditions are, therefore, appliedalong y = [110], on both potential and wave functions.43

The width of the unit cell is WSD ≥ W + 14 nm. Thelength of the source and drain films included in the sim-ulation domain is LSD = 16 nm. The source and drainare separated from the gate by Si3N4 spacers with lengthLsp = 6 nm (ε = 7.5).

Surface roughness (SR) and charge disorders are ex-plicitly included in the geometries. For SR, we use aGaussian auto-correlation function model44 with differ-ent correlation lengths Λ and rms ∆ on different facets:Λ = 1.5 nm and ∆ = 0.25 nm on the bottom inter-face with the BOX, Λ = 1.5 nm and ∆ = 0.35 nm onthe top (001) facet, and Λ = 2.0 nm, ∆ = 0.45 nm onthe side (110) facets. We also include Remote Coulomb

−20 −15 −10 −5 0 5 10

z (nm)

1016

1017

1018

1019

1020

1021

Nd(cm

−3)

ChannelSpacerSource

Reference 1.5× 1020 cm−3

Reference 3.0× 1020 cm−3

Small underlap

Small overlap

Strong underlap

FIG. 2. Target doping profiles on the source side (the profilesare symmetric on the drain side).

Scattering (RCS) in the gate stack as a distribution ofpositive charges at the SiO2/HfO2 interface with densitynRCS = 2 × 1013 cm−2. The SR parameters for the topand bottom interfaces are chosen to reproduce the ex-perimental mobility in planar FDSOI devices (along thelines of Ref. 11). The SR parameters of the side facetsas well as nRCS are chosen to reproduce the experimen-tal mobility in W = 10×H = 10 nm trigate devices (seesection III). The sidewalls appear rougher than the topand bottom interfaces because they are etched in the SOIfilm.

The source, drain and spacers are n-doped with pointcharges as models for ionized donor impurities. Thesepoint charges are randomly scattered using the dopingprofiles shown in Fig. 2 as target distribution functions.The “Reference” profiles have Nd = 1.5 × 1020 cm−3 orNd = 3 × 1020 cm−3 in the contacts, and a single decaylength λ1 = 3 nm/decade under the spacers and chan-nel. The other profiles will be discussed in detail in sec-tion V B. Near the edges of the simulation box the pointcharge distributions are mixed with the target continu-ous background charge distributions in order to smooththe variations of the potential at the boundaries wherethe contact self-energies41 are plugged in.45

B. The NEGF approach

The current is computed in a self-consistent NEGFframework,41,46 within the effective mass approximation(EMA).47 The NEGF equations are solved on a finitedifferences grid (with 2 A step), in a fully coupled modespace approach (160 to 352 modes depending on the crosssection). Electron-phonon scattering is described by anacoustic deformation potential Dac = 14.6 eV,5 and bythe three f and three g inter-valley processes of Ref. 3.Technical details about the electron-phonon self-energies

Page 3: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

3

and solution of the NEGF equations can be found inthe Appendix of Ref. 10. The n-doped substrate is bi-ased at back-gate voltage Vbg = 0 V and is treated semi-classically.48

Electron-electron interactions are treated in a self-consistent “Schrodinger-Poisson” approximation, i.e. thecarriers are moving in the potential created by the aver-age density of conduction band electrons in silicon. Theeffects of valence band electrons in silicon and other ma-terials are accounted for by the dielectric constants in-troduced in Poisson’s equation. This approximation ne-glects Coulomb correlations that might, however, becomeimportant in nanoscale devices.49 As a step further, wehave introduced (in a few test cases) a GW -like self-energy accounting for the long-range correlations broughtby the dielectric mismatch between the different mate-rials, supplemented with a local density approximationfor the short-range exchange-correlation effects.8,12,50,51

This has a moderate impact on the contact resistances(5 − 10% increase). Since the calculation of this GW -like self-energy is very demanding, these corrections havebeen neglected in the following.

One of the main advantages of NEGF is that allstructural scattering mechanisms (SR, impurity and RCSscattering) are treated explicitly. There is no need formodels for the interactions with these disorders. Asa matter of fact, such models are still missing un-der the spacers, which are very inhomogeneous butcritical regions on the resistive path (see sections IIIand IV). Also, at variance with most previous NEGFcalculations,16,17,22,23,25,39,52,53 the source and drain con-tacts are wide enough with respect to the channel to actas bulk reservoirs (with a 3D-like density of states). Thisis essential for a quantitative description of the contactresistances. The convergence with respect to tSD, WSD

and LSD has, in particular, been carefully checked.

C. Methodology

At low drain bias Vds, the total resistance of the deviceis expected to increase linearly with gate length L:54

R(L) =Vds

Ids= R0 +Rs +Rd +

L

n1dµe, (1)

where Ids is the drain current, Rs is the diffusive resis-tance of the source, Rd is the diffusive resistance of thedrain, µ is the carrier mobility and n1d the carrier den-sity per unit length in the channel, and R0 is the bal-listic resistance of the device (residual resistance in theabsence of scattering mechanisms).54,55 At zero tempera-ture, the ballistic conductance of a homogeneous conduc-tor, G0 = R−1

0 = (12.9 kΩ/N)−1, is limited by the num-ber N of 1D sub-bands carrying current. At room tem-perature, G0 reads, assuming Maxwell-Boltzmann statis-tics and a single transport mass m∗ for all sub-bands:25,54

G0 =n1de

2

√2πm∗kT

. (2)

It is, therefore, proportional to the carrier density in theconductor – we will come back to that point in sectionIV.

The “apparent” contact resistance Rc = R0 +Rs +Rd

can hence be extracted from the linear extrapolation ofR(L) down to L = 0 (an approach known as the trans-mission line method42). For that purpose, we prepare aseries of devices with lengths L = 30 nm and L = 60nm (and up to L = 90 nm), sharing the same con-tacts (source/drain/spacers geometry and dopant distri-butions). In order to limit the noise on the R(L) datathat might arise from different disorders at different L’s,we repeat the 30 nm long sample of surface roughness andRCS charges along the 60 and 90 nm long channels.10 Wecan extract that way the resistance of the chosen contactgeometry and dopant distribution very accurately. Wethen average the contact resistance over 3 to 8 surfaceroughness profiles and dopant distributions in order toassess local variability.

In order to ease the comparison between devices withdifferent cross sections, the carrier densities n2d =n1d/Weff and the resistances R = RWeff are normalizedto the total effective width of the channel Weff = W+2H.The drain bias is Vds = 10 mV, and the threshold voltageVt is extracted from the Y = Ids/

√gm function (where

gm = dIds/dVgs is the transconductance and Vgs the gatebias).56 There are no significant short-channel effects (inparticular, no Vt roll-off) in any of the investigated de-vices.

D. Example: The W = 10×H = 10 nm trigate

In this paragraph, we illustrate the above methodologyon a specific example: The W = 10 × H = 10 nm tri-gate device with the “Reference 1.5×1020 cm−3” dopingprofile of Fig. 2.

The total resistance of a series of devices is plotted as afunction of the channel length in Fig. 3, at gate overdriveVgt = Vgs − Vt = 0.69 V. As expected, the resistancescales linearly with length in the inversion regime. Thecontact resistance Rc (red star on Fig. 3) and the channelmobility µ can therefore be extracted from these data.Rc is plotted as a function of the gate overdrive Vgt andcarrier density 〈n2d〉ch in the channel in Fig. 4. 〈n2d〉chis defined as the average carrier density in the central, 30nm long segment of the 60 nm long channel. Rc follows an

approximate 〈n2d〉−βch (or V −βgt ) trend (with β close to 1).As the conductivity is proportional to the carrier densityin the simplest Drude model, this trend suggests thatthe contact resistance is dominated by the near channel(spacers) region.57 This will be confirmed by a quasi-Fermi level analysis in the next section.

Data for five devices with different disorders are repre-sented in Fig. 4. The local variability due to fluctuationsin surface roughness and dopant distributions remainsweak in 10 × 10 nm devices. Rc is compared to the to-tal resistance R(L = 30 nm) of the 30 nm long devices,

Page 4: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

4

0 20 40 60 80 100

L (nm)

150

200

250

300

350

400

450

500

550R

(Ω.µm)

Vds = 10 mV

Vgt = 0.690 V

FIG. 3. Resistance R(L) as a function of gate length L fora W = 10 × H = 10 nm trigate device with the “Reference1.5×1020 cm−3” doping profile of Fig. 2. The gate voltage isVgt = Vgs−Vt = 0.69 V, which corresponds to a carrier density〈n2d〉ch ' 9×1012 cm−2 in the channel. The shaded gray areais the 90% confidence interval for the linear regression (dottedred line), and the red star is the extrapolated Rc ≡ R(L = 0).

1012 1013

〈n2d〉ch (cm−2)

101

102

103

R(Ω

.µm)

Rc

R(L = 30 nm)

R0

0.13 0.41 0.69 0.97 1.25Vgt (V)

FIG. 4. Contact resistance Rc as a function of the gate over-drive (upper axis)/carrier density (lower axis), in a W =10×H = 10 nm trigate device with the “Reference 1.5×1020

cm−3” doping profile of Fig. 2. Data for five devices (dottedlines, each corresponding to different disorders in the channeland contacts), and their average (solid line) are represented.Rc is compared to the total resistance R(L = 30 nm) of the 30nm long devices, and to the resistance R0 of the correspondingballistic device.

and to the the resistance R0 of the corresponding “bal-listic” device (neither phonons, nor SR and RCS, witha continuous background dopant distribution in the con-tacts). The contact resistance is around half the totalresistance in a 30 nm long device – it would hence be aslarge as two-thirds the total resistance in a ∼ 15 nm long

channel. It is, moreover, much larger than the ballis-tic resistance (especially in the strong inversion regime),which is a component and lower bound of Rc. The opti-mization of the source/drain/spacers region is, therefore,at least as important as the design of the channel in thesetransistors.

III. A QUASI-FERMI LEVEL ANALYSIS OF THERESISTIVE PATH

The above conclusions can be further supported bya quasi-Fermi level analysis, which highlights where thepotential drops in the system. We first define the quasi-Fermi level in the NEGF framework, then discuss limit-ing cases (ballistic and diffusive conductors), and finallymake a detailed analysis on the trigate device of para-graph II D.

A. Definition

Drift-diffusion models48 assume that the carriers arein local equilibrium with a quasi-Fermi level εf (r). Thecurrent density in the device is then proportional to thegradient of εf (r):

j = nµ∇rεf . (3)

In the NEGF framework, the carriers can be driven farout-of-equilibrium, so that the quasi-Fermi level is, ingeneral, ill-defined. The distribution of electrons can in-stead be characterized by the local distribution function:

f(r, E) = Docc(r, E)/D(r, E) , (4)

where Docc(r, E) = ImG<(r, r, E)/(2π) is the density ofoccupied states (with G< the lesser Green’s function),and D(r, E) = −ImGr(r, r, E)/π is the total density ofstates (with Gr the retarded Green’s function).41 Yet adetailed analysis shows that f(r, E) remains close to aFermi function at low bias. In that limit, we can definethe quasi-Fermi level εf (r) as the chemical potential thatreproduces the NEGF carrier density n(r) assuming localFermi-Dirac statistics:

n(r) =

∫dE D(r, E) f [(E − εf (r)) /kT ] , (5)

where f(x) = 1/(1 + ex) is the reduced Fermi function.The quasi-Fermi level analysis is an efficient way to

bridge NEGF with the drift-diffusion models widely usedin industrial TCAD tools. For practical purposes, wedefine a one-dimensional quasi-Fermi level εf (z) in thedevice from the relation:

n1d(z) =

∫dE D1d(z, E) f [(E − εf (z)) /kT ] , (6)

where n1d(z) and D1d(z, E) are the NEGF carrier den-sity and density of states per unit length, which are theintegrals of n(r) and D(r, E) in the cross section at z.

Page 5: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

5

B. Limiting cases: Ballistic and diffusive conductors

In this paragraph, we address two limiting cases: theballistic and diffusive conductors. We discuss how thequasi-Fermi level drops in these conductors, in order toprovide guidelines for the analysis of trigate and FinFETdevices.

Although the carrier distribution in a ballistic conduc-tor is definitely not a Fermi-Dirac distribution,58 the so-lution of Eq. (6) remains unambiguously defined. Let usconsider a ballistic transistor in the ON state, at low Vds

and low temperature. The potential in such a transistorcan be modeled in a first approximation as a square bar-rier. Carriers with positive group velocities (labeled aswave vectors kz > 0 for simplicity) are injected from thesource and are in equilibrium with the chemical potentialµs. Carriers with negative group velocities (kz < 0) areinjected from the drain and are in equilibrium with thechemical potential µd = µs − eVds. If the carrier den-sity is much lower in the channel than in the source anddrain (which is usually the case), most of the electronsare backscattered by the barrier, and only a small frac-tion flows in the channel. The quasi-Fermi level hencetends to µs in the source, and to µd in the drain. In thechannel, there is a population of carriers with positivegroup velocities flowing from the source and in equilib-rium with µs, and a population of carriers with negativegroup velocities flowing from the drain and in equilibriumwith µd. Hence, all ±kz states are filled below µd, whileonly the kz > 0 states are occupied between µd and µs.If the density of states is approximately constant in the[µd, µs] range, the quasi-Fermi level in the channel mustbe εf = (µs +µd)/2 in order to match the carrier densitywith a Fermi-Dirac distribution.

This argument shows that the quasi-Fermi level essen-tially drops at both ends of the channel in a ballisticdevice.59,60 Ballistic device simulations confirm that thisis indeed the case. The width of the drop on each end ofthe channel depends, in particular, on the shape of thebarrier and on the temperature. The drop is typicallyfaster in the strong than in the weak inversion regime,where the quasi-Fermi level can show significant varia-tions over the whole channel in short devices.

In a diffusive conductor on the other hand, the cur-rent must be sustained by a finite electric field. Thequasi-Fermi level shall show the same variations as thepotential if the density is approximately constant, i.e. itshall drop linearly in the channel at low bias.

As a concluding remark, we would like to emphasizethat the above analysis only applies at low bias, andmight break down at large Vds in both ballistic58 anddiffusive cases.

C. Application to the W = 10×H = 10 nm trigate

The quasi-Fermi level computed in the same device asin paragraph II D is plotted in Fig. 5 at different gate

−40 −30 −20 −10 0 10 20 30 40

z (nm)

−10

−8

−6

−4

−2

0

ε f(m

eV)

Source Channel Drain

Vds = 10 mV

Vgt = −0.01 V

Vgt = 0.27 V

Vgt = 0.69 V

Vgt = 1.25 V

FIG. 5. Quasi-Fermi level in a 30 nm long, W = 10×H = 10nm trigate device with the “Reference 1.5×1020 cm−3” dopingprofile of Fig. 2, at different gate overdrives. The drain biasis Vds = 10 mV. The shaded gray area delimits the segmentwhere the slope dεf/dz of the quasi-Fermi level is computed.

overdrives (drain bias Vds = 10 mV).

As expected from the above discussion, the quasi-Fermilevel tends to zero in the source (reference chemical po-tential µs = 0), and to µd = −eVds in the drain. More-over, the quasi-Fermi level is almost constant in thesource and drain, and drops under the spacers and in thechannel. These regions are, therefore, the most resistiveparts of the device. The resistance of the heavily doped,bulk source and drain is negligible because the densityof carriers and cross sectional area are much larger therethan in the channel, even though the mobility is expectedto be very low in the contacts (< 100 cm2/V/s).61

The quasi-Fermi level profile is, nonetheless, stronglydependent on the gate bias. Near or under the thresh-old, the transport is strongly limited by the bell-shapedbarrier under the gate. The density is, moreover, veryinhomogeneous in the channel (see Fig. 6). The quasi-Fermi level drops unevenly in the channel, and no cleardistinction can be made between a ballistic and a diffu-sive (scattering-limited) component.

The picture is clearer at moderate to strong inversion.The carrier density is much more homogeneous under thegate. The quasi-Fermi level drops quasi-linearly in thechannel, and exhibits two steps under the spacers. Wecan conjecture from paragraph III B that i) the transportin the channel is diffusive (the carrier mobility µ shallhence be proportional to slope of the quasi-Fermi levelunder the gate) ; ii) the drop of the quasi-Fermi levelunder the spacers is proportional to the contact resistanceRc (including the ballistic component R0).

To support these conclusions, we have recomputed thechannel mobility µ and contact resistance Rc from the

Page 6: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

6

−40 −30 −20 −10 0 10 20 30 40

z (nm)

1010

1011

1012

1013

1014

n2d

(cm

−2)

Source Channel Drain

(a)

Vgt = −0.01 V

Vgt = 0.27 V

Vgt = 0.69 V

Vgt = 1.25 V

FIG. 6. (a) Carrier density n2d in a 30 nm long, W = 10×H =10 nm trigate device with the “Reference 1.5 × 1020 cm−3”doping profile of Fig. 2, at different gate overdrives. (b, c) 3Dcarrier density in a longitudinal and transverse cross sectionthrough the center of the channel (Vgt = 1.25 V). The strongdisorder in the contacts is due to dopants.

relations:

1

〈n1d〉chµe=

1

eIds

⟨dεfdz

⟩ch

(7a)

Rc =Vds

Ids− L

〈n1d〉chµe, (7b)

where 〈dεf/dz〉ch is the slope of the quasi-Fermi level ex-tracted from a linear regression in the inner z ∈ [−10, 10]nm range of the 30 nm long channel.

The mobility and contact resistance obtained that wayare plotted in Fig. 7. They are in very good agree-ment with those extracted with the transmission linemethodology of paragraph II C in the strong inversionregime, where the quasi-Fermi level drops quasi-linearlyunder the gate. The contact resistance is, therefore, in-deed dominated by the spacers.57 The mobility also com-pares well with that extracted from the R(L) data on“homogeneous” nanowire channels without spacers andbulk source/drain contacts (the gate then runs across thewhole simulation box – see Ref. 10). The differencesat low 〈n2d〉ch are due to the source/channel and chan-nel/drain junctions, whose finite depletion widths result

1012 1013

〈n2d〉ch (cm−2)

102

Rc(Ω

.µm)

(a)

R(L) [Eq. (1)]

dεf/dz [Eqs. (7)]

0.13 0.41 0.69 0.97 1.25Vgt (V)

1012 1013

〈n2d〉ch (cm−2)

50

100

150

200

250

300

µ(cm

2/V

/s)

(b)

R(L) [Eq. (1)]

dεf/dz [Eqs. (7)]

Homogeneous channel [Ref. 10]

Experimental [Ref. 27]

FIG. 7. (a) Contact resistance Rc and (b) carrier mobility µ inthe channel of a W = 10×H = 10 nm trigate device with the“Reference 1.5×1020 cm−3” doping profile of Fig. 2, extractedfrom: i) a linear regression on the R(L) data [Eq. (1)], and ii)the slope dεf/dz of the quasi-Fermi level under the gate of the30 nm long device [Eqs. (7)]. The data are averaged over fivedevices as in Fig. 4. The mobility computed in homogeneousnanowire channels without spacers and S/D contacts is alsoplotted (see Ref. 10) along with experimental data (Ref. 27).

in significant variations of the carrier density near bothends of the channel (see Fig. 6). As a matter of fact, ho-mogeneous nanowire channels without such junctions arebetter suited to the calculation of the long-channel mo-bility at low carrier density.10 The calculated mobilitiesare close to the experimental data of Ref. 27 whateverthe methodology.

The analysis of the quasi-Fermi level provides an alter-native to the R(L) methodology of paragraph II C (fasterbut often less accurate at low densities). Although Fig.5 is consistent with a drift-diffusion model using differentmobilities in the channel, spacers and bulk source/drain,we point out that Eq. (3) is not able to capture theballistic component of the contact resistance on physi-cal grounds. R0 might indeed be lumped in the mobility

Page 7: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

7

under the spacers, but that mobility would hardly betransferable to other device designs.

In the next section, we discuss the different (ballisticand diffusive) contributions to the contact resistance.

IV. ANALYSIS OF THE CONTACT RESISTANCE ANDMODELS

In this section, we break down the contact resistanceinto ballistic and diffusive contributions and discuss sim-ple models for the NEGF data.

A. Breakdown of the contact resistance into ballistic anddiffusive contributions

In order to disentangle the main contributions to thecontact resistance, we have computed the following seriesof W = 10×H = 10 nm devices:

1. The “Ballistic” device without scattering (neitherphonons nor SR, RCS and impurities). In order topreserve the electrostatics, the point-like dopantsin the source and drain are replaced by the back-ground distribution of Fig. 2, and the RCS chargesby a uniform background at the SiO2/HfO2 inter-face. As expected, the resistance R0 of this deviceis independent on the device length in the L ≥ 30nm range. It is plotted in Fig. 4.

2. Same as 1, with phonons enabled (PH).

3. Same as 2, with surface roughness and RCS enabled(PH+SR+RCS).

4. Same as 3, with Coulomb scattering enabled in thesource and drain (PH+SR+RCS+IMP, i.e. full de-vice with point-like dopants).

The different contact resistances are plotted as a func-tion of 〈n2d〉ch in Fig. 8a, while the contributions fromphonons (PH), surface roughness and remote Coulomb(SR+RCS), and impurity (IMP) scattering, defined asthe differences between the contact resistance of two suc-cessive devices, are plotted in Fig. 8b. As before, thedata were averaged over different realizations of the dis-orders.

The ballistic resistance is the dominant contributionto Rc, but represents only 40% to 50% of the total con-tact resistance. The main diffusive component is impu-rity scattering, followed by phonons, then SR+RCS scat-tering. As expected, the ratio impurity/phonon scatter-ing decreases with increasing carrier density due to thescreening of the ionized impurities by the electron gas un-der the spacers and gate. The SR+RCS contribution isactually dominated by RCS scattering near both ends ofthe channel at low densities, and by SR scattering underthe spacers at large densities. It is, however, difficult to

1012 1013

〈n2d〉ch (cm−2)

102

Rc(Ω

.µm)

RNW0

Ballistic device

PH

PH+SR+RCS

PH+SR+RCS+IMP

0.13 0.41 0.69 0.97 1.25Vgt (V)

(a)

1012 1013

〈n2d〉ch (cm−2)

101

102

Rc(Ω

.µm)

Ballistic

PH

SR+RCS

IMP

0.13 0.41 0.69 0.97 1.25Vgt (V)

(b)

FIG. 8. (a) Contact resistance Rc as a function of the carrierdensity in a W = 10 × H = 10 nm trigate device with the“Reference 1.5 × 1020 cm−3” doping profile of Fig. 2. Dif-ferent scattering mechanisms have been included (see text).(b) The ballistic, phonon (PH), surface roughness (SR+RCS),and impurity (IMP) scattering contributions to the contact re-sistance, obtained as differences between the data of Fig. 8a.The dashed lines are simple models for the ballistic resistanceand phonons contribution (see text).

disentangle these two smaller contributions. Yet SR scat-tering clearly makes a minor contribution to Rc, while itis a dominant mechanism in the channel. Actually, mostof the contacts undergo volume accumulation rather thansurface inversion, so that the carriers do not probe muchthe surface of the spacers. The distribution of carriersunder the spacers will be discussed in more detail in thenext paragraph.

B. Models and discussion

In order to get a deeper understanding of Fig. 8, wehave computed the ballistic resistance RNW

0 (n2d) of a ho-

Page 8: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

8

mogeneous (infinitely long) 10×10 nm nanowire channel(dotted line in Fig. 8a).62 RNW

0 (n2d) is significantly lowerthan the ballistic resistance R0 of the whole device. Thecross section and carrier density are, actually, very inho-mogeneous in the source/drain and under the spacers. Inthe simplest “top of the barrier” model,63,64 the ballisticresistance of such an inhomogeneous device shall be theballistic resistance of the homogeneous conductor builtfrom the cross section at the top of the 1D conductionband edge profile Ec(z). At low bias, the top of the bar-rier is very close to the point z = zmin where the carrierdensity n2d(z) ≡ nmin

2d reaches its minimum. We have,therefore, plotted RNW

0 (nmin2d ) as a dashed black line on

Fig. 8b. This simple estimate matches the resistance ofthe ballistic device much better. The point z = zmin isunder the gate in the subthreshold regime (Fig. 6), butmoves under the spacers in the ON regime. The ballis-tic resistance is then more dependent on the design ofthe spacers than on the design of the channel, and trulyappears in this respect as a “contact” resistance.

We have also attempted to recompute the phonon con-tribution from a simple local mobility model:

RPH =

∫device

dz

n2d(z)µPH(z)e− L

〈n2d〉chµPHch e

, (8)

where n2d(z) is the carrier density, µPH(z) the phonon-limited mobility at z, and µPH

ch the inversion layer mo-bility in the channel. The first term is resistance of thewhole device, while the second term is the resistance of anideal channel with homogeneous carrier density 〈n2d〉ch.We can further approximate this expression as follows:

RPH ∼1

µPHch e

[∫device

dz

n2d(z)− L

〈n2d〉ch

]∼ Lref⟨

∆n−12d

⟩−1µPH

ch e, (9)

where⟨∆n−1

2d

⟩=

1

Lref

[∫device

dz

n2d(z)− L

〈n2d〉ch

], (10)

and Lref is an (arbitrary) reference length. 〈∆n−12d 〉 is

the difference between the inverse carrier densities inte-grated in the whole device and in the ideal channel, nor-malized to Lref . We choose Lref = 2Lsp = 12 nm, so thatEq. (9) appears as an effective spacer resistance with

an effective spacer carrier density 〈n2d〉sp ≡⟨∆n−1

2d

⟩−1.

The above approximation on the local mobility is justi-fied by the fact that the main contributions to 〈∆n−1

2d 〉and RPH are collected under the spacers (where 1/n2d(z)reaches its maximum in the strong inversion regime), andnear the entrance/exit of the channel (where the differ-ences between n2d(z) and 〈n2d〉ch are the largest). Inthese regions, µPH(z) can be replaced with µPH

ch , as µPH

shows a rather weak dependence on the carrier density.This model, plotted as a dashed blue line in Fig. 8, re-produces the NEGF data very well. It also reproduces

0.0 0.2 0.4 0.6 0.8 1.0 1.2

Vgt (V)

0.0

0.5

1.0

1.5

2.0

n2d

(cm

−2)

×1013

(a)

〈n2d〉ch〈n2d〉sp ≡ 〈∆n−1

2d 〉−1

nmin2d

0.0 0.2 0.4 0.6 0.8 1.0 1.2

Vgt (V)

0.0

0.5

1.0

1.5

2.0

2.5

3.0

C(µF/cm

2)

(b)

〈n2d〉ch〈n2d〉sp ≡ 〈∆n−1

2d 〉−1

FIG. 9. (a) Average carrier density 〈n2d〉ch in the channel and〈n2d〉sp under spacers, and minimal density nmin

2d in the deviceas a function of the gate overdrive Vgt, in a W = 10×H = 10nm trigate device with the “Reference 1.5×1020 cm−3” dopingprofile of Fig. 2. The point where the density reaches itsminimum nmin

2d is under the spacers when Vgt > 0.4 V. (b)Gate-channel and gate spacers capacitances. The data wereaveraged over five different devices.

the main trends, but is not as accurate for the surfaceroughness and impurity contributions, whose mobilitiesshow stronger dependences on the carrier density. Thecarriers indeed move from a “bulk-like” distribution inthe source to “surface-inversion-like” distribution in thechannel (Fig. 6), so that neither the inversion layer northe bulk mobilities hold under the spacers. In particular,the SR-limited inversion layer mobility µSR

ch overestimatesthe surface roughness contribution to Rc. Indeed, n2d(z)can be pretty large near the source/spacer junction, butthe carriers are distributed in the bulk of the nanowire,and are, therefore much less scattered by surface rough-ness than expected from µSR

ch .

As discussed above, the conductivity of the contacts(as well as the ballistic conductance in the strong inver-sion regime) are essentially controlled by the density of

Page 9: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

9

carriers under the spacers. The latter can be adequately

characterized by 〈n2d〉sp ≡⟨∆n−1

2d

⟩−1. By definition

[Eq. (10)], the spacers are depleted (as regards trans-port) with respect to the channel when 〈n2d〉sp . 〈n2d〉ch.〈n2d〉ch and 〈n2d〉sp exhibit different dependences on thegate voltage (Figs. 6 and 9). The channel is indeedwell controlled by the gate, the gate to channel capac-itance Cch = d〈n2d〉ch/dVgt reaching a large fraction ofthe oxide capacitance Cox. On the contrary, the spac-ers are doped but are poorly controlled by the gate (be-cause they do not overlap). The gate to spacers ca-pacitance Csp = d〈n2d〉sp/dVgt is actually much lowerthan Cch (and weakly dependent on Vgt). The spacersdo not, therefore, accumulate much excess charge in theON regime. In the present device, 〈n2d〉ch and 〈n2d〉spcross at gate overdrive Vgt ' 0.5 V. Above that gatevoltage, the carrier density shows a significant dip underthe spacers (with density nmin

2d ) that makes a significantcontribution to the ballistic and diffusive resistances.

To assess charge control in different devices, we canfurther approximate 〈n2d〉ch and 〈n2d〉sp by:

〈n2d〉ch ' Cox

(Vg − Vt

)〈n2d〉sp ' Csp

(Vg − V ′t

), (11)

where Cox ≈ 2.7 µF/cm2 for all devices considered in

this work, and Vt and V ′t are threshold voltages for thechannel and spacer, respectively.65 There is, therefore, alinear relation between 〈n2d〉sp and 〈n2d〉ch:

〈n2d〉sp − n0 = α (〈n2d〉ch − n0) , (12)

where α = Csp/Cox < 1 and n0 = CspCox(Vt−V ′t )/(Cox−Csp) is the crossover density. α and n0 are characteristicof a design, but do not depend on the channel length (atleast in the absence of short-channel effects). As an ex-ample, α = 0.47 and n0 = 6.1× 1012 cm−2 in the 10× 10nm device of Fig. 9. Data for other devices will be givenin the next paragraphs. In the strong inversion regime,Rc is approximately proportional to 〈n2d〉−1

sp . It can

therefore be written Rc ∼ 2Lsp/(〈n2d〉spµspe), where µsp

is an “effective” spacer mobility (typically in the 55− 65cm2/V/s range). We would like to stress, though, thatµsp embeds the ballistic resistance and shall not, there-fore, be interpreted as a diffusive mobility transferable toother spacer lengths.

To conclude, we would like to remind that thepresent NEGF calculations miss the contribution fromthe metal/semiconductor contact resistance. The latteris expected to be weakly dependent on Vgt, and shalltherefore appear as a rigid shift R′c = R′cWeff , whereR′c is independent of Vgt and of Weff (as long as themetal/semiconductor contact remains the same whateverthe design of the channel).

V. IMPORTANCE OF THE DESIGN OF THE NEARSPACER REGION

As discussed above, the design of the spacers can have agreat impact on the contact resistance. We investigate inthis section the effect of the cross section of the nanowire(for a given doping profile), then the effect of the dopingprofile (at constant cross section), and finally the effectof charges trapped in the spacer material.

A. Influence of the cross section

The contact resistance and channel mobility are plot-ted as a function of the carrier density 〈n2d〉ch for trigateand FinFET devices with different cross sections W ×Hin Fig. 10. They have been extracted from linear re-gressions on the R(L) data on full devices and on ho-mogeneous channels,10 respectively. In order to improvethe readability of the figure, we have actually plottedRc × 〈n2d〉ch as a function of 〈n2d〉ch in Fig. 10a,c. Fig.10a,b display the results for “square” trigate devices withside W = H ranging from 4 to 10 nm, and Fig. 10c,dthe results for horizontal trigate and vertical FinFET de-vices with either W = 8 nm or H = 8 nm. The dopingprofile is “Reference 1.5× 1020 cm−3” from Fig. 2. Thestandard deviation on Rc (±1σ) computed on a set ofeight different devices is also reported for square trigatedevices as a measure of local variability.

The carrier mobility in the channel of square trigatesdecreases when the cross section is reduced due to the en-hancement of scattering by structural confinement. Thepicture is, however, very dependent on the carrier den-sity. At high density, inversion mostly takes place at thesurface of the nanowire (except in the 4× 4 nm device).Hence the current flows on the top and two side facets,and transport along each facet is mainly limited by SRscattering on that facet. The mobility is, therefore, littledependent on the cross section. The carriers, however,flow closer to the axis of the nanowire when decreasing〈n2d〉ch and W , and get eventually scattered by the otherfacets of the nanowire. This explains the strong depen-dence of the mobility on the cross section at intermediatecarrier densities. At small carrier densities, the transportbecomes limited by Remote Coulomb scattering.

In square trigate devices, the normalized contact re-sistance Rc also tends to increase with decreasing crosssection, especially below W = H = 8 nm. In the stronginversion regime, Rc is about three times larger in a 4×4nm than in a 10 × 10 nm device, hence the denormal-ized contact resistance Rc/Weff is almost 7 times larger.The same analysis as in section IV shows that the ballis-tic and diffusive components of Rc are both enhanced byconfinement in the 4× 4 nm device. The main contribu-tors at high inversion are now impurity scattering, thensurface roughness scattering, the ballistic resistance, andfinally electron-phonon scattering. The increase of diffu-sive components partly results, as for the channel mobil-

Page 10: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

10

0.0 0.5 1.0 1.5 2.0

〈n2d〉ch (cm−2) ×1013

0

1

2

3

4

5

6

7

Rc×〈n

2d〉 ch

(Ω/cm)

×1011

(a)

10× 10 nm

8× 8 nm

5× 5 nm

4× 4 nm

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

〈n2d〉ch (cm−2) ×1013

0

50

100

150

200

250

300

µ(cm

2/V

/s)

(b)

10× 10 nm

8× 8 nm

5× 5 nm

4× 4 nm

0.0 0.5 1.0 1.5 2.0

〈n2d〉ch (cm−2) ×1013

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Rc×〈n

2d〉 ch

(Ω/cm)

×1011

(c)

24× 8 nm

16× 8 nm

8× 16 nm

8× 24 nm

(001) FDSOI

(110) FDSOI

(110) DG

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

〈n2d〉ch (cm−2) ×1013

0

50

100

150

200

250

300

350

400

µ(cm

2/V

/s)

(d)

24× 8 nm

16× 8 nm

8× 16 nm

8× 24 nm

(001) FDSOI

(110) FDSOI

(110) DG

FIG. 10. Contact resistance Rc×〈n2d〉ch (a, c) and channel mobility µ (b, d) as a function of the carrier density 〈n2d〉ch in thechannel, for devices with different cross sections W ×H. The doping profile is “Reference 1.5× 1020 cm−3” from Fig. 2. (a, b)are “square” trigate devices (W = H). The standard deviation on the contact resistance, representative of local variability, isrepresented by the error bars (±1σ). (c, d) are“rectangular” trigate and FinFET devices with either W = 8 nm or H = 8 nm.The mobility and contact resistances of planar (001)/(110) FDSOI devices and symmetric (110) double gate devices (DG) arealso reported for comparison.66

ity, from the enhancement of scattering by confinement.Yet the degradation of Rc is mostly due to the decrease ofthe carrier density under the spacers. The crossover den-sity indeed decreases when reducing cross section (downto n0 ' 5.4× 1011 cm−2 in the 4× 4 nm trigate), so thatthe spacers become a strong bottleneck already at weakinversion densities. This results from the decrease of thenumber of dopants under the spacers and from the in-crease of the surface to volume ratio. Indeed, on the onehand, the linear density of carriers and conductivity un-der the spacers are expected to be proportional (in thesimplest approximation) to the number of dopants perunit length, nd ∝ WH. On the other hand, the spacersfeed a channel with total width Weff = W + 2H. There-fore, the normalized contact resistance of a square trigateis expected to behave as Rc ∝ (W + 2H)/(WH) ∝ 1/W .As a consequence, the contact resistance can be as largeas 75% of the total device resistance in a 30 nm long,

4× 4 nm trigate at gate overdrive Vgt ' 0.7 V (see TableI).

The device to device variability also increases a lotwhen reducing cross section. As the average number ofdopants under each spacer decreases from ' 13 in the10×10 nm device to only ' 2 in the 4×4 nm device, thefluctuations of the number and position of these dopantshave increasing impact on the electrostatics and trans-port properties of the transistors.52,53,67

The contact resistance and mobility in rectangular tri-gate and FinFET devices (Figs. 10c, d) also show sizabledependence on the cross section. Actually, vertical Fin-FET devices withH W are expected to reach the samecontact resistance and mobility µ110 as planar (110),symmetric double-gate (DG) devices,66 while horizontaltrigate devices with W H are expected to reach thesame contact resistance and mobility µ001 as planar (001)FDSOI devices.66 µ110 is smaller than µ001 due to the

Page 11: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

11

W ×H 〈Nimp〉 σimp

〈Nimp〉n0 (cm−2) α

Rc

R(30 nm)4× 4 nm 1.9 0.74 5.4× 1011 0.36 0.755× 5 nm 3.1 0.58 1.8× 1012 0.39 0.678× 8 nm 8.1 0.36 4.9× 1012 0.43 0.6110× 10 nm 12.7 0.28 6.1× 1012 0.47 0.578× 16 nm 16.3 0.24 5.0× 1012 0.46 0.598× 24 nm 24.6 0.19 4.7× 1012 0.48 0.6216× 8 nm 16.3 0.24 7.2× 1012 0.46 0.6024× 8 nm 24.6 0.19 9.8× 1012 0.49 0.58

TABLE I. Average number of impurities 〈Nimp〉 under eachspacer and normalized standard deviation σimp/〈Nimp〉 in tri-gate devices with the “Reference 1.5 × 1020 cm−3” dopingprofile of Fig. 2. Crossover density 〈n0〉, ratio α = Csp/Cox,and ratio between the average contact resistance Rc and thetotal resistance R(L = 30 nm) of 30 nm long devices, at gateoverdrive Vgt = 0.7 V.

101 102

WH (nm2)

103

104

105

Rc(Ω

)

Vgt = 0.7 V

W = H

W < H (FinFETs)

W > H (Horizontal trigates)

FIG. 11. Contact resistance Rc (per wire) as a function ofthe cross-sectional area S = WH in different kinds of de-vices (square and horizontal trigates, vertical FinFETs) atgate overdrive Vgt = 0.7 V.

stronger surface roughness and heavier transport massalong (110) facets [the ground state valley is the Z valleywith transport mass m∗ = 0.19m0 in (001) films and theheavier X,Y valleys with transport mass m∗ ≈ 0.32m0

in (110) films]. The mobility is therefore larger in tri-gate devices with W > H than in FinFET devices withH > W . As an example, the W = 24×H = 8 nm trigateshows 30% larger mobility than the W = 8×H = 24 nmFinFET at high inversion density (〈n2d〉ch = 1013 cm−2).It also shows 30% lower contact resistance, but this per-formance can hardly be related with the better channelmobilities. As a matter of fact, the contact resistanceof planar (001) and (110) FDSOI devices are quite sim-ilar while their mobilities are very different (primarilybecause the spacers do not undergo as strong surface in-version as the channel). Likewise, the denormalized con-tact resistance Rc = Rc/Weff of the W = 8 nm×H = a

FinFET and of the W = a×H = 8 nm trigate are veryclose. As pointed out for square trigates, and confirmedby Fig. 11, Rc is almost inversely proportional to thecross sectional area S = WH of the spacers in this rangeof dimensions (except for the smallest devices). However,the spacers of the vertical FinFET feed a much largerchannel (Weff = 2a + 8 nm) than those of the horizon-tal trigate (Weff = a + 16 nm). This explains why thenormalized contact resistance Rc of the vertical FinFETsshows a weak dependence on H, and remains very closeto the reference (110) double gate device, while the nor-malized contact resistance of horizontal trigate devicesrapidly improves with increasing W . In general, devicesthat maximize effective channel width for a given crosssection will show larger normalized contact resistances.

B. Influence of the doping profile

In this paragraph, we discuss the impact of the dopingprofile on the contact resistance in a W = 10 ×H = 10nm trigate device.

We have considered the different doping profiles plot-ted in Fig. 2. On the source side, the ionized impurityconcentration Nd(z) reads:

Nd(z) = N0d ×

1

1 + 10(z−z1)/λ1× 1

1 + 10(z−z2)/λ2, (13)

where N0d is the dopant concentration deep in the source

(N0d = 1.5 × 1020 cm−3 or N0

d = 3 × 1020 cm−3).Nd(z) is approximately equal to N0

d when z < z1, thendecreases by almost one order of magnitude every λ1

when z1 < z < z2, and one order of magnitude every(λ−1

1 +λ−12 )−1 when z > z2. In the following, z1 = zc−6

nm is the entrance of the source spacer (zc being the en-trance of the channel), while z2 can be varied around zc(see Table II). The “Reference 1.5×1020 cm−3” and “Ref-erence 3 × 1020 cm−3” profiles feature one single decaylength λ1 = 3 nm/decade (λ2 →∞), while the other pro-files feature two distinct decay lengths. The “small un-derlap” profile is an optimal distribution resulting froma process simulation. The doping profiles are symmetricunder the drain.

The contact resistances Rc computed in the W = 10×H = 10 nm trigate device are plotted as a function ofgate overdrive/carrier density in Fig. 12. The deviceswith N0

d = 3×1020 cm−3 all show lower resistances thanthe devices with N0

d = 1.5 × 1020 cm−3. Indeed, largerN0d ’s tend to increase the carrier density in the spacers

[larger n0 in Eq. (12)], hence to decrease the contactresistance. As a matter of fact, n0 is greater than 1013

cm−2 in all N0d = 3× 1020 cm−3 devices.

In general, slower decay under the spacer also decreasesthe contact resistance, as shown by the comparison be-tween the devices with N0

d = 3 × 1020 cm−3. The effectis not spectacular though, as in a 10×10 nm channel thetail of impurities near the gate only contains on averageone donor per nanometer at Nd(z) = 1019 cm−3 – the

Page 12: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

12

Name N0d (cm−3) λ1 (nm) z2 − zc (nm) λ2 (nm) 〈Nimp〉 n0 (cm−2) α

Reference 1.5× 1020 cm−3 1.5× 1020 3.0 0.0 +∞ 12.7 6.1× 1012 0.47Reference 3× 1020 cm−3 3× 1020 3.0 0.0 +∞ 25.4 1.3× 1013 0.61Strong underlap 3× 1020 4.0 −2.0 1.5 28.4 1.4× 1013 0.67Small underlap 3× 1020 4.0 0.0 1.5 32.4 2.1× 1013 0.64Small overlap 3× 1020 4.0 4.0 1.5 34.2 2.3× 1013 0.49

TABLE II. Parameters for the different doping profiles of Fig. 2 [See Eq. (13)]. The average number of impurities 〈Nimp〉 underthe spacers of a W = 10×H = 10 nm trigate device, the crossover density 〈n0〉, and the ratio α = Csp/Cox are also given.

0.0 0.5 1.0 1.5 2.0

〈n2d〉ch (cm−2) ×1013

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

Rc×〈n

2d〉 ch

(Ω/cm)

×1011

Reference 1.5× 1020 cm−3

Reference 3.0× 1020 cm−3

Small underlap

Small overlap

Strong underlap

0.41 0.69 0.97 1.25Vgt (V)

FIG. 12. Contact resistance Rc × 〈n2d〉ch in a W = 10 ×H = 10 nm trigate device as a function of the gate overdrive(upper axis)/carrier density in the channel (lower axis), forthe different doping profiles of Fig. 2.

number of dopants at the entrance/exit of the channelremains therefore very low.

The trends evidenced in Fig. 12 have been confirmedon a W = 8 × H = 16 nm FinFET device. We wouldlike to point out that the above doping profiles have littleimpact on the carrier mobilities in the channel, and onthe threshold voltage Vt at small Vds. Yet larger N0

dand slower decay under the spacers can increase drain-induced barrier lowering at large Vds ; the problem ofthe contact resistances at high fields is, however, beyondthe scope of the present work and will be discussed in another paper.

C. Influence of charges trapped at the Si/Si3N4 interface

Charges trapped in the channel at the Si/SiO2 orSiO2/HfO2 interface are known to be strong scatterers,especially at low carrier densities. In this paragraph, weinvestigate the effects of charges trapped at the interfacebetween the wire and the spacer material.

For that purpose, we have computed the contact resis-tances Rc of a W = 10×H = 10 nm trigate device, withdifferent densities σs of positive or negative charges ran-

σs (cm−2) 〈Qsp〉 (e) n0 (cm−2) α−7.5× 1012 −0.3 8.0× 1011 0.38−2.5× 1012 8.3 3.5× 1012 0.45

(Reference) 0 12.7 6.1× 1012 0.472.5× 1012 17.1 7.9× 1012 0.477.5× 1012 25.7 1.3× 1013 0.48

TABLE III. Average dopant plus trapped charges 〈Qsp〉 un-der the spacers of a W = 10×H = 10 nm trigate device withthe “Reference 1.5×1020 cm−3” doping profile of Fig. 2, as afunction of the density of trap charges σs at the Si/Si3N4 in-terface. The crossover density 〈n0〉 and the ratio α = Csp/Cox

are also given.

domly distributed at the Si/Si3N4 interface. Rc is plot-ted in Fig. 13, as a function of gate overdrive for a givenσs, and as a function of σs for different gate overdrives.The threshold voltage Vt and the carrier mobility in thechannel are little dependent on the density of traps.

The effects of positive and negative trap charges areopposite. Positive trap charges slightly decrease the con-tact resistance, whereas negative trap charges stronglyincrease it. The picture is symmetric for a p-type tran-sistor: positive (resp. negative) trap charges increase(resp. decrease) the contact resistance. Also, the effectof the trap charges is almost the same whether they aredistributed randomly at the Si/Si3N4 interface [curveslabeled “(A)” on Fig. 13], or replaced by a continuousinterface charge distribution [curves labeled “(C)” on Fig.13]. Therefore, the variations of Rc are essentially drivenby electrostatics, and not by scattering (at variance withRCS for example).

As a matter of fact, interface traps behave as sur-face dopants, and can therefore enhance accumulationunder the spacers (n0 increases when σs > 0), or depletethe wire (n0 decreases when σs < 0, see Fig. 14 andTable III). This either increases (σs > 0) or decreases(σs < 0) the conductivity under the spacers. The effectof these traps shall, therefore, be captured by any methodthat reproduces the correct electrostatics, including drift-diffusion models.

The effects of surface traps increase with increasingsurface to volume ratio under the spacers, i.e. with de-creasing channel cross section, and with decreasing dop-ing. As an example, Rc(σs = −2.5 × 1012 cm−2) −Rc(σs = 0) is around twice larger in a 5 × 5 nm thanin a 10×10 nm trigate at high inversion densities. Inter-

Page 13: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

13

1012 1013

〈n2d〉ch (cm−2)

101

102

103

Rc(Ω

.µm)

(a)

σs = −7.5× 1012 cm−2 (C)

σs = −7.5× 1012 cm−2 (A)

σs = −2.5× 1012 cm−2 (A)

σs = 0 (Reference)

σs = +2.5× 1012 cm−2 (A)

σs = +7.5× 1012 cm−2 (A)

σs = +7.5× 1012 cm−2 (C)

0.13 0.41 0.69 0.97 1.25Vgt (V)

−1.0 −0.5 0.0 0.5 1.0 1.5

σs (cm−2) ×1013

0

100

200

300

400

500

Rc(Ω

.µm)

(b)

Vgt = 0.27 V (〈n2d〉ch ≈ 2.8× 1012 cm−2)

Vgt = 0.69 V (〈n2d〉ch ≈ 8.9× 1012 cm−2)

Vgt = 1.01 V (〈n2d〉ch ≈ 1.6× 1013 cm−2)

FIG. 13. (a) Contact resistance Rc as a function of the gateoverdrive (upper axis)/carrier density in the channel (loweraxis), for different densities of trapped charges σs at theSi/Si3N4 interface. The device is a W = 10 × H = 10 nmtrigate with the “Reference 1.5 × 1020 cm−3” doping profileof Fig. 2. “(A)” means atomistic (point-like) charge distri-butions, and “(C)” means continuous background charge dis-tributions at the Si/Si3N4 interface. (b) Contact resistanceRc as a function of σs, for different gate overdrives/carrierdensity in the channel.

estingly, the contact resistance remains almost the sameif a 0.8 nm thick layer of SiO2 is inserted between siliconand Si3N4, and the traps are moved at the SiO2/Si3N4 in-terface. This results, again, from the electrostatic natureof the mechanism at play here, and can be contrastedwith the behavior of RCS in the channel, whose scatter-ing strength decreases exponentially with the thicknessof the interfacial layer of SiO2.7 The effect of chargestrapped in the bulk of the spacers is, as expected, weakerthan the effect of charges trapped at the interface withsilicon. Namely, Rc(σs) ' Rc(ρsts) in the 10 × 10 nmtrigate of Fig. 13, where ρs is the density of traps in thebulk of Si3N4, and ts = 2.5 nm is an effective thicknessrelating the effects of bulk charges to those of surface

−40 −30 −20 −10 0 10 20 30 40

z (nm)

1012

1013

1014

n2d

(cm

−2)

Source Channel Drain

ns

Vgt = 0.69 V

ns = −7.5× 1012 cm−2

ns = −2.5× 1012 cm−2

ns = 0 (Reference)

ns = +2.5× 1012 cm−2

ns = +7.5× 1012 cm−2

FIG. 14. Carrier density n2d in a 30 nm long, W = 10×H =10 nm trigate device with the “Reference 1.5 × 1020 cm−3”doping profile of Fig. 2, for different densities of trappedcharges σs at the Si/Si3N4 interface (gate overdrive Vgt = 0.69V).

charges.As a consequence of these results, a change of spacer

material might improve the contact resistance of one kindof carrier (e.g., electrons) but degrade the contact resis-tance of the other one (e.g., holes) if the spacer and/orits interface with silicon gets charged differently.

VI. COMPARISON WITH EXPERIMENTAL DATA

We conclude this work by a comparison with experi-mental data on trigate devices with cross sections W =14×H = 12 nm and W = 40×H = 12 nm fabricated atCEA/LETI.26 The gate stacks are made of 0.9 nm of SiO2

and 2 nm of HfO2 (effective oxide thickness EOT = 1.3nm). The spacers are 9 nm long. The target doping inthe source and drain is N0

d ' 3× 1020 cm−3; the dopingprofiles are modeled with Eq. (13) using z1 − zc = −9nm, λ1 = 6 nm, and z2 = zc nm, λ2 = 1.5 nm in thesource. These values are characteristic of slightly under-lapped devices.

The experimental contact resistances and mobilitiesare extracted with the same transmission line methodas in the simulations in the L = 50− 200 nm range. TheR(L) data are averaged over 14 devices. The total resis-tance of the W = 40 nm device is plotted as a function ofgate length in Fig. 15a (with ±1σ error bars). It showsthe expected linear behavior, although there is a smallquadratic correction that might be due to the fact thatlong wires tend to thin in the middle. The mobility inthe W = 14 nm device, measured as the slope of theR(L) data, is close to the mobility in 10×10 nm trigates(Fig. 7), and is in excellent agreement with NEGF calcu-lations. The contact resistances obtained from the linearregressions down to L = 0 are plotted as a function of Vgt

Page 14: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

14

0 50 100 150 200

L (nm)

0

200

400

600

800

1000

1200

1400R

(Ω.µm)

(a)

W = 40 nm

Vgt = 0.4 V

Vgt = 0.6 V

Vgt = 0.8 V

Vgt = 1.0 V

0.2 0.4 0.6 0.8 1.0

Vgt (V)

102

Rc(Ω

.µm)

(b)

W = 14 nm Exp.

W = 14 nm NEGF

W = 40 nm Exp.

(001) FDSOI NEGF

FIG. 15. (a) Total resistance measured as a function of gatelength in W = 40 × H = 12 nm trigate devices, for dif-ferent gate overdrives. (b) Measured contact resistances inW = 40 × H = 12 nm and W = 14 × H = 12 nm trigatedevices, as a function of gate overdrive. They are comparedwith simulations for a W = 14 ×H = 12 nm trigate and fora planar, 12 nm thick (001) FDSOI device.

in Fig. 15b, for both the W = 40 nm and the W = 14nm devices. They are compared with NEGF simulationsfor the W = 14 nm device and for a planar, 12 nm thick(001) FDSOI device.

The experimental data show the same ∝ V −βgt depen-dence as the simulations. The experimental contact re-sistances for the W = 14 nm device are, nonetheless,10−20 Ω.µm larger than the simulations in the Vgt ≥ 0.4V range. We remind, though, that the simulations missthe metal/semiconductor contact resistance, which shallactually be around 20 Ω.µm.68,69 The experimental datafor the W = 40 nm device lie, as expected from sectionV A, between the simulations for the W = 14 nm deviceand for the planar (001) FDSOI device. The larger dis-crepancies at Vgt ∼ 0.2 V might result from the loweraccuracy of the extrapolations down to L = 0 (enhanceddevice-to-device variability near the threshold and larger

slope of the R(L) data), and/or from missing Coulombcorrelations in the simulations.

The experimental data therefore support the main con-clusions of this work about the dependence of the contactresistance on the gate bias and on the cross section. Such

a V −βgt behavior, if not properly accounted for in compactmodels for the transistors, can strengthen the dependenceof the mobility on gate length in short channels.57,70,71

VII. CONCLUSIONS

We have computed the contact resistance Rc of tri-gate and FinFET devices using Non-Equilibrium Green’sFunctions. At low drain bias, Rc can represent a verylarge fraction of the total resistance of these devices, sothat the design of the source/drain is as or even moreimportant than the design of the channel in sub-20 nmtechnologies. The spacers between the heavily dopedsource/drain and the gate are the most resistive partsof the devices and dopant fluctuations under these spac-ers a major source of variability. The conductance underthe spacers is typically limited by the poor electrostaticcontrol over the charge density in these areas. The car-rier density can indeed show a dip there, which enhancesboth ballistic and diffusive components of Rc. As a conse-quence, the resistance of the spacers has a near ∝ 1/Vgt

dependence, which, if not properly accounted for, canpartly explain the apparent dependence of the channelmobility on length in short devices.57,70,71 We have in-vestigated the impact of the channel width W and heightH, and the impact of the source/drain doping profiles onthe device performances. The contact resistance Rc ismore dependent on the cross sectional area S = WH ofthe spacers than on the effective width Weff = W + 2Hof the channel. Hence, devices such as FinFETs thatmaximize the effective width for a given cross sectionalarea tend to show larger normalized contact resistancesRc = RcWeff . Also, the contact resistance and variabilitydramatically increase when the cross sectional area of thechannel is . 50 nm2. Finally, we have shown that defectsat the interface between silicon and the spacer materialact as surface dopants, which, depending on their charge,improve or degrade the contact resistance.

This work was supported by the French National Re-search Agency (ANR project NOODLES). The NEGFcalculations were run on the TGCC/Curie machine us-ing allocations from PRACE and GENCI.

1“The international technology roadmap for semiconductors(itrs),” .

2K. Kuhn, IEEE Transactions on Electron Devices 59, 1813(2012).

3C. Jacoboni and L. Reggiani, Reviews of Modern Physics 55, 645(1983).

4F. Gamiz and M. V. Fischetti, Journal of Applied Physics 89,5478 (2001).

5D. Esseni, A. Abramo, L. Selmi, and E. Sangiorgi, IEEE Trans-actions on Electron Devices 50, 2445 (2003).

Page 15: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

15

6K. Uchida and S.-I. Takagi, Applied Physics Letters 82, 2916(2003).

7M. Casse, L. Thevenod, B. Guillaumot, L. Tosti, F. Martin,J. Mitard, O. Weber, F. Andrieu, T. Ernst, G. Reimbold, T. Bil-lon, M. Mouis, and F. Boulanger, IEEE Transactions on ElectronDevices 53, 759 (2006).

8S. Jin, M. Fischetti, and T.-W. Tang, IEEE Transactions onElectron Devices 54, 2191 (2007).

9P. Toniutti, P. Palestri, D. Esseni, F. Driussi, M. D. Michielis,and L. Selmi, Journal of Applied Physics 112, 034502 (2012).

10Y.-M. Niquet, V.-H. Nguyen, F. Triozon, I. Duchemin, O. Nier,and D. Rideau, Journal of Applied Physics 115, 054512 (2014).

11V. Nguyen, Y. Niquet, F. Triozon, I. Duchemin, O. Nier, andD. Rideau, IEEE Transactions on Electron Devices 61, 3096(2014).

12Y. M. Niquet, I. Duchemin, V.-H. Nguyen, F. Triozon, andD. Rideau, Applied Physics Letters 106, 023508 (2015).

13R. Kotlyar, B. Obradovic, P. Matagne, M. Stettler, and M. D.Giles, Applied Physics Letters 84, 5270 (2004).

14S. Jin, M. V. Fischetti, and T. wei Tang, Journal of AppliedPhysics 102, 083715 (2007).

15E. B. Ramayya, D. Vasileska, S. M. Goodnick, and I. Knezevic,Journal of Applied Physics 104, 063711 (2008).

16S. Poli, M. Pala, T. Poiroux, S. Deleonibus, and G. Baccarani,IEEE Transactions on Electron Devices 55, 2968 (2008).

17S. Poli, M. Pala, and T. Poiroux, IEEE Transactions on ElectronDevices 56, 1191 (2009).

18M. P. Persson, H. Mera, Y.-M. Niquet, C. Delerue, and M. Di-arra, Physical Review B 82, 115318 (2010).

19N. Neophytou and H. Kosina, Physical Review B 84, 085313(2011).

20J. W. Lee, D. Jang, M. Mouis, G. T. Kim, T. Chiarella, T. Hoff-mann, and G. Ghibaudo, Solid-State Electronics 62, 195 (2011).

21S. Kim, M. Luisier, A. Paul, T. Boykin, and G. Klimeck, IEEETransactions on Electron Devices 58, 1371 (2011).

22M. Aldegunde, A. Martinez, and A. Asenov, Journal of AppliedPhysics 110, 094518 (2011).

23M. Luisier, Applied Physics Letters 98, 032111 (2011).24K. Akarvardar, C. D. Young, M. Baykan, I. Ok, T. Ngai, K.-W.

Ang, M. Rodgers, S. Gausepohl, P. Majhi, C. Hobbs, P. Kirsch,and R. Jammy, IEEE Electron Device Letters 33, 351 (2012).

25V.-H. Nguyen, F. Triozon, F. Bonnet, and Y. M. Niquet, IEEETransactions on Electron Devices 60, 1506 (2013).

26R. Coquand, S. Barraud, M. Cass, P. Leroux, C. Vizioz, C. Com-boroure, P. Perreau, E. Ernst, M.-P. Samson, V. Maffini-Alvaro,C. Tabone, S. Barnola, D. Munteanu, G. Ghibaudo, S. Monfray,F. Boeuf, and T. Poiroux, Solid-State Electronics 88, 32 (2013).

27R. Coquand, M. Casse, S. Barraud, D. Cooper, V. Maffini-Alvaro, M. Samson, S. Monfray, F. Boeuf, G. Ghibaudo,O. Faynot, and T. Poiroux, IEEE Transactions on Electron De-vices 60, 727 (2013).

28S.-D. Kim, C.-M. Park, and J. Woo, IEEE Transactions on Elec-tron Devices 49, 457 (2002).

29S.-D. Kim, C.-M. Park, and J. Woo, IEEE Transactions on Elec-tron Devices 49, 467 (2002).

30A. Dixit, A. Kottantharayil, N. Collaert, M. Goodwin, M. Jur-czak, and K. De Meyer, IEEE Transactions on Electron Devices52, 1132 (2005).

31P. Magnone, V. Subramanian, B. Parvais, A. Mercha, C. Pace,M. Dehan, S. Decoutere, G. Groeseneken, F. Crupi, andS. Pierro, Microelectronic Engineering 85, 1728 (2008).

32D. Tekleab, S. Samavedam, and P. Zeitzoff, IEEE Transactionson Electron Devices 56, 2291 (2009).

33M. G. Parada, C. Malheiro, P. G. Agopian, and R. C. Giacomini,ECS Transactions 39, 255 (2011).

34S. J. Park, D.-Y. Jeon, L. Monts, S. Barraud, G.-T. Kim, andG. Ghibaudo, Semiconductor Science and Technology 28, 065009(2013).

35C.-W. Sohn, C. Y. Kang, M.-D. Ko, D.-Y. Choi, H. C. Sagong,E.-Y. Jeong, C.-H. Park, S.-H. Lee, Y.-R. Kim, C.-K. Baek, J.-S.

Lee, J. Lee, and Y.-H. Jeong, IEEE Transactions on ElectronDevices 60, 1302 (2013).

36T. An, K. Choe, K.-W. Kwon, and S. Kim, Journal of Semicon-ductor Technology and Science 14, 425 (2014).

37A. Pereira and R. Giacomini, Microelectronics Reliability 55, 470(2015).

38J.-S. Yoon, E.-Y. Jeong, S.-H. Lee, Y.-R. Kim, J.-H. Hong, J.-S.Lee, and Y.-H. Jeong, Japanese Journal of Applied Physics 54,04DC06 (2015).

39S. Berrada, M. Bescond, N. Cavassilas, L. Raymond, andM. Lannoo, Applied Physics Letters 107, 153508 (2015).

40J.-P. Colinge, Solid-State Electronics 48, 897 (2004).41M. P. Anantram, M. S. Lundstrom, and D. E. Nikonov, Pro-

ceedings of the IEEE 96, 1511 (2008).42W. Shockley, “Research and investigation of inverse epitaxial

UHF power transistors”, Report No. A1-TOR-64-207, Air ForceAtomic Laboratory, Wright-Patterson Air Force Base, Ohio,September 1964.

43The system is therefore modeled as a periodic array of nanowiresconnected to thick source and drain films. The convergence withrespect to tSD, WSD and LSD has been carefully checked.

44S. M. Goodnick, D. K. Ferry, C. W. Wilmsen, Z. Liliental,D. Fathy, and O. L. Krivanek, Physical Review B 32, 8171(1985).

45The probability to find a dopant at at given point of the sourceis proportional to the background dopant concentration Nback

shown in Fig. 2. A distribution of point charges Npoint is pickedrandomly, and mixed with Nback near the edge z = zmin ofthe simulation box: N = α(z)Nback + [1 − α(z)]Npoint, where

α(z)−1 = 1 + 10(z−zmix)/λmix , with zmix = zmin + 4 nm andλmix = 0.66 nm.

46G. Stefanucci and R. van Leeuwen, Nonequilibrium Many-BodyTheory of Quantum Systems: A Modern Introduction (Cam-bridge University Press, 2013).

47G. Bastard, Wave mechanics applied to semiconductor het-erostructures (Les Editions de Physique, 1988).

48M. S. Sze and K. N. Kwok, Physics of Semiconductor Devices(John Wiley and Sons, New-York, 2006).

49M. Fischetti, S. Jin, T.-W. Tang, P. Asbeck, Y. Taur, S. Laux,M. Rodwell, and N. Sano, Journal of Computational Electronics8, 60 (2009).

50C. Li, M. Bescond, and M. Lannoo, Phys. Rev. B 80, 195318(2009).

51R. Lavieville, F. Triozon, S. Barraud, A. Corna, X. Jehl, M. San-quer, J. Li, A. Abisset, I. Duchemin, and Y.-M. Niquet, NanoLetters 15, 2958 (2015).

52N. Seoane, A. Martinez, A. Brown, J. Barker, and A. Asenov,IEEE Transactions on Electron Devices 56, 1388 (2009).

53A. Martinez, M. Aldegunde, N. Seoane, A. Brown, J. Barker,and A. Asenov, IEEE Transactions on Electron Devices 58, 2209(2011).

54S. Datta, F. Assad, and M. Lundstrom, Superlattices and Mi-crostructures 23, 771 (1998).

55M. Shur, IEEE Electron Device Letters 23, 511 (2002).56G. Ghibaudo, Electronics Letters 24, 543 (1988).57D. Rideau, F. Monsieur, O. Nier, Y. Niquet, J. Lacord,

V. Quenette, G. Mugny, G. Hiblot, G. Gouget, M. Quoirin, L. Sil-vestri, F. Nallet, C. Tavernier, and H. Jaouen, in 2014 Inter-national Conference on Simulation of Semiconductor Processesand Devices (SISPAD) (2014) pp. 101–104.

58J.-H. Rhew, Z. Ren, and M. S. Lundstrom, Solid-State Electron-ics 46, 1899 (2002).

59M. Buttiker, Y. Imry, R. Landauer, and S. Pinhas, PhysicalReview B 31, 6207 (1985).

60S. Datta, Electronic Transport in Mesoscopic Systems (Cam-bridge University Press, Cambridge, 1995).

61C. Jacoboni, C. Canali, G. Ottaviani, and A. A. Quaranta, SolidState Electronics 20, 77 (1977).

62(), RNW0 can either be obtained from a ballistic NEGF calculation

in a homogeneous nanowire or from the band structure of that

Page 16: Contact resistances in trigate and FinFET devices in …Contact resistances in trigate and FinFET devices in a Non-Equilibrium Green’s Functions approach L eo Bourdet, 1,2Jing Li,

16

nanowire (see, e.g., Refs. 63 and 64).63K. Natori, Journal of Applied Physics 76, 4879 (1994).64A. Rahman, J. Guo, S. Datta, and M. Lundstrom, IEEE Trans-

actions on Electron Devices 50, 1853 (2003).65(), Vt can be slightly different from Vt defined in paragraph II C,

as the former is based on a density criterion, and the latter on acurrent criterion.

66(), the planar (001) FDSOI device used for comparison in Fig. 10is a 8 nm thick Si film on a 25 nm thick BOX with the same gatestack and surface roughness parameters as the trigate devices(namely Λ = 1.5 nm and ∆ = 0.25 nm on the bottom interfacewith the BOX, Λ = 1.5 nm and ∆ = 0.35 nm on the top interfacewith the gate stack). The planar (110) FDSOI device is the same,but with Λ = 2.0 nm and ∆ = 0.45 nm on both top and bottominterfaces. The planar (110) double-gate device is, likewise, a

symmetric stucture with a 8 nm thick film, and Λ = 2.0 nm,∆ = 0.45 nm on both interfaces.

67S. Markov, B. Cheng, and A. Asenov, IEEE Electron DeviceLetters 33, 315 (2012).

68(), the source and drain can be modeled as 90×90 nm pads withcontact resistivities ρc ' 2× 10−8 Ω.cm2. For a typical effectivewidth Weff = 50 nm, these pads therefore make a contributionto the contact resistance R′c ' 24 Ω.µm.

69K. Ohuchi, C. Lavoie, B. Yang, M. Kondo, K. Matsuzawa, andP. M. Solomon, Japanese Journal of Applied Physics 51, 101302(2012).

70M. Zilli, D. Esseni, P. Palestri, and L. Selmi, IEEE ElectronDevice Letters 28, 1036 (2007).

71V. Barral, T. Poiroux, D. Munteanu, J. Autran, andS. Deleonibus, IEEE Transactions on Electron Devices 56, 420(2009).