Toward the End of the MOSFET Roadmap:

advertisement
Toward the End of the MOSFET Roadmap:
Investigating Fundamental Transport Limits
and Device Architecture Alternatives
by
Anthony Joseph Lochtefeld
Bachelor of Science in Electrical Engineering
Ohio Northern University, May 1990
Master of Science in Electrical Engineering
Purdue University, August 1996
Submitted to the Department of Electrical Engineering and Computer
Science in partial fulfillment of the requirements for the degree of
Doctor of Philosophy
in Electrical Engineering and Computer Science
at the
Massachusetts Institute of Technology
May 2001
©Massachusetts Institute of Technology, 2001. All Rights Reserved.
A u th o r ............ . . ...
....
. . ....
. ........
e n.
Departmen olfrical Engineefing and Computer Science
May 15, 2001
Certified by .........................................
Dimitri A. Antoniadis
Professor of Electrical Engineering
-fTl9sis Supervisor
.........
Accepted by ...................
.ASSOCUSE
s
JUL . 1200
LIBRARIES
Arth rC. Smith
rofessor of Electrical Engineering
Graduate Officer
ARKER
Toward the End of the MOSFET Roadmap: Investigating Fundamental
Transport Limits and Device Architecture Alternatives
by
Anthony J. Lochtefeld
Submitted to the Department of Electrical Engineering and Computer
Science on May 15, 2001, in partial fulfillment of the requirements
for the degree of Doctor of Philosophy in Electrical Engineering
and Computer Science
Abstract
MOSFET scaling and performance has progressed rapidly in recent years, with physical gate
lengths for electrostatically sound devices reaching 30 nm or below: near the prospective scaling
limits for traditional bulk MOSFETs. This work investigates several key issues for this "end of
roadmap" regime. Focus is on understanding the limitations to carrier velocity in MOSFET
inversion layers as channel lengths are scaled well below 100 nm, and on relaxing these limits
through architectural alternatives to bulk MOSFETs.
It has been proposed that drain current is ultimately limited by the rate at which carriers can be
thermally injected from the source into the channel. In this work it is shown that commonly used
techniques for experimentally determining carrier velocity are insufficient to determine how close
modem MOSFETs operate to this ballistic or "thermal limit". A new technique is proposed, and
applied to two advanced industry technologies with deep-sub-100-nm channel lengths. It is
shown that a IV NMOS technology with Leff < 50 nm operates at no more than -40% of the limiting thermal velocity. Furthermore, no indication is found that continued scaling is bringing us
closer to the thermal velocity limit.
Via simulation, the relationship between mobility and scaling is investigated for bulk silicon
NMOSFETs and FDSOI (Fully-Depleted Silicon-On-Insulator) alternatives, focusing on the 50
and 25 nm Leff generations. Scaling of bulk MOSFETs well below 100 nm Leff requires heavy
channel doping, leading to degraded low-field mobility. Provided that the gate workfunction is
used to determined the threshold voltage, FDSOI devices do not suffer from this trade-off, by virtue of the fact that their channel can be undoped. It is shown that single-gate FDSOI is, accordingly, an attractive alternative down to 50 nm Leff. For deeper scaling, double-gate FDSOI should
have approximately a 3X mobility advantage over bulk NMOS.
With careful determination of channel length, inversion-layer charge, and series resistance, it is
possible to study experimentally the relation between channel length and mobility in deep-sub100-nm MOSFETs. With the aid of inverse modeling techniques, evidence is found that in the
very shortest modern MOSFETs, mobility is less then would be expected from "universal" mobility, and independent of transverse field. This may be indicative of a transition in the dominant
scattering mechanism, from surface- to Coulomb- scattering.
3
The relevance of low-field mobility to the performance of deep-sub-100-nm MOSFETs is not
well understood. In this work this relationship is studied experimentally: mobility (with low lateral electric field) is modified by externally applying uniaxial stress to NMOS devices, and the
corresponding shift in carrier velocity (with high lateral field) is measured. The dependence of
velocity on mobility is found to be significant, and is found to correspond well with the predictions of energy-balance modeling.
Given their promise for scalability without mobility degradation, the design space for FDSOI
MOSFETs merits detailed study. Via 2D simulation, scalability and drive current are investigated
for three basic FDSOI alternatives: single-gate, and double-gate either with symmetrical workfunction mid-gap gates or asymmetrical workfunction n+p+ gates. For the single-gate device, it
is shown that scaling below Leff = 35 nm may not be achievable for practical silicon film thickness, unless Ioff requirements are relaxed. For double-gate devices we have shown that hypothetical mid-gap top- and bottom- gates are superior to n+/p+ poly gates, for both scalability and drive
current.
Realization of the ideal double-gate device structure involves three major technical challenges:
formation of gates above and below a thin single-crystalline silicon layer, achievement of fine
alignment between top- and bottom-gates, and achieving low source/drain resistance for the thin
silicon film. These issues are addressed in this work through demonstration of three primary technologies: wafer bonding with pre-patterned features, interferometric alignment, and selective epitaxy for raised source/drains.
Thesis Supervisor: Dimitri A. Antoniadis
Title: Professor of Electrical Engineering
4
Acknowledgments
Many individuals have supported me, both professionally and personally, throughout the course
of this work. I would like to thank Professor Dimitri Antoniadis for years of guidance and
patience, and for allowing me considerable freedom in pursuing this research. His technical
direction and vision have made this work possible. I would also like to thank, especially, professors Hank Smith, Jim Chung, Eugene Fitzgerald, Anantha Chandrakasan, and Judy Hoyt for guidance in various stages of this work, and professors Smith and Hoyt for serving as readers of this
thesis. I would also like to thank my former advisor, Professor Michael Melloch at Purdue University, for the opportunity to work in his group and my introduction to this field.
The exchange of knowledge and perspectives with members of my research group has been
invaluable. I am thankful to have had the chance to work with Isabel Yang, Melanie Sherony,
Zachery Lee, Mark Armstrong, Andy Wei, and Keith Jackson, all of whom now have a "Dr."
before their name. And many others: Andrew Ritenour, Siva Narendra, Hasan Nayfeh, Ihsan
Djomehri, Jim Fiorenza, Isaac Lauer, Ilia Sokolinski, Corina Tanasa, Dr. Jong-Ho Lee, Dr.
Duheon Song, and Dr. Hitoshi Wakabayashi. Collaboration with members of other research
groups at MIT has been important as well. In particular I would like to thank Dr. Thomas Langdo,
Mitchell Meinhold, and Dr. Maya Farhoud. Thanks as well to Euclid Moon, Dr. David Pflug, Dr.
Juan Ferrara, Dr. Matthew Currie, and Dr. Lalitha Parameswaran.
Many thanks to the research, support and administrative staff at the Microsystems Technology
Laboratory and the Nanostructures Laboratory. Among many names I would like especially to
mention Bernard Alamariu, Paul Tierney, Patricia Burkhardt, James Carter, Dan Adams, and
Wayne Price. Many others as well: Dr. Vicky Diadiuk, Joe Walsh, Barry Farnsworth, Paudely
Zamora, Kurt Broderick, Paul Mcgrath, Ron Stoute, Joe DiMaria, Dennis Sullivan, Brian Foley,
Jim Daley, Deb Hodges-Pabon, Christina Gordy, and Mark Mondol. And more. And at Lincoln
Laboratory: special thanks to Andy Loomis; thanks also to Jim Reynolds, Jim Burns, and Skip
Hoyt.
Beyond the research, life here has been enriched by the people, friends in the lab, office, and pub.
Andy, Andy, and Andy. Mark, Fletch, Tom, and Isaac. It's been some trip, hasn't it?
And beyond MIT, I think of a few special people who have given me support through these long
years. Old friends and new, you know who you are. And Brenda, who has kept me going through
the end of all this. And finally my family, and most of all my parents.
5
6
Contents
1 Introduction
1.1 Background .................................................
Velocity and Scaling in MOSFETs: the Roadmap Perspective .........
Velocity and Scaling in MOSFETs: Device Physics Perspective .......
Goals of This Work ..........................................
1.2 Outline .......................................................
21
21
22
24
25
26
28
2 Experimental Determination of Carrier Velocity in Deeply Scaled NMOS
28
2.1 Introduction .................................................
29
2.2 The Thermal Injection Limit .....................................
2.3 Investigation of Velocity Measurement Techniques by Simulation .......... 31
32
Carrier Velocity from Saturated Transconductance ..................
Carrier Velocity from Drain Current and Short-Device CV............ 34
35
Simulation Results: Comparison of Techniques.....................
39
2.4 Experimental Results ..........................................
39
Technology Parameters .......................................
43
M easurements ..............................................
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
How Close? . .
46
----..................................
Dependence of veff on Vdd
2.5 Further Investigation of the Inversion-Charge Relationship Between Long and
47
Short Devices ..............................................
51
Conclusion
..................................................
2.6
3 Mobility and Scaling I: Universal Mobility in Bulk and FDSOI MOSFETs
3.1 Introduction .................................................
Universal Mobility ..........................................
3.2 Mobility and Scaling in Bulk MOSFETs ............................
Bulk MOSFET Scaling .......................................
Transverse Field Determination .................................
M obility versus Generation ....................................
3.3 Mobility and Scaling in FDSOI MOSFETs...........................
3.4 Results .......................................................
3.5 Conclusion ..................................................
7
52
52
52
53
53
58
60
60
62
64
4 Mobility and Scaling II: Channel Length Dependence in Super-Halo NMOS
4.1 Introduction .................................................
Scattering Mechanisms in Deeply Scaled Bulk MOSFETs .............
4.2 On Experimental Determination of Low-Field Mobility. ..........................
4.3 Experimental Results: Long Channel Mobility ........
65
65
66
67
.70
.70
71
.75
.75
.78
79
80
Mobility versus Vgs ..........................
4.4
4.5
4.6
Mobility versus Effective Transverse Field .........
Experimental Results: Mobility versus Channel Length.
Effective Mobility versus Channel Length at Constant
Effective Mobility versus Channel Length at Constant
Effective Mobility versus Channel Length at Constant
Technology B ................................
Investigation of Sources of Error...................
Effects of Error in Rsd and Leff Estimates ..........
Results for Longer Channels.....................
C onclusion ...................................
Gate Bias ...
Overdrive ...
Effective Field
.... .... ....
. . . . . . . . . .. .
... ..... ....
.81
81
83
.83
5 Electron Velocity Dependence on Mobility in Deep-sub-100-nm bulk NMOS
5.1 Introduction .................................................
5.2 Experiment ......................................
M obility with Uniaxial Stress .......................
5.3
Velocity versus M obility ..........................
On Sources of Error in RR Estimation.................
Series Resistance Effects ..........................
Drain Bias for Low-Field Measurements ..............
5.4
Energy Balance M odeling ...........................
5.5
On Energy- and Momentum- Relaxation Interdependence
Conclusion ......................................
85
85
.86
86
89
.91
91
93
.94
95
.96
6 FDSOI Design-Space
6.1 Introduction ..............
Methodology ............
6.2 Single Gate Scaling Results. .
6.3 Double Gate Scaling Results.
DG Alternatives .........
Simulation Results .......
6.4 Relaxing Ioff Criteria .......
6.5 Energy Balance Simulation of Ion
6.6 Conclusion ...............
97
.. 97
7 Double-Gate Fabrication Technology
7.1 Introduction ......
7.2 Formation of Gates
7.3 Alignment of Gates
7.4 Reduction of S/D Parasitic Resistance. .
Raised Source and Drain Process . ..
113
.113
.113
.116
.116
.119
. 97
.. 99
.101
.101
.103
.108
.110
.112
8
7.5
C onclusion ...................................................
121
8 Conclusion
8.1 Summary of Results...........................................
Experimental Results ........................................
Simulation Results ..........................................
8.2 Future Work................... .... ..........................
Experimental Analysis of Deep-Sub-100-nm Devices ...............
Fabrication ...............................................
122
122
122
122
123
123
125
A Simulation Techniques
134
9
10
List of Figures
Figure 1.1: Effective electron velocity veff required to meet 1999 ITRS Roadmap targets
(NMOS). Unscaled T0 X: > 1.5 nm. Unscaled Vt: 0.25 V. Vt values used (Table 1.1) correspond to short-channel devices in each technology, so are well below roadmap guide23
line of 0.4 V . ..........................................................................................................
Figure 1.2: Potential profile and electric fields in a bulk NMOSFET channel, for Vgs =
Vds = Vdd. For acceptably low DIBL (Drain Induced Barrier Lowering) IEy must be sig. 25
nificantly greater than Ex . ........................................................................................
Figure 1.3: FDSOI M OSFET Architectures ..............................................................
26
Figure 2.1: Carrier velocity considered at source-to-channel barrier. Unless transport is
ballistic, average carrier velocity (veff) at the source-to-channel barrier must be less than
the thermal injection velocity vO. For deep-sub-100-nm devices conduction-band peak
is typically within several nm of source-channel transition (x=0, defined as the point
where source doping falls to 2x10 19 cm- 3 [19]). Note that simulations for bulk MOSFETs
indicate a conduction-band peak much less pronounced than shown in this schematic30
Figure 2.2: Illustration of relationship between carrier velocity and inversion-charge density along MOSFET channel. Since there is no single ve for the channel, different extraction techniques for ve can give significantly different answers.................................30
Figure 2.3: Monte-Carlo simulation results: electron velocity as a function of position in
an ultra-scaled bulk NMOS channel. T0 X phys=1.5 nm. Vgs=Vds = 1V. Courtesy of D.
1
F ran k (IB M )....................................................................................................................3
Figure 2.4: NMOSFET simulation structure (Avant! MEDICI TM [13]). Mobility models used in this chapter: (a) for low Ex, concentration-dependent pt; (b) for EY, Si-Si0 2
universal g assumed, with coefficients from [13]; (c) for high Ex, either a Caughey-Thomas relationship [64] for DD simulations, or an electron-temperature dependent model as
described in Appendix 1 for EB simulations. For DD simulations, electron saturation ve-
11
locity vsat = 1.3x 10 7 cm/s. For EB simulations, energy relaxation time t
vsat=1x10 7
= 0.1 ps, and
cm/s. From simulation, T0 xelec was assessed from the slope of the dQi/dVgs
curve for mid-channel for a long-channel device at Vds=O; slope was taken at Vgs=Vdd-
Classical distribution was assumed for inversion layer electrons for all simulations. Leff
is determined by the 2x10 19 cm- 3 S/D doping points. Fermi-Dirac statistics. External
lumped resistance of 90 Q2pm per side, to give total Rsd = 220 Qtm........................32
Figure 2.5: EB simulation results for the device of Figure 2.4. Investigating components
of transconductance as a function of position in channel. (Avant! MEDICI TM). ........ 33
Figure 2.6: Example of Qi(xo) determination. Capacitance is measured on large device.
For large device, overlap C is inconsequential, so area under curve corresponds to inversion layer charge. Integration limit is adjusted by AVt (= DIBL + Vt-rolloff - IonRs) as
determined from the target short device. Qi(xo) = (Qio+AQi) / (W*L). NMOS. ....... 35
Figure 2.7: EB simulation for the device of Figure 2.4: area between curves gives
which corresponds to Qi as determined in Eqn. (2.9). ...............................................
Qcgs,
36
Figure 2.8: DD simulation results (solid): output characteristics for the device of Figure
2.4, matching saturated gm and Ion of experimental technology A (discussed in Section
2 .4 ). ................................................................................................................................
37
Figure 2.9: Simulation results: conduction-band and carrier velocity vs. position for bulk
super-halo device of Figure 2.4. Velocities at five marked points along the channel correspond to values given by different MOSFET carrier velocity extraction techniques (Table 2.1). Solid symbols are corrected for S/D series resistance, open symbols are
uncorrected. Vt, Txelec, Rsd, and Ion approximately match the "Technology A" under in-
vestigation. DIBL here is 66 mV/V. (a) DD model: vsat = 1.3x107 cm/s (b) EB model:
= 0 . I s.........................................................................................................................3
8
Figure 2.10: Monte-Carlo results: output characteristics and electron inversion-layer velocity, for a 25 nm Leff NMOSFET design presented in [2]. Courtesy of D. Frank (IBM).
Inversion-layer distribution is classical. Our estimates of vid and vgm shown are approxim ate v alues. ...................................................................................................................
40
Figure 2.11: Example output characteristics for NMOS devices from two technologies
under investigation. DIBL is similar for both devices shown, 60-70 mV/V.............42
Figure 2.12: Measured Vt and DIBL for technology A. Results shown are for same set
of devices used for inverse-modeling determination of Rsd and Leff. Leff corresponds to
the points where S/D doping falls to 2x10 19 cm-3 [19]. DIBL is via Eqn. (2.10). ........ 42
12
Figure 2.13: Experimental results: carrier velocity by four techniques vs. DIBL, for technology A. Solid symbols are corrected for S/D series resistance, open symbols are uncorrected. The implicit variable in the x-axis is gate length.....................................44
Figure 2.14: Experimental results: carrier velocity by four techniques, for technology B.
Solid symbols are corrected for S/D series resistance, open symbols are uncorrected. .45
Figure 2.15: Experimental results: effective electron velocity measured from vidiDIBL=100 mV/V in all cases. For technologies A and B, as a function of Vdd. ..... 47
Figure 2.16: EB simulation result for the device of Figure 2.4: Incremental gate capacitance vs. channel position, with Vgs = Vds = IV. Compared with and without lumped external resistance, to long channel value C'iong-inv (Vds = 0). x0 , indicated above, is
position of Ec-peak from simulation at Vgs=Vds=lV. (Determined from a lateral cut
through the device, at approximately the inversion layer depth)...............................49
Figure 2.17: EB simulation result for the device of Figure 2.4: Inversion layer charge vs.
channel position for first half of channel. Vds=1V. Upper-right insert shows size of corrections made to long-channel Qj to obtain Qi(xo) via Eqn. (2.6). Short-channel correc50
tion increases, and Rsource correction decreases, Qi(xo). ..........................................
Figure 2.18: Illustration of three methods of determining conduction-band potential with
position: top two are weighted averages of electron potential, to the depth specified. Bottom is simple 1-D cut along the length of the device, at 0.4 nm. DD simulation, for structure of Figure 2.4; only first half of channel is shown...............................................51
Figure 3.1: Measured NMOS effective mobility for FDSOI and bulk MOSFETs. From
[37]. The deviation for Eeff < 0.6 MV/cm is likely due to overestimation of Qj that can
occur unless Vgs >> Vds [38] when measuring via the split-CV method [52]. Vds =50 mV.
54
Modern bulk MOSFETs operate at Eeff > 1 MV/cm .................................................
Figure 3.2: Simulation structures for NMOS bulk doping alternatives investigated. Junctions are spatially abrupt. For (a) 09 = 4.19 V (for n+ Si). For (b) c9 = 4.19 V and <Db
= 5.25 V (for p+ Si). For Nb and Tstep values by scaling node, see Figures 3.4 - 3.6.
54
Source and drain are n+. .............................................................................................
Figure 3.3: Results for uniform-doped bulk NMOSFET: shown are 2D simulation data
points for Vt vs. Nb, from which Nb values meeting criteria (2) are interpolated. (Avant!
M ED IC I TM )..................................................................................................................56
Figure 3.4: 2D simulation results for uniform-doped bulk NMOSFET: doping level (Nb)
13
correlated with Leff, across four generations. Requisite doping levels to meet two separate criteria are show n. ............................................................................................
57
Figure 3.5: Results for perfect step-doped bulk NMOSFET: shown are 2D simulation
data points for Vt vs. step depth, from which Tstep values meeting criteria (2) are interpolated. (Avant! M ED ICI TM ).....................................................................................
57
Figure 3.6: 2D simulation results for perfect step-doped bulk NMOSFET: step depth
(Tstep) correlated with Leff, across four generations. Tstep required for meeting two separate criteria are show n. ............................................................................................
58
Figure 3.7: 2D simulation results: maximum effective field Eeff along NMOSFET channel for two bulk doping alternatives, with n+ poly gates, using Eqn. (3.4). Shown vs. scaling across four generations. Devices are in saturation (Vds = Vds = Vdd)- Qi/q =
C' ,(Vdd-Vt)/q = 1x10 13 cm-3 was used for all generations. Also shown is dashed line
for Eeif = Qj / 2 Fsi. This corresponds to zero channel charge and no built-in potential between top and bottom contacts (Qb =0, (bi = 0), which is the situation for SG- and DGFDSOI devices with mid-gap workfunction gates......................................................59
Figure 3.8: Calculated universal-curve mobilities for two bulk NMOSFET doping alternatives, correlated with scaling. Values are plotted using Eqn. (3.2) with Eeff values from
Figure 3.7. For Eqn. (3.2), go = 670 cm2/Vs, Eo= 0.67 MV/cm, n = 1.6, all according to
[31]. Included is dashed line corresponding to the limit of Qb = 0 and cDbi = 0, applicable
to SG- and DG- FDSOI devices with mid-gap workfunction gates. .........................
61
Figure 3.9: Universal effective NMOSFET mobility (for low lateral field) from Eqn.
(3.2), with regions of operation delineated for four different bulk and FDSOI device architectures. For each device architecture, the range of Eeff corresponds to: 0.8x10 13 <
Qi/q < 1.2x1013 cm- 3 . In the case of DG-Dmg, Qj is the sum of inversion charge in both
inversion layers, except for special case noted in (b). For DG-Dmg and SG-mg, since
Qb = 0 Eqn. (3.3) was used to estimate Eeif, independent of channel length. For bulk architectures, peak Eeff along channel was determined via 2D simulation, as Esurf - Qi/2cs.
........................................................................................................................................
63
Figure 4.1: Universal vs. bulk-limited mobility as a function of channel doping. NMOS
Universal mobility is from Eqn (3.2) using coefficients from [31], and is based on Ni =
1x1013 cm-3 . Calculation of Eeff is one-dimensional from Eqn. (3.3): lateral charge sharing effects are ignored. Bulk-limited mobility is for electrons in p-region. ............. 67
Figure 4.2: Simulation results: comparison of mid-channel Qi-Vgs relationship in short
vs. long NMOS devices, with TOxelec = 1.9 nm and uniform channel doping: Na =
1.25x10 18 cm-3 for short devices and 0.6x10 18 cm- 3 for long. Abrupt junctions are 21 nm
14
deep. DIBL > 300 mV/V for 35 nm device, 100 mV/V for 50 nm device, and < 10 mV/
70
V for the 500 nm device. .............................................................................................
Figure 4.3: (a) Gate leakage current I9 for technology A, with effect on source and drain
currents. (b) Correction to allow Id determination for mobility measurements..........72
Figure 4.4: Measured effective mobility vs. Vgs for technologies A and B...............73
Figure 4.5: Two experimental methods of Eeff determination....................................73
Figure 4.6: Measured effective mobility compared for two methods of Eeif determination.
Technology B. Qj and Qb are from split-CV meas. [52]. Toxelec = OX /C' gsd(Vgs) -...7
Figure 4.7: Measured effective mobilities for technologies A and B, plotted along with
universal curve mobilities. For measured data, Eeff is determined via Eqn. (4.8).........75
Figure 4.8: Measured Iteff vs. Leff for different gate biasing. Technology A. Toxelec=2.4
76
nm . DIBL at Leff = 45 nm = 120 mV/V......................................................................
Figure 4.9: Intrinsic drain-to-source bias in (= Vds - IdRs) channel for pgff measurements
77
of Figure 4.8. Applied Vds is 50 mV. ......................................................................
Figure 4.10: Measured dependence of short-device pgef extraction on applied Vds, for a
77
. .................................................................................................
range of Leff. Vgs =
Figure 4.11: Measured Reff vs. Leff for different gate overdrive. Technology A. Applied
78
V gs is indicated in inset.............................................................................................
Figure 4.12: Range of Qj corresponding to Figure 4.11. ...........................................
79
Figure 4.13: Measured pge vs. Leff for different Eeff. Technology A. Eeif is determined
80
via Eqn. (4.8). Applied Vgs is indicated in inset........................................................
Figure 4.14: Effects of error in Leff estimation, +/- 5nm. Actual estimated AL is 65 nm.
82
(a) assumes AL = 60 nm. (b) assumes AL = 70 nm. ................................................
Figure 4.15: Effects of error in S/D series resistance estimation, +/- 10%. Actual estimated Rsd is 190 Qpm. (a) assumes Rsd = 209 2gm. (b) assumes Rsd = 171 2gm..82
15
Figure 4.16: Measured pge vs. Leff for different gate biasing. Technology A. T0 "elec=2.4
83
nm. Results for points of Leff < 150 nm same as in Figure 4.8. ..............................
Figure 5.1: Schematic illustration of strain experiment............................................87
Figure 5.2: Experimental results: normalized mobility shift in long and short (Leff = 10
gm, -45 nm) devices vs. uniaxial strain. Each point represents a fractional shift relative
to unstrained value (which is represented by *). ........................................................
88
Figure 5.3: Experimental results: normalized velocity shift (near the source) in short (Leff
-45 nm) device vs. uniaxial strain. Each point represents a fractional shift relative to unstrained value (which is represented by *). ..............................................................
89
Figure 5.4: Experimental results: 6 ve extracted by two methods, versus long-device kteff.
Mobilities are measured at (VgsVds) = (IV, 50 mV). Velocity viad is measured at
(Vgs,Vds) = (IV, IV). Velocity vgmi is from peak saturated intrinsic transconductance,
measured at Vds= 1V; the peak occurs for Vgs somewhat less than 1V. For this data 6 ve
/ pgeff = 0.31 when using vgmi, and 0.30 for Vidi. Each point represents a fractional shift
relative to unstrained value (which is represented by *) ...........................................
90
Figure 5.5: Experimental results: 8ve extracted by two methods, versus short-device
pg9ff. Gate and drain biases are as described for Figure 5.4. For this data, the ratio Rg
= 6 ve / 4Peff = 0.48 when using vgmi, and 0.46 for vidi. Each point represents a fractional
shift relative to unstrained value (which is represented by *). ..................................
91
Figure 5.6: Resistance test structure setup. Current was forced through terminals 1-4, and
the corresponding voltage drop measured between terminals 2 and 3. Test structure
resistance Rt was measured for strain = 0, 0.04, and 0.08%. Rt V3 -V2 / 114 .-........ -93
Figure 6.1: Scaling simulations for SG-CDmg device. (Avant! TMA Medici [13]). Driftdiffusion with Lombardi mobility model. No gate depletion. S/D are n+ 1.5x1020 cm~3 ,
abrupt junctions. Na = 1x1015 cm~3 in channel. In (a) and (b) each point corresponds to
a set of IV transfer characteristics from device with specified Tsi, from which DIBL (according to Eqn. 2.10) and Ioff were extracted. Tox and Vdd from Table 6.1...............100
Figure 6.2: Interpolated results from Figure 6.1: Tsi required per scaling node of Table
6.1. TX and Vdd vary with Lcff according to Table 6.1...............................................101
Figure 6.3: Alternative double-gate FDSOI structures.................................................102
16
Figure 6.4: Ratio in inversion charge in DG-n+p+ vs. DG- 4 mg MOS, according to longchannel charge-sheet m odel..........................................................................................102
Figure 6.5: Scaling simulations for DG-n+p+ device. Model and doping specifications
from Figure 6.1 apply. Tox and Vdd from Table 6.1...................................................103
Figure 6.6: Scaling simulations for DG-mg device. Model and doping specifications
from Figure 6.1 apply. Tox and Vdd from Table 6.1...................................................104
Figure 6.7: Interpolated results from Figure 6.5-6.6: Tsi required to meet short-channel
criteria for each scaling node of Table 6.1. Tox and Vdd vary with Leff according to Table
104
6 .1 . ...............................................................................................................................
Figure 6.8: Simulation results: DIBL and Vt-rolloff for DG structures. Vt-rolloff =
Vt(long-device) - Vt(short device), where Vt is linear extracted at Vds=50 mV. Tsi values
used are from Figure 6.7, meeting the Ioff requirement. Other parameters as per Table
105
6.1. D IB L is via Eqn. (2.10). .......................................................................................
Figure 6.9: DG-n+p+ long-channel Vt: dependency on Tsi and Tbox, with corresponding
range of DG-Img long-channel Vt indicated, upper right. Analytical are ID Laplace solutions (appropriate in the limit of zero channel depletion charge) for Qi/q=1 x 1011 cm-3 .
10 6
......................................................................................................................................
Figure 6.10: Simulation results: effect of lateral S/D doping gradient on Tsi required to
meet Ioff! 1x10- 8 A/gm. Shown versus scaling nodes of Table 6.1. Simulation grid spacing is 0.6 nm at transitions from S/D to channel. Tox and Vdd vary with Leff according to
10 7
T ab le 6 .1. .....................................................................................................................
Figure 6.11: SG-Dmg MOSFET, interpolated results from Figure 6.1: Tsi required per
scaling node of Table 6.1. Tox and Vdd vary with Leff according to Table 6.1. .......... 108
Figure 6.12: DG-n+p+ MOSFET, interpolated results from Figure 6.5: Tsi required per
scaling node of Table 6.1. Tox and Vdd vary with Leff according to Table 6.1...........109
Figure 6.13: DG-dmg MOSFET, interpolated results from Figure 6.6: Tsi required per
scaling node of Table 6.1. Tox and Vdd vary with Leff according to Table 6.1. .......... 109
Figure 6.14: Energy balance simulation results for Ion and Ioff for the three FDSOI architectures. For solid symbols, lumped resistances were added to the source and drain (80
Qgm per side). Total Rsd for solid symbols is 190-200 Qitm; for open symbols, 30-40
17
Qsm. Tsi values chosen to meet Ioff criterion(Figure 6.2-6.7); other parameters are for the
50 nm node in Table 6.1. rw=0. 15 ps, Lombardi mobility model. .............................. 110
Figure 7.1: Three fundamental structural possibilities for DG-FDSOI MOSFETs; illustration from [77 ]............................................................................................................114
Figure 7.2: Illustration of the SOIAS flip-and-bond process, applied to double-gate MOSFET fabrication . ............................................................................................................
115
Figure 7.3: Illustration of the IBBI alignment scheme, applied to double-gate fabrication
via X -ray lithography. ..................................................................................................
117
Figure 7.4: Test double-gate structures fabricated via pre-patterned flip-and-bond, with
IB B I alignment. ............................................................................................................
117
Figure 7.5: Sacrificial SiNX spacer / selective epitaxial raised S/D process. (a) Gate stack
is etched using LTO (low-temperature oxide) hard-mask, followed by re-ox (not shown)
and S/D arsenic implants. (b) LTO liner and SiNx LPCVD deposition (13 and 90 nm respectively), followed by anisotropic spacer RIE etch in NF 3 :0 2 , stopping on LTO liner.
(c) Pre-epitaxial wet clean (H2 0 2 :H 2 SO 4 followed by HF). (d) Selective removal of
SiNx spacers in H 3PO 4 , followed immediately by UHVCVD selective growth..........118
Figure 7.6: Selective epitaxial raised S/D results, shown for a bulk MOSFET. Epitaxial
layer is 75 nm thick. The 13 nm LTO liner between epi and n+ poly is not visible here,
but its presence can be inferred from the lack of epi-growth on the n+ poly side-wall.
(SEM im age taken at 10000O X , uncoated)...................................................................119
Figure 7.7: XTEM images. (a) Faceting at Si / LTO interface, with 750C UHVCVD
growth via SiH 2 Cl 2 /H2 - (b) Faceting suppressed, w/ 750C growth via SiH 2 Cl 2 /H2 /PH 3 ......................................................................................................................................
1 20
Figure 8.1: Illustration of alternative method for 9gef determination in short NMOS devices, adapted from [93]. For given Lmask, mobility is found from dRtot / dLmask. Inversion charge density Qj is determined from long-device CV integration, corrected for short
device according to Eqn. (4.6). Results are extremely sensitive to the polynomial fit
(here, to degree 3) to the data, suggests that more points are required for reliable results.
Techn ology A . ..............................................................................................................
124
18
List of Tables
Table 1.1: 1999 ITRS Roadmap Parameters .............................................................
23
Table 2.1: Summary of the Carrier Velocity Techniques Investigated........................36
Table 2.2: Parameters for NMOS Technologies Investigated ..................................
41
Table 2.3: Comparison of carrier velocity, ballistic efficiency, and transmission coefficients. Experimental results for technologies A and B, plus Monte-Carlo results for 25 nm
from [2]. 0 = veff/vo = T/(2-T). vo values estimated from [18] as 1.6, 1.7, and 1.9x10 7 cm!
s, respectively, moving from right to left in the table. veff = vidi at DIBL = 100 mV/V. For
Monte-Carlo device, Qj(xO) was estimated from (FOx / Toxelec)(V gs-Vt). As this method of
charge determination was different from the capacitive techniques used for technologies A
and B, some uncertainty was introduced in the comparison, reflected in that range of values
46
for M C results. ..................................................................................................................
Table 3.1: Design N ode Param eters ...........................................................................
55
Table 6.1: FDSOI Design Node Parameters ..............................................................
98
Table 6.2: Normalized Transconductance for Three FDSOI Architectures; Leff = 50 nm
11 1
..........................................................................................................................................
19
20
Chapter 1
Introduction
1.1 Background
For several decades silicon bulk MOSFET microprocessor scaling and performance has
progressed exponentially: transistor gate density and microprocessor speed have doubled
approximately every 18 months [1]. This progress has been made possible by enormous
advances in all areas of semiconductor device fabrication: lithography, thin-film growth
and deposition, reactive ion etching, channel implant engineering, as well as device contact, interconnect and packaging technologies. For digital logic applications silicon
CMOS has long outperformed alternatives, particularly for mass-market applications like
desktop and portable computing, where power consumption and heat dissipation must be
limited. Many past predictions of a slowdown or end to this era of rapid progress have
been based on over-conservative assumptions regarding one or another of these fabrication technologies. However, in recent years research MOSFET devices have been fabricated which may be approaching more fundamental limits: it has been projected that both
band-to-band and gate-oxide tunneling will prevent channel length scaling below -20 nm
[2].
Electrostatically sound devices with gate lengths reaching down to 25-30 nm [3][4]
have been recently demonstrated.
Although industry projections, as found in the 1999
International Technology Roadmap for Semiconductors (ITRS) [1], suggest devices of
21
these dimensions will not be in production until 2011-13, we appear to be approaching the
"end of roadmap" regime.
This work focuses on understanding the limitations to carrier velocity in MOSFET inversion layers as channel lengths are scaled well below 100 nm, and on relaxing these limits
through architectural alternatives to bulk MOSFETs.
Velocity and Scaling in MOSFETs: the Roadmap Perspective
For a given gate capacitance and gate overdrive, drain current corresponds directly to carrier velocity at the source-side of a MOSFET channel. From the 1999 ITRS, which presents estimated targets for many aspects of device fabrication that will be required to
continue the trend of exponential performance increase, target carrier velocities can be
derived. Figure 1.1 represents effective carrier velocity veff for an "on" device (Vgs = Vds
=
Vdd), as required for NMOS, derived from roadmap targets for Tox and Vdd via the fol-
lowing relationship, as in [5]:
veff =
=on
Ion
(1.1)
i Cox(Vdd - Vt)
Where Qj is inversion layer charge, C'ox is the normalized gate-oxide capacitance, Vdd is
power-supply voltage, and Vt threshold voltage. Relevant roadmap parameters are listed
in Table 1.1. The roadmap gives a range of values for Tox and Vdd for each generation; the
thicker values were chosen for Tox for this this comparison. For Vdd low values, corresponding to low-power, were chosen.
Effect of unscaled threshold voltage, unscaledgate oxide thickness
The 1999 ITRS roadmap does not list Vt per generation, instead giving Vt = 0.4 V (for
long-channel devices) as a guideline.
This is a reflection of the fact that without innova-
tive techniques to manage static power dissipation, Vt cannot be scaled concomitantly
with Vdd while simultaneously maintaining acceptable Ioff.
As is evident in Figure 1.1,
unscaled Vt places high demands upon carrier velocity for the shorter roadmap generations.
22
0 Vt, Tox scaled
U Unscaled Vt
A Unscaled To
0
0 Unscaled Vt, Tox
1-2
E 10
4-
-
_6 6 0
01
100
70
180
130
ITRS Node (nm)
Figure 1. 1: Effective electron velocity veff required to meet 1999 ITRS Roadmap targets
(NMOS). Unscaled T0 x: > 1.5 nm. Unscaled Vt: 0.25 V. Vt values used (Table 1.1) correspond to short-channel devices in each technology, so are well below roadmap guideline of 0.4 V.
TABLE 1.1
1999 ITRS ROADMAP PARAMETERS
YEAR
1999
2002
2005
2008
NODE (nm)
180
130
100
70
NMOS Ion (pA/pm)
750
750
750
750
Lgate (nm)
140
85
65
45
Vdd (V)
1.5
1.2
0.9
0.6
Toxelectrical (nm)
2.5
1.9
1.5
1.2
Vt (V)
0.25
0.25
0.2
0.2
In addition, the SiO 2 gate insulator can be scaled only so far before leakage current due to
tunneling becomes unmanageable. Predicted limits range from 2 nm (for equivalent electrical thickness in inversion Toelec, which includes gate-poly depletion and non-zero
inversion layer depth) to 1 nm (physical thickness) [6][7][8]. If an acceptable high-k gate
dielectric process cannot be developed, the most aggressive roadmap targets for To, elec
23
may not be practical. For Figure 1.1, to demonstrate the additional demands that unscaled
T0 Xelec will place upon carrier velocity, a limiting minimum thickness of 1.5 nm was chosen. Alternately, achieving high carrier velocity in deeply scaled MOSFETs can relax
other challenging roadmap targets.
Velocity and Scaling in MOSFETs: Device Physics Perspective
To understand what may happen to carrier velocity with continued device scaling, we
must consider two perspectives on limits to carrier velocity:
The "tyranny of universal mobility"
Figure 1.2 illustrates MOSFET fields and potentials. Vdd is not expected to scale proportionally to channel length, so that lateral electric field
IEx| will increase for shorter devices:
this would tend to improve veff with scaling. However, for electrostatic integrity the gate
bias must have much more control over the potential at the source-side of the channel than
drain bias; otherwise the device cannot be turned off fully via zero gate bias. This requires
JEyJ >> ExJ near the source for a device in saturation (gate and drain bias high), where EY
is transverse electric field. High Ey leads to degraded carrier mobility: it is well known
that MOSFET effective mobility peff has a universal relationship with Ey, decreasing as
EY increases [9]. Because of this "tyranny of universal mobility" (geff degraded by scaling) in bulk MOSFETs it is not clear whether further scaling will improve, or degrade,
veff. This is difficult to predict theoretically, since traditional device analysis based on the
Drift-diffusion (DD) or energy balance (EB) approaches may not be physically justifiable
in this "end of roadmap" regime, primarily because EX and potential can change significantly over one carrier scattering length [10].
The ballistic or "thermal limit"
It has been suggested that advanced MOSFET devices may in fact be approaching the fundamental limit of ballistic transport [11][12] (zero scattering events for electrons), where
veff reaches its maximum possible value, determined by the thermal velocity of carriers
24
Vgs
V
Vds
-Ex
EX
V S,
~0S
Vd
X
SLCH
Figure 1.2: Potential profile and electric fields in a bulk NMOSFET channel, for Vgs =
Vds = Vdd. For acceptably low DIBL (Drain Induced Barrier Lowering) IEy must be significantly greater than IEJI.
entering the channel from the source [14]. In this case DD and EB approaches are clearly
inappropriate.
Goals of This Work
The two primary goals of this work are:
(1) To further the experimental understanding of the behavior of veff and peff in deeplyscaled bulk MOSFET channels. To this end, we investigate NMOS devices from two
advanced deep-sub-100-nm CMOS technologies.
(2) To examine both design-space and fabrication technology for alternative MOSFET
architectures that have the potential for scaling without high transverse electric fields: single and double-gate Fully Depleted Silicon-on-Insulator (FDSOI) MOSFETs (Figure 1.3)
may allow scaling without a mobility penalty.
25
Double Gate
Single Gate
LG
I
I
Top Gate
Si02
LG
I
II
T4=
Si
Channel
Si02
Bottom Gate
Figure 1.3: FDSOI MOSFET Architectures
1.2 Outline
In Chapter 2, we investigate experimentally (and with the aid of 2D simulation) electron
velocity in advanced NMOSFETs to see how near they operate to the thermal limit.
In Chapter 3, we investigate via 2D simulation the relationship between scaling and lowfield mobility for bulk NMOSFETs as well as for FDSOI alternatives.
In Chapter 4, we explore experimentally the effect of scaling on mobility in bulk NMOSFETs, to see if the universal relationship of 9ef with EY holds in heavily doped deep-sub100-nm channels.
In Chapter 5, we explore experimentally the relation of low-field mobility to the carrier
velocity of NMOSFETs in saturation: mobility change is induced via mechanical strain,
and the corresponding mobility and velocity shifts are correlated.
In Chapter 6, we explore via 2D simulation the design-space for single- and double-gate
FDSOI MOSFETs, focusing primarily on scalability and transport.
26
In Chapter 7, we investigate two novel fabrication technologies which may make deeplyscaled FDSOI MOSFETs feasible: gate alignment via Interferometric Broad-Band Imaging (IBBI) and selective epitaxial raised source and drains.
Chapter 8 summarizes results and suggests future work.
27
Chapter 2
Experimental Determination of Carrier
Velocity in Deeply Scaled NMOS
2.1 Introduction
Continued success in scaling bulk MOSFETs has brought increasing focus on fundamental
performance limits. It has been proposed that drain current is ultimately limited by the
rate at which carriers can be thermally injected from the source into the channel [12]. In
this chapter we show that commonly used techniques for experimentally determining carrier velocity are insufficient to determine how close modem MOSFETs operate to the ballistic or "thermal limit". We propose a new technique, and apply it to two advanced
industry technologies with deep-sub-100-nm channel lengths. We show that an advanced
1V NMOS technology with Leff < 50 nm operates at no more than -40% of the limiting
thermal velocity. Furthermore, we find no indication that continued scaling is bringing us
closer to the thermal velocity limit.
28
2.2 The Thermal Injection Limit
It has been shown that electrons in a MOSFET channel can exceed significantly the saturation velocity for isotropic field regions vsat (~1x10
7
cm/s) [15][17]. The ultimate limit
to performance is, rather, thought to be the thermal injection velocity vo (1.2-2x,0
7
cm/s)
from the source accumulation layer into the channel [14][18]. Applying the formalism of
1-flux scattering theory [14] this can be stated as:
Ion / W = vo Qi(xo) T/(2-T)
(2.1)
where Ion is the saturated drain current and Qi(xo) is the areal inversion layer density at the
conduction-band peak (with Vgs=Vds=Vdd) at the source-side of the channel. The effec-
tive carrier velocity at xO is:
veff = v 0 T/(2-T)
(2.2)
T is the transmission coefficient at x0 . T=1 (and veff = v0 ) represents the thermal limit:
transport is fully ballistic, and there is no backscattering of carriers to the source. This
viewpoint is illustrated in Figure 2.1. As a measure of how close to the thermal limit a
device operates, it is conceptually useful to define a thermal or ballistic efficiency
$ veff/vo = T/(2-T)
1:
(2.3)
To estimate T and $, veff must be determined experimentally (vo can be estimated theoretically [18]). The point xo, where Eqs. (2.1-2.3) are evaluated, is identified with the conduction-band peak (where transport is purely diffusive) because v0 is the limit of diffusive
velocity only: moving toward the drain carriers are accelerated (to well beyond v0 eventually) as
Q1 decreases,
regardless of P. This is illustrated schematically in Figure 2.2.
Monte-Carlo simulation results emphasize this point, showing that carrier velocity
increases dramatically along the MOSFET channel (Figure 2.3). This demonstrates the
necessity of determining veff as close to x. as possible, if the goal is to assess how near to
the thermal limit a modern MOSFET operates. The answer to this question has important
ramifications: large $ for a modern "standard" MOSFET would suggest only minor drivecurrent benefit could be expected from continued scaling, or from technology alternatives
for mobility improvement (e.g. strained-Si or undoped thin-film SOI).
29
EC
I I
I
0 0
LCH
-- ,
--
X
Figure 2.1: Carrier velocity considered at source-to-channel barrier. Unless transport is
ballistic, average carrier velocity (veff) at the source-to-channel barrier must be less than
the thermal injection velocity v0 For deep-sub-100nm devices conduction-band peak is
typically within several nm of source-channel transition (x=O, defined as the point where
source doping falls to 2x10 19 cm- 3 [19]). Note that simulations for bulk MOSFETs indicate a conduction-band peak much less pronounced than shown in this schematic.
Vs
VD
VG
-1
1
Q!
C)
IdIW = QiVe
0
LCH
= constant in x
x
Figure 2.2: Illustration of relationship between carrier velocity and inversion-charge density along MOSFET channel. Since there is no single ve for the channel, different extraction techniques for ve can give significantly different answers.
30
Leff
3
E
00
2
0
,
0
5
10
15
20
25
Channel Position (nm)
Figure 2.3: Monte-Carlo simulation results: electron velocity as a function of position in
an ultra-scaled bulk NMOS channel. Toxphys=1.5 nm. Vgs=Vds ~ 1V. Courtesy of D.
Frank (IBM).
2.3 Investigation of Velocity Measurement Techniques
by Simulation
We explore the physical significance of several experimental carrier velocity extraction
approaches via 2-dimensional drift-diffusion (DD) and energy balance (EB) simulations
(Appendix 1). The goal is to find a technique which measures velocity near the conduction-band peak. We use a "super-halo" MOSFET structure (Figure 2.4) with effective
channel length (Left) of 50 nm, where Left is defined as in [19], by the points at which the
source/drain doping fall to 2x10
19
cm- 3 . Note that the metallurgical junction distance is
much less in this case, 36 nm. The highly non-uniform doping profile is used in advanced
bulk MOSFETs to provide some degree of immunity to short channel effects, reducing
short-channel threshold voltage rolloff [20]. The structure was designed to match closely
31
measured characteristics (Vt, T0 X, DIBL, Leff, gm) of the experimental technologies investigated below (technology A, Section 2.4).
Gate
elec
TT0oxe
2.4
Leff
=50
nm
Source
-E0
Drain
n-type
-
3x1O1 9
1x10 19
p-type
y
1x10
18
Figure 2.4: NMOSFET simulation structure (Avant! MEDICI TM [13]). Mobility models
used in this chapter: (a) for low Ex, concentration-dependent g; (b) for EY, Si-SiO 2 universal p assumed, with coefficients from [13]; (c) for high Ex, either a Caughey-Thomas
relationship [64] for DD simulations, or an electron-temperature dependent model as
described in Appendix 1 for EB simulations. For DD simulations, electron saturation
velocity vyt = 1.3x 107 cm/s. For EB simulations, energy relaxation time TW = 0.1 ps, and
vsat=lxlO cm/s. . From simulation, Toxelec was assessed from the slope of the dQi/dVgs
curve for mid-channel for a long-channel device at Vds=O; slope was taken at Vgs=VddClassical distribution was assumed for inversion layer electrons for all simulations. Leff
is determined by the 2x10 19 cm-3 S/D doping points. Fermi-Dirac statistics. External
lumped resistance of 90 ipm per side, to give total Rsd = 220 Qgm.
Carrier Velocity from Saturated Transconductance
Effective carrier velocity can be measured from extrinsic or intrinsic saturated
transconductance gm and gmi [15]:
vgm = gm / WC'0
X
(2.4)
vgmi = gnm / WC',x
(2.5)
32
C'10
is the gate oxide capacitance per unit area in inversion. gmi is saturated
transconductance corrected for source/drain parasitic resistance (Rsd) as in [16].
As
carrier velocity is inherently an intrinsic quantity, vgmi should be a more accurate
reflection. Nonetheless vgm is a useful estimate when Rsd is not known. According to
[21] vgmi is related to carrier velocity ve at the point in the channel where ave/aVgs= 0;
EB simulations show that this occurs 10-30 nm beyond the source for a 100 nm device.
Repeating the simulations in [21] for deep-sub-100-nm devices, we find that this distance
(relative to channel length) becomes more significant, occurring at a point almost 45%
across the channel. This is shown in Figure 2.5. Note also that aQ/ aVgs at this point so
far into the channel can not be approximated by C'x. Thus, as devices are more deeply
scaled, vgm and vgmi become less appropriate as a measure of velocity at the conductionband peak (x = xO).
Simulated results of this approach are compared with other
approaches, below.
e
o
Sgs
0
10
30
20
40
50
Channel Position (nm)
Figure 2.5: EB simulation results for the device of Figure 2.4. Investigating components
).
of transconductance as a function of position in channel. (Avant! MEDICI
33
Carrier Velocity from Drain Current and Long-Device CV
Conceptually, a more straightforward way to extract veff is to directly measure Io.
WQi(xo). Determining Qi(xo) in deeply scaled devices is problematic due to uncertainties
in channel length, large (relative) overlap and fringing capacitances, and non-uniform
charge distribution along the channel. However, in strong inversion and in the gradual
channel approximation (IEy >> JExJ, which is true for electrostatically-sound bulk
MOSFETs), Qi(xo) in a short channel should correspond closely to the long-channel
inversion layer charge Qo=fOCgsd Vds=O, where C'gsd is the gate-to-source/drain (tied)
capacitance, normalized to unit area. This assumption is examined further in Section 2.5.
Accordingly we let:
Qi(xo)(short-chan.) = fVgs*Cgsd Vds=O (long-chan.)
(2.6)
where Vgs* = Vgs + AVgs with AVgs accounting for differences between long and short
devices. The most important component in AVgs is AVt due to DIBL and thresholdvoltage rolloff. Adjustment due to AVt is illustrated in Figure 2.6. The expression for
effective velocity as extracted from Ion becomes:
vid = Ion / WQi(xo)(short-chan.) =Ion / (WfOtygs+A> C'gsd (long-chan.))
(2.7)
A second component in AVgs is due to voltage drop on the source resistance, IonRsAdding this correction to the upper integration limit in Eqn. (2.7) gives the expression
used for "intrinsic" effective velocity, vidi:
vidi = Ion
/
/(W
(W
C1
Vgs + AVt-IRs)
C gsd (long-chan.))
js+
(2.8)
The extraction techniques for vid and vidi were developed for this work. Simulated results
of this approach are compared with other approaches, below.
Carrier Velocity from Drain Current and Short-Device CV
In [22] a carrier velocity extraction technique is presented, which we denote vid2:
Vid2
I
vgs C'gs (vds=vdd))
C'gs is the capacitance (per unit area) measured from the short device.
34
(2.9)
The method
W=L=10gm
Vds = 0
AVt
gs
1.5 -
S
1.0
QiO
AQi
0.5
0
0.5
0
1.0
Gate Voltage (V)
Figure 2.6: Example of Qi(xo) determination. Capacitance is measured on large device.
For large device, overlap C is inconsequential, so area under curve corresponds to inversion layer charge. Integration limit is adjusted by AVt (= DIBL + Vt-rolloff - IonRs) as
determined from the target short device. Qi(xo) = (Qio+AQi) / (W*L). NMOS.
requires no Rsd correction, as the Qi obtained from integrating C'gs in saturation is
intrinsic. In addition, no correction is required for short-channel effects (as with vid and
vidi) since IV and CV are measured in the same device. From a practical standpoint, this
technique is difficult to apply accurately because knowledge of Left is required to
normalize Cgs to unit area. In addition, for a short device fringing capacitances are a
much larger relative contribution to Cg, requiring more correction. From a device theory
standpoint, since Cgs is measured in saturation, integrating gives a value of Qi which is an
average along the channel
Since this is much less than Qi at xO (the
(Figure 2.7).
conduction-band peak), vid2 will be much higher than veff at xO.
Simulation Results: Comparison of Techniques
Simulation allows investigation of quantities unobtainable by experimental methods, such
as potential, and carrier velocity and density, as functions of position along the channel.
35
-
I
2
E
0
Vd,=lV
V ,=OV
CO.
Vds=Vgs=lV
1
/
x
Qlong-channel
1r
I(Vds=O)
0
10
50
40
30
20
Channel Position (nm)
Figure 2.7: EB simulation for the device of Figure 2.4: area between curves gives Qcgs,
which corresponds to Qj as determined in Eqn. (2.9).
Name
Ref.
Measurement
Vgm
gm / WC'%x
[15]
Vgmi
gmi / WC'ex
[15]
Vid2
Ion / W
vS
C
vt
[22]
VdsVdd
U
d~d
Vid
Ion/ WJ(Vgs+ Avt)ctgsd
f0
Vidi
Ion/ W(Vgs + AVt - Ion Rs)
Eqn. (2.7)
I
Eqn. (2.8)
0
TABLE 2.1: Summary of the carrier velocity techniques investigated.
36
Using our 50 nm super-halo simulation structure, velocity values corresponding to five
extraction methods (summarized in Table 2.1) are extracted from the simulated output
characteristics (DD example is shown in Figure 2.8). For vgm and vgmi, C'0 x was determined from the slope of the dQi/dVgs curve, taken at mid-channel for a long-channel
device at Vds=O, Vgs=Vdd. These values are plotted on the simulated carrier velocity profile along the channel, as shown in Figure 2.9. Note that results are relatively insensitive
to model chosen: DD model produces very similar velocity profile for first 15 nm of chanSaturated extrinsic transconductance gm for both models are matched
nel as EB model.
approximately to an experimental device from technology A (described in Section 2.4)
with approximately equal DIBL. For DD model this matching is achieved through adjusting electron saturation velocity vsat; calibration of EB model is through adjusting energy
For both DD and EB models "universal curve" mobility is used (see
relaxation time tc.
Section 3.1). The simulated conduction-band depicted is a cut at 0.4 nm below the SiSi0
2
interface, which is the depth of the inversion-layer centroid at the source end of the
channel.
0.8 o 0 o Tech. A, measured
DD simulation
0.6
E
-E
*
S0.
*
0.2
0
e
0
0.2
0.4
0.6
0.8
1.0
Vds (V)
Figure 2.8: DD simulation results (solid): output characteristics for the device of Figure
2.4, matching saturated gm and Ion of experimental technology A (discussed in Section
2.4).
37
Ec-peak (x=xo)
(a)
Vds=Vgs=1V
0
3
.Vid
-0.2
AVidi
2
-0.4
SVid2
1
-0.6
0
0
10
20
30
40
5(
Channel Position (nm)
(b)
Ec-peak
I
Vds=Vgs=lV
I
0
3
-Vid
-0.2
AVidi
0 vgm
N-
-
-0.4
2
-.1i
0
Vg~.
-id
1
-0.6
0
10
20
30
40
Jo
50
Channel Position (nm)
Figure 2.9: Simulation results: conduction-band and carrier velocity vs. position for bulk superhalo device of Figure 2.4. Velocities at five marked points along the channel correspond to values given by different MOSFET carrier velocity extraction techniques (Table 2.1). Solid symbols are corrected for S/D series resistance, open symbols are uncorrected. Vth, TOXelec, Rsd, and
Ion approximately match the "Technology A" under investigation. DIBL here is 66 mV/V. (a)
DD model: vsat= 1.3x10 7 cm/s (b) EB model:T,=0.1 ps.
38
These results show clearly that the extraction methods developed for this work (vid, vidi)
correspond to inversion-layer velocities closer to xo than the methods from the literature
(Vgm,
vgmi, vid2). Using any of the latter methods will significantly overestimate ballistic
efficiency
1.
However, vid and vidi also correspond to points in the channel somewhat
beyond xo. We must interpret, then, the subsequent experimental results (based on vigd) as
putting a reasonable upper bound on veff, and therefore
1, for the technologies
investi-
gated.
Carrier velocity ve(x) was extracted from simulation by taking Ion / WQi(x). In the case of
EB simulations, this leads to some overestimation at the drain end, where the inversion
layer is found to "spread out" deep below the surface. Since the integration depth was
fixed (at 5 nm) for the purpose of extracting Qi(x) from the simulation, the deep tail of the
charge distribution was missed at the drain end.
Velocity extraction near the scaling limit
Extending this investigation closer to the projected limit of MOSFET scaling requires
Monte-Carlo simulation. Figure 2.10 shows results for the simulated 25 nm super-halo
device presented in [2].
S/D series resistance of the simulated device is not known accu-
rately; with the available data only an approximate comparison of vid vs. vgm is possible.
For vid in this figure, Qi(xo) was estimated from C'0 x(Vgs-Vt) since CV data was not available. Furthermore, in this case gm is not a differential quantity but a difference: gm = AIon
/ AVgs for one Vgs step. The results support our contention that vic is more suited than vgm
for measuring velocity near xo.
2.4 Experimental Results
Technology Parameters
We measure vid and vidi (and compare with transconductance methods) for NMOS
devices from two advanced CMOS technologies, referred to in this work as technologies
39
1.0
VgS=1.1I V-
.
0.9 V
S0.5
0.7 V-
0.5 V
0
0.4
0.8
1.
Vds (V)
Leff
0
3
-0.4
<
CD
g
-i
2
(D
WI
04
x
0
-0.8
1
C
3
C,,
-1.2
0
0
5
10
15
20
25
Channel Position (nm)
Figure 2.10: Monte-Carlo results: output characteristics and electron inversion-layer
velocity, for a 25 nm Leff NMOSFET design presented in [2]. Courtesy of D. Frank
(IBM). Inversion-layer distribution is classical. Our estimates of vid and vgm shown are
approximate values.
40
"A" and "B" (Table 2.2)*. Figure 2.11 gives example output characteristics. Figure 2.12
gives Vt for the shorter devices of technology A. Corresponding information for technology B is available in [24]. In Chapters 3 and 5 the more scaled technology A will be
examined in greater detail. Toxelec was determined experimentally from
Fox
/
C'gsd (with
Vgs=Vdd) measured in a large (L/W = 10/10 pm) device. Regarding the range of source-
drain parasitic series resistance (Rsd) presented in Table 2.2: for both technologies, the low
estimate is determined from inverse modeling [25][53].
However, at the time of the
investigation summarized in this chapter, only data from the shift and ratio technique [23]
for was available for technology B. This higher estimate (270 i-
m) was used for the
subsequent analysis. The inverse-modeled result for "A" (190 i- tm) was considered to
be preliminary, and somewhat low compared to the best reported results in the literature.
For this reason a more conservative estimate of 220 0-pm was used here.
The use of
these estimates, that are likely high (by 30 9-gm for both technologies), introduces negligible error for the relative comparisons between measured velocity for "A" vs. "B" to be
shown below. For absolute values of vidi for comparing to the thermal limit, the final values (Table 2.3) presented are small overestimates (1.0-1.5%).
TABLE 2.2
PARAMETERS FOR NMOS TECHNOLOGIES INVESTIGATED
*INDICATES RSD VALUE USED IN THIS CHAPTER
tINDICATES VALUE USED IN SUBSEQUENT CHAPTERS
Technology
T elec
Vt (Vds=
Nominal
50mV linear
extracted,
long chan)
Vdd
Rsd
Leff for DIBL
< 130 mVN
A
2.4 nm
0.3 V
1.0 V
1 9 0 t- 2 2 0 *
-40 nm
B
4.3 nm
0.35 V
1.8 V
2 40t-2 7 0 *
-65 nm
92-gr
* These devices were obtained courtesy of industrial partners.
41
Technology A
Technology B
I
1.2
I
Vgs = 1.5 V 00000000C
0000
000000
E
0.8
Vgs = 1.8 V
000000000000000000
- 00 00*00000*00000
000000*
0000
E
0
00
4-0
C
CD
0000000000000000000000
0.4
.
00000000000000000000
oO
,%0***
0 00**0
0 00 00000000***
00
C
0 00
0 00
0 0
0
0.5 V-^- ^^ ^^0~
0 00000000000000000OOOOC 0
0.6 V
0---------0 000000000o00000000000000
*00000
C:
00
0
0
0
0.5
1.0
1.5/0
0.5
1.0
1.5
Drain Voltage (V)
Figure 2.11: Example output characteristics for NMOS devices from two technologies
under investigation. DIBL is similar for both devices shown, 60-70 mV/V.
40C I
E
0
_0
.c
300 -
300
200-
200u
10 F-
100
~~_~ -----
OF
H/ -100OF
-200
Leff = 3 5 nm
-
--
0
3
- 0
A Vds = 50 mV Lin. Ext.
0 Vds = 50 mV, Icc = 1e-7W/L A
LVs = 1V mV, Icc = 1e-7W/L A
100
500
Log Effective Channel Length (nm)
Figure 2.12: Measured Vt and DIBL for technology A. Results shown are for same set
of devices used for inverse-modeling determination of Rsd and Leff. Leff corresponds to
the points where S/D doping falls to 2x10 19 cm-3 [19]. DIBL is via Eqn. (2.10).
42
Measurements
To determine experimentally Qi(xo) we integrate C'gsd obtained from a large (10x
device, and make adjustments according to Eqn. (2.6).
im)
An example measured CV for
technology A is shown in Figure 2.6. For AVt determination, rolloff was defined as the
difference between linearly extracted threshold voltages (50 mV Vds): compared for the
long (L=10 pim) device versus each short device characterized. DIBL was defined as:
AVt
V t(Vd
AVS
= Vdd) -
t(VdS, =
0.2)
(2.10)
(Vdd - 0.2)
where Vt is determined for a constant current of le-7(W/L) A. C'ox for vgm and vgmi was
determined experimentally from Eox/C'gsd (Vgs=Vdd)- The dependence of Rsd on gate bias
is not taken into account; modest inaccuracies in Rsd do not significantly affect the
results. Experimental results for technology A (Figure 2.13) corroborate the simulation
results, showing relative differences between vid, vidi, vgm, and vgmi of the same order as
shown in Figure 2.9.
It is important to compare velocities of different technologies at
equal DIBL (regardless of measurement technique), because for a given technology
carrier velocity increases as electrostatic integrity decreases [26].
Similar results were
found for the longer-channel technology B but with moderately less spread among the
values (Figure 2.14). That this difference is most pronounced at shorter channel lengths
(either within one technology, or moving from "B" to "A") suggests that with deeper
scaling, effective velocity extraction via an Ion / WQi(xo) method such as Vid or Vidi is
increasingly necessary
How Close?
The thermal injection velocity is a function of channel doping and inversion layer density
[18], increasing with both. Our estimates for vo are taken from [18]. A more detailed
treatment would take into account the non-uniform doping profile and 2D effects at xo, as
well as the non-parabolicity of the bands.
However, the approximate values used are
sufficient in the context of setting a reasonable upper-bound estimate on P.
43
Using
Technology A
1.0
. .gm
E
OVgm
A vidi
0
8
Q0.6 .P
0
C
00
A
AA A,- ,&
- - - - - -
-
-
-
A
AAIL
2
jU
Veff
00
0.4
Vgs=Vds=1 V
0
100
50
150
Drain Induced Barrier Lowering (mVN)
Figure 2.13: Experimental results: carrier velocity by four techniques vs. DIBL, for technology A. Solid symbols are corrected for S/D series resistance, open symbols are
uncorrected. The implicit variable in the x-axis is gate length.
vo=1.7x10 7 cm/s for technology A, and taking veff = vidi from Figure 2.13, we find that
for maximum acceptable DIBL (100 mV/V) $ = veff/vo = 0.39 corresponding to an upper
bound on T of 0.56. Table 2.3 summarizes experimental results for technologies A and
B, as well as for 25 nm (Leff) Monte Carlo simulation results from [2]. Although caution
must be taken drawing general conclusions from comparisons of different technologies
(and Monte-Carlo simulations as well) the results suggest that
1 is
not increasing as we
scale to shorter generations. The significant (15%) decrease in veff from "B" to "A" may
result partly from lower mobility, due both to heavier channel doping and thinner gate
oxides.
In [27] a mobility difference of -15% was shown between 3.2- and 1.5-nm
physical T0 X: thicknesses very close to our estimates for "B" and "A" respectively.
In addition, for this work devices are tested at the technology's nominal Vdd; this means
that Vds / Leff (at DIBL=100 mV/V) is moderately higher for "B". For a more "fair"
44
Technology B
1.0
-
Vgmi
O Vgm
E
AVidi
0
AVid
-
0.8
0
4tt
0
W
C
A
AA
0
X
0
0.6
QA
A
0
A
A
A
A
>
QD
0
0.4Vgs=VdSl
.8V
100
50
150
Drain Induced Barrier Lowering (mVN)
Figure 2.14: Experimental results: carrier velocity by four techniques, for technology B.
Solid symbols are corrected for S/D series resistance, open symbols are uncorrected.
comparison we compare also for constant overdrive (for technology A: Vdd [=Vgs=Vds]
=1V; for B: Vdd=1. 6 5 V), which determines the transverse field. Comparison at constant
DIBL (100 mV/V) determines the lateral field. For these conditions we find the decrease
from "B" to "A" is essentially the same: 14%.
Current drive corresponds directly to
carrier velocity; with continued scaling (to 25 nm) NMOS current drive does not appear
to be significantly above 40% of the thermally limited value.
These results for technology B (P = 0.47) are consistent with results reported for an Lgate
= 130 nm technology in [29], where
f
was estimated at 0.45 by comparing Ion (measured)
with Ion (ballistic-limit from simulation). However, the authors of [29] have reported
P
as high as 0.55 for devices of similar dimensions at DIBL ~ 100 mV/V [30]. Results in
[29] were not given for a specified DIBL.
45
25 nm M.C.
Tech. A
Tech. B
Veff (cm/s) 6.7-7.6x10 6
6.6x10 6
7.7x10 6
0
0.35-0.40
0.39
0.47
T
0.52-0.57
0.56
0.64
-o*--
Decreasing Leff, Tox
Increasing v0
TABLE 2.3: Comparison of carrier velocity, ballistic efficiency, and transmission
coefficients. Experimental results for technologies A and B, plus Monte-Carlo results for
25 nm from [2]. @= veff/vo = T/(2-T). vo values estimated from [18] as 1.6, 1.7, and
1.9x10 7 cm/s, respectively, moving from right to left in the table. veff = vidi at DIBL =
100 mV/V. For Monte-Carlo device, Qi(xo) was estimated from (sex / Toxelec)(Vgs-Vt).
As this method of charge determination was different from the capacitive techniques
used for technologies A and B, some uncertainty was introduced in the comparison,
reflected in that range of values for MC results.
Dependence of veff on Vdd
In order to separate effects of Leff scaling from the corresponding changes in lateral electric field, the above experiments were repeated for different Vdd (=Vgs = Vds). Figure 2.15
summarizes the results. It is significant to note that, beyond the lowest Vdd node, for each
technology there is relatively little increase in carrier velocity with increasing drain bias.
One possible explanation is the "tyranny of universal mobility" touched on in Chapter 1.
From the standpoint of carrier velocity, there are competing trends corresponding to
increasing Vdd: decreasing mobility (because of increasing transverse field EY due to
increasing Vgs) versus increasing lateral field EX.
Another explanation is offered by the
scattering theory approach to estimating MOSFET drain current [14]. As noted in Section
2.2, scattering theory relates Ion to vo by Eqs. (2.1-2.3):
Ion = W v0 Qi(xo)
@
(2.11)
Scattering theory makes the further assumption that for modem short devices at high drain
bias, to first order
1 is determined
only by the backscattering of carriers within a distance
46
( of the source, determined by Ec(C) - Ec(peak) = -KT. In this view, for a MOSFET in
strong inversion and with high drain bias, ( and hence I0. is only weakly dependent upon
lateral field and drain bias [28].
E
1.0
NMOS -
Technology B
C
0.8
r-
0
0)
0.6
Technology A
0
0.4
III
0.5
1.0
1.5
2.0
Vdd = Vgs = Vds(V)
Figure 2.15:
Experimental results: effective electron velocity measured from vidi.
DIBL=100 mV/V in all cases. For technologies A and B, as a function of Vdd.
Although it would seem of interest to try to separate the Ex and EY dependencies of veff by
examining veff as a function of Vds at fixed Vgs, this approach would be limited by the fact
that lateral field should not increase significantly for Vds > Vgs -Vt.
2.5 Further Investigation of the Inversion-Charge Relationship Between Long and Short Devices
Through simulation we further explore the physical significance of our value of Qj(x 0 ),
the estimated inversion-layer charge at the conduction-band peak for a short device.
In
our velocity extraction methods we compensate for short-channel effects and series resistance; nevertheless it may not be appropriate to use long-channel low-Vds CV to estimate
47
short-channel high-Vds Qi, for two reasons. First, 2D effects significantly modify the
electric field configuration. Second, 2D doping profiles can be significantly different due
to the use of halo doping. If this is not physically justifiable, the importance of our methods of velocity extraction (vid, vidi) rests only on the fact that they give estimates of velocity closer to the conduction-band peak than other methods.
First we look at long- vs. short- channel capacitance, in Figure 2.16. We compare (a) simulated long-channel capacitance (C'long-inv = dQi/dVgs at mid-channel of a long device
with zero Vds) with (b) simulated short-channel incremental capacitance C'short-inv(x) =
dQi(x) /dVgs (at the operating bias point, i.e. with Ion current flowing).
For the short
device, the structure of Figure 2.4 is used; the long device has super-halo doping identical
to that of the short device near the source and drain, but with the channel more lightly
doped far from the source and drain.
We see that even at x=0, C'short-inv (dashed line)
does not match the long-channel value. However, given the basic relation for inversionlayer charge in a MOSFET:
Q (x) = C'ox[Vgs -Vt - (Ds(x)]
(2.12)
(where Os is surface potential) the incremental capacitance along the channel will be
dQi(x) =
O
dVgs
ds(x)
(2.13)
dVgs
At the source end of the channel, dcDs/dVgs depends upon the potential drop across the
source resistance. This suggests that nonzero source resistance may be why C'short-inv(x=o)
< C'i ong-in,.
Re-doing the simulation with no lumped resistors to account for extended
source/drain regions reduces the total simulated Rsd down to about 40 igm, and we get
the solid line in Figure 2.16. As would be expected in the gradual channel approximation,
the gate capacitance at the source end of a short device approaches the long-channel gate
capacitance value. This suggests that there is in fact a physical basis for our estimation of
Qi(xo) for the short device from capacitance measurements of the long device.
Nonetheless, when in simulation we determine Qi(xo) according to Eqn. (2.6) we do not in
fact get a value that corresponds to x=x0 (the simulated conduction-band peak). This is
48
X0
midchannel
C'long-inv
C\1
El
IL
dQi = C'rt-n
dVgs
(short channel
Rsd = 220 Qiprm)
.0--..i
dVgS
(short channel
Rsd = 40 2tm)
0
20
10
0
Channel Position (nm)
Figure 2.16: EB simulation result for the device of Figure 2.4: Incremental gate capacitance vs. channel position, with Vgs = Vds = 1V. Compared with and without lumped
external resistance, to long channel value C'iong-inv (Vds = 0). xo, indicated above, is
position of Ec-peak from simulation at Vgs=Vds=lV. (Determined from a lateral cut
through the device, at approximately the inversion layer depth).
seen in Figure 2.17. The dashed line is Qi(xo) derived from simulated CV measurements
of a long channel (10 gm) device at zero Vds, then adjusted for short-channel effects and
series resistance via Eqn. (2.6). The solid lines are Qi(x) determined, from simulation, as
a function of channel position in a short (Leff = 50 nm) device. It is not apparent from this
figure that Qi(xo) via Eqn. (2.6) corresponds to any special point near the source. However, the picture is complicated by the "extra" carriers that can be thought of as "spilling
over" from the heavily doped source region, present even when Vgs =0. In this case we
have:
Qi(xo)=fsdQI/dVgs
where
Qivo is the channel charge
at x. when Vgs=O.
+ Qivo
(2.14)
Note that xO is defined as the position
of the conduction-band peak with Vgs =Vds=Vdd: when Vgs = 0, the conduction-band peak
49
is further into the channel. Thus when the device is off (Vgs = 0) the charge at the conduction-band peak will be less than Qiv,0 .
x0|
*
midchannel
2
Adjustments:
VgSlV
C~~~j~ __
A
Rsource
COl
1.
C7,
Qixo) by Eqn. (2.6)
a0
20
10
0
Channel Position (nm)
Figure 2.17: EB simulation result for the device of Figure 2.4: Inversion layer charge vs.
channel position for first half of channel. Vds=lV. Upper-right insert shows size of corrections made to long-channel Qi to obtain Qj(xO) via Eqn. (2.6). Short-channel correction increases, and Rsource correction decreases, Qj(xO).
Relevance to determinationof P
The presense of these "extra" carriers, not induced by the gate, make our experimental
method of determining Qj(xO) via Eqn. (2.6) an underestimation of the actual inversioncharge at xO. This is the reason that our extracted velocity vid is an over-estimate of
velocity at xO, as seen in Figure 2.9.
If taken literally, the results of Figure 2.9 and 2.17
suggest the experimental viad could overestimate the true value of velocity at the conduction-band peak by
-
70%. This means that the actual ballistic efficiency of technology A
may only be approximately 25% versus the experimental value of 39%.
However, EB simulations, with classical vertical distribution of inversion charge, are
inadequate to accurately make this assessment. To further demonstrate one aspect of the
50
'
5 nm
2.5 nm
Vds=Vgs=1 V
Integration
Depth
w
1-D cut approx.
at min. inversion
centroid depth
(method used in this study)
20
10
0
Channel Position (nm)
Illustration of three methods of determining conduction-band potential with
Figure 2.18:
position: top two are weighted averages of electron potential, to the depth specified. Bottom is
simple 1-D cut along the length of the device, at 0.4 nm. DD simulation, for structure of Figure
2.4; only first half of channel is shown.
problem: Figure 2.18 demonstrates the uncertainty in determination of xo. Since inversion
charge varies strongly with position at the beginning of the channel (according to our simulation results) differences of a few nm in xo will affect critically the actual value of $,
since D oc1/ Qi(xo) from Eqs. (2.1-2.3).
2.6 Conclusion
We have demonstrated a technique for measuring effective carrier velocity near the
source-side of the MOSFET channel, which is more appropriate than existing techniques
for the purpose of determining how close modem technologies are to the thermal
(ballistic) limit. With this we have shown that a deeply-scaled (Leff < 50 nm) IV NMOS
technology operates, at most, at 40% of the limiting thermal velocity. Based on our
analysis of two sub-100 nm experimental technologies plus Monte-Carlo simulations near
the bulk scaling limit, we conclude that continued bulk NMOS scaling should not bring
us closer to the limit of source-to-channel carrier injection.
51
Chapter 3
Mobility and Scaling I: Universal Mobility in Bulk and FDSOI MOSFETs
3.1 Introduction
We investigate via simulation the relationship between mobility and scaling for bulk
NMOSFETs and FDSOI (Fully-Depleted Silicon-On-Insulator) alternatives, focusing on
the 50 and 25 nm Leff generations. Scaling of bulk MOSFETs well below 100 nm Leff
requires heavy channel doping, leading to degraded low-field mobility. Provided that the
gate workfunction is used to determined the threshold voltage, FDSOI devices do not
suffer from this trade-off, by virtue of the fact that their channel can be undoped. We
show that single-gate FDSOI is, accordingly, an attractive alternative down to 50 nm Leff.
For deeper scaling, double-gate FDSOI should have approximately a 3X mobility
advantage over bulk NMOS.
Universal Mobility
In deriving analytical expressions for MOSFET drain current, it is convenient to define an
"effective" mobility as a weighted average through the inversion layer:
52
1-
fo i gn(y)dy
eff =
.(3.1)
fo n(y)
where yj is inversion-layer thickness and go is low-field mobility: that is, with lateral electric field EX << Esat (electric field for velocity saturation in an isotropic field region). In
bulk-Si and SOI MOSFETs egf behaves according to a "universal" relationship depending
only on Eeff, the effective or average transverse field seen by carriers in the inversion layer
[9][31][37]:
-
Eeff
=
(3.2)
0
Eeff
(1Qi+ Qb)
(33)
Esi
where po, Eo, v and q are fitting parameters which depend on carrier type. ("Low-field"
in this context refers to lateral, not transverse, electric field). 1 = 1/2 for electrons. Qj and
Qb are the inversion and channel depletion region charge areal densities, respectively.
Figure 3.1 illustrates with experimental results from [37]. Note that Eqn. (3.3) applies to
FDSOI only in the case of Vgb = 0. Typically for channel doping heavier than 2-3 x 1018
cm-3 , Qb becomes the dominant contribution to Ecif.
Although lateral electric fields in a modem MOSFET are far in excess of Esat, theoretical
[32][14] and experimental work [33][34], as well as our own work in Chapter 5, suggests
that carrier velocity (and hence drive current) will still depend strongly on gef in the sub100 nm regime.
3.2 Mobility and Scaling in Bulk MOSFETs
Bulk MOSFET Scaling
To maintain electrostatic integrity while shrinking channel length, depletion depth W must
be reduced [35]. We investigate two basic strategies for controlling W in bulk MOSFETs
53
700
-- o--
600 -
s1=53 nm
tsi=60 nm
E 500
a.
nm
Stsi90
-
400
300
200 V
100
V,0V
'
0
-0.2
0
0,2
0.4
0.6
0.8
1.2
1
Effective Electric Field (MV/cm)
Figure 3.1: Measured NMOS effective mobility for FDSOI and bulk MOSFETs. From
[37]. The deviation for Eeff < 0.6 MV/cm is likely due to overestimation of Qj that can
occur unless Vgs >> Vds [38] when measuring via the split-CV method [52]. Vds = 50
mV. Modem bulk MOSFETs operate at Eeff > 1 MV/cm.
Tstep
xi
I
ON.
I
PP_ I
I Gate:(g
- LeI
Ge
SI Gate:(Dg
I
1I.
15 nm
(b) Step-doped
(a): Uniform-doped
Figure 3.2: Simulation structures for NMOS bulk doping alternatives investigated. Junctions are spatially abrupt. For (a) ( = 4.19 V (for n+ Si). For (b) Dg = 4.19 V and (Db
= 5.25 V (for p+ Si). For Nb and Tstep values by scaling node, see Figures 3.4 - 3.6.
Source and drain are n+.
54
(Figure 3.2) via 2D Drift-diffusion simulations, for four design nodes of decreasing Left.
Table 3.1 lists key technology parameters scaled concomitantly with Left; numbers are
based approximately on ITRS technology nodes [1], with Vdd chosen for Qi/q = 1x10 13
cm-2 at Vt=0.2 V. It is necessary to estimate from simulation (with classical distribution of
electrons in the inversion layer, and no gate depletion) the relation between physical and
electrical gate oxide thicknesses Toxphys and Toxelec. T0 elec was taken as the slope of the
simulated dQi/dVgs curve, extracted at mid-channel of a long-channel device, at Vds=O
and Vgs=Vdd- It is found that across all four scaling nodes, for bulk MOSFETs the correction is in the range of 0.8 - 1.2 nm, consistent with 1-D solutions from [36]. Accordingly,
the physical Tox in simulation is 0.1 nm less than the values shown in Table 3.1.
When assessing scalability, for all device architectures treated in this study we adopt two
criteria, to ensure electrostatic integrity and acceptable loft: (1) DIBL
100 mV/V and (2)
Vt (Vds = Vdd) > 0.2 V. In [20] it is suggested that Vt = 0.2 V is the limiting "worst case"
value--including worst DIBL and rolloff effects for the shortest device in a technology-for MOSFETs with channel lengths below 100 nm.
TABLE 3.1 DESIGN NODE PARAMETERS
Approximate ITRS
Year
Leff
(nm)
Toxelec (nm)
S/D extension Xi
(nm)
Vdd
(V)
2000-1
2002
2005
2008
70
50
35
2.5
1.9
1.5
25
1.2
48
34
26
21
1.35
1.1
0.9
0.75
Uniform channel doping
For bulk MOSFETs with uniformly doped channels, depletion depth W is scaled by
increasing channel doping Nb. Figure 3.3 shows the effect of varying channel doping on
high-Vds Vt for the four scaling nodes of Table 1.1. Interpolating between simulation
points to find Nb required for Vt = 0.2V (and doing likewise to find points of DIBL = 100
mV/V) results in Figure 3.4. A sharp increase in Nb is indicated for scaling to 25 nm Left.
55
Note that for the uniform-doped device criterion (2) is more stringent (requiring heavier
channel doping) except for the shortest node.
0. 14,
0.3-
Lef (nrr
'u
0
50
35
25
0.2
CO,
~0
0.11
0'
18
18.5
19
Log 10 (Nb) cm-3
Figure 3.3: Results for uniform-doped bulk NMOSFET: shown are 2D simulation data
points for V vs. Nb, from which Nb values meeting criteria (2) are interpolated. (Avant!
MEDICI TNI)
Perfect step-doping
A hypothetical alternative is the perfect step-doped bulk MOSFET (also known as the
ground-plane MOSFET [39]), where Nb = 0 down to depth Tstep, and is arbitrarily high
beyond. In this case W is simply Tstep- In practice this could be approximated by epitaxial growth of a lightly doped layer on a heavily doped substrate, or super-steep retrograde
doping. Vt is highly sensitive to Tstep, as shown in Figure 3.5. Interpolating values from
Figure 3.5 gives the step depth required to meet scaling criterion (2) which for the step
doped device is much more stringent than criterion (1), as shown in Figure 3.6.
56
10
Criteria:
c.0
C
o DIBL = 100 mVN
7.5
e Vt (drain hi) = 0.2 V
E
3C?
51
C To
0)
2.51
0
D
0
30
70
60
50
40
Effective Channel Length (nm)
Figure 3.4: 2D simulation results for uniform-doped bulk NMOSFET: doping level (Nb)
correlated with Leff, across four generations. Requisite doping levels to meet two separate criteria are shown.
0.4
0.3
II
0.2
70
0.1
Leff (nn ):
0
10
35 50
2
'
20
15
'
25
'
30
35
Tstep (nm)
Figure 3.5: Results for perfect step-doped bulk NMOSFET: shown are 2D simulation
data points for Vt vs. step depth, from which Tstep values meeting criteria (2) are interpolated. (Avant! MIEDICI TM)
57
I
60-
I
I
1
1
50
60
70
Criteria:
0 DIBL = 100 mVN
50 - * Vt (drain hi)
=
0.2 V
E
40--
0- 30CD
o 2 0 -.............
.4-4
30
40
Effective Channel Length (nm)
Figure 3.6: 2D simulation results for perfect step-doped bulk NMOSFET: step depth
(Tstep) correlated with Leff, across four generations. Tstep required for meeting two separate criteria are shown.
Transverse Field Determination
Eqn. (3.3) will overestimate the Qb contribution to of Eeff for short MOSFETs, due to the
omission of lateral charge-sharing effect between the channel and source/drain. However,
our simulations indicate that the contribution due to Qj does not change appreciably when
2D effects are taken into account.
Therefore, for short bulk-Si devices Eeif can be
obtained from:
Eeff = Esurface - Qi/2Esi
(3.4)
where Esurface is the simulated maximum transverse field in the channel, at the SiO 2 /Si
interface, and Qj is approximated as C'ox(Vdd -Vt), where Vt is the high-Vds value. The
high Eeff that results from scaling is shown in Figure 3.7; note the reduction for the stepdope device at equivalent scaling node.
For the step-dope device, high Eeif is not due to
high Qb but rather to the large built-in potential (CDbi) between the n+ gate and the p+ doping at depth Tstep (NMOS case). Reducing Tstep to meet scalability criteria results in
58
1.8
Bulk, Uniform-Doped
1.6 .
-9
*
Bulk, Step-Doped
1.4-
W
(D
>
E
1.2
1
CZ
>
0 .8 -- - - - - - - - - - - - - 0
0.6-
- - --
Limit for Ob =0, (Ibi = 0
30
20
40
50
60
70
Effective Channel Length (nm)
Figure 3.7: 2D simulation results: maximum effective field Eeif along NMOSFET channel for two bulk doping alternatives, with n+ poly gates, using Eqn. (3.4). Shown vs.
scaling across four generations. Devices are in saturation (Vds = Vds = Vdd)- Qi/q =
C'ox(Vdd-Vt)/q = 1x10 13 cm- 3 was used for all generations. Also shown is dashed line for
Eeif = Qj / 2Esi. This corresponds to zero channel charge and no built-in potential
between top and bottom contacts (Qb= 0, Obi = 0), which is the situation for SG- and DGFDSOI devices with mid-gap workfunction gates.
increased built-in transverse field. This can be understood qualitatively from the relation
for effective transverse field for a long-channel step-doped device: assuming Vgs=O and
zero channel depletion charge,
EOxT0 x + EsiTsi = Fbi
where EOX and Esi are the transverse fields in the oxide and silicon. Since Fsi/EOx ~ 3 and
since EOXEOX = EsiEsi at the Si-SiC 2 interface to conserve electric flux density, we have
Esi = Dbi / (3TOx + Tsi) = (bi / Teq
as the contribution from (bi to the transverse field. Therefore instead of Eqn. (3.3), we
have:
Eeff = TAQ+
Esi
59
TDbi
Teq
(3.5)
Together these two doping alternatives--uniform and step--correspond to the limits of Eeff
for bulk devices of a given generation: for the same Q results for halo and/or retrograde
profiles can be expected to fall in-between.
Mobility versus Generation
Figure 3.8 gives the corresponding universal curve mobilities via Eqn. (3.2): significant
mobility degradation occurs near the projected bulk-Si scaling limit (-20 nm [2]).
However, using the universal mobility relationship appropriate for the Si0 2 -Si interface
to illustrate bulk scaling trends may not be appropriate. Experiments have shown that
nitrided gate oxides improve NMOS gerf at high Eeff [40][41].
This results in a
significant improvement for deeply scaled bulk NMOSFETs (see Figure 4.7). Conversely,
for the high dopings required for extremely scaled bulk MOSFETs, it has been projected
that Coulomb scattering may dominate [42]. In Chapter 4 we explore experimental
evidence for this.
We must conclude that assumptions regarding peff in deep-sub-100-nm bulk MOSFETs
are speculative. This is especially true if we assume that to reach roadmap targets for
T0 Xelec < 1.5 nm, an alternative high-k gate dielectric material will be required (Section
1.1).
Considering the difficulty in demonstrating alternative dielectrics with interface
quality on Si comparable to the Si-SiO 2 system, it may well be that assuming nonnitrided Si0 2 -Si universal curve mobility is not, in fact, pessimistic for bulk MOSFETs.
In light of these observations we consider this a reasonable choice for demonstrating the
relation between mobility and scaling.
3.3 Mobility and Scaling in FDSOI MOSFETs
In fully depleted SOI devices (Figure 1.3) short-channel effects are suppressed by
limiting silicon and oxide film thicknesses.
The deleterious effect of drain bias on
source-side channel potential is limited by device geometry, so that the requirement of
Ey >>IEx (Section 1.1) is relaxed; FDSOI devices thus do not suffer from the "tyranny
of universal mobility." If alternative gate material processes can be developed such that
60
Limit for Qb = 0, (Dbi = 0
300-----------------------------------C,)
E
O
0
2
200-
100--
0 Bulk, Uniform-Doped
* Bulk, Step-Doped
.>
W
0
30
40
50
60
70
Effective Channel Length (nm)
Calculated universal-curve mobilities for two bulk NMOSFET doping
Figure 3.8:
alternatives, correlated with scaling. Values are plotted using Eqn. (3.2) with Eeif values
from Figure 3.7. For Eqn. (3.2), go = 670 cm 2 /Vs, Eo= 0.67 MV/cm, v = 1.6, all according to [31]. Included is dashed line corresponding to the limit of Qb = 0 and Obi = 0,
applicable to SG- and DG- FDSOI devices with mid-gap workfunction gates.
gate work-functions alone set an acceptable threshold voltage, Qb can be essentially zero,
and Beff and gf are decoupled from device scaling. Figures 3.7 and 3.8 include Eeif and
peff corresponding to the limit of Qb=0 and Obi =0, illustrating this potential benefit of
FDSOI. However, the simplest FDSOI implementation, Single-Gate with mid-gap gate
(SG-mg) may not scale beyond -35 nm (for Tsi > 5 nm) because of fringing fields in the
buried oxide (Section 6.2; also [43][39]).
Double-Gate MOSFETs, either with
symmetrical-workfunction mid-gap gates (DG-(cmg) or with asymmetrical n+/p+ gates
(DG-n+p+), do not suffer from this problem [39]; even so it has been suggested that DG
MOSFETs may not scale far beyond bulk [2]. Scalability considerations aside, we find
that SG and DG MOSFETs still have a significant advantage over bulk, a significantly
reduced effective transverse field. For SG-mg, with Qb<<Qi Eqn. (3.3) reduces to Eeff
Q/
2Esi.
For DG-Dmg Eiff is further reduced by half, if
61
Qj is
=
interpreted as the sum of
inversion layers at both gate interfaces: Eeff =
Qj / 4Esi-
This points to an important
difference between transport in DG-mg versus DG-n+p+ devices. Since the latter has a
single inversion layer (at the n+ interface for NMOS), Eeif will be twice as high when
compared with a DG-Img device with the same total
Qj, resulting in a much-reduced
pgJf.
We accordingly find that DG-n+p+ effective mobility is 30-40% reduced versus the DGDmg device, across all four generations investigated.
For ultra-thin SOI films, it may be questioned whether the bulk universal mobility
applies: mobility may be degraded by the presence of a second scattering interface near
the inversion layer. Recent experimental results [44] show that this is a concern when Tsi
<< 10 nm. For the case of Qi/q= 1x10
13
cm- 3 , degradation for a 9.4 nm film (compared
to a control device with 54 nm Tsi) appears to be less than 5%. However, this increases to
-10% degradation for Tsi = 5.2 nm.
3.4 Results
We illustrate the trends discussed above for the Leff = 50 and 25 nm generations,
respectively, in Figure 3.9. In Figure 3.9a, for each of the discussed device architectures
a region of operation is plotted against the universal mobility curve, for the Leff = 50 nm
node. For determining Eeff, a range of
Qj is used:
high and low values of Eeff correspond
to Vdd chosen for higher-performance vs. low-power operation.
For Leff = 50 nm, SG-
Dmg is a viable option, offering a 35-50% mobility advantage over bulk designed for the
same length, depending on the bulk doping profile, and the operating
Qj.
The advantage
for DG-mg (when operated at the same total inversion-layer density as bulk) is higher,
greater than 2X. Moving to the Leff = 25 nm generation (Figure 3.9b) the advantage for
DG-mg over bulk increases to 2.6-3.5X, again depending upon both the bulk doping
profile and Qj.
Recent research has suggested that the mobility advantage of thin FDSOI MOSFETs may
be counteracted by a decrease of gate capacitance (-12%, due to looser confinement of
62
(a) 50 nmn Leff
(b) 25 nm Leff
DG-Img
DG-(mg
CD
cmJ
E
0
0
SG-4mg
.5
DG-(mg*
UD
0
0
0
4-
0
SD
DG:
SG:
SD:
UD:
0
UD
SD
SOl Double Gate
SOI Single Gate
Bulk Step-Doped
Bulk Uniform-
0.5
1.0
1.5
*
double Q,5
0
0.5
0
,
1.0
1.5
Effective Transverse Field (MV/cm)
Figure 3.9: Universal effective NMOSFET mobility (for low lateral field) from Eqn.
(3.2), with regions of operation delineated for four different bulk and FDSOI device
architectures. For each device architecture, the range of Eeff corresponds to: 0.8x10 13
< Qi/q < 1.2x10 13 cm- 3 . In the case of DG-Dmg, Qi is the sum of inversion charge in
both inversion layers, except for special case noted in (b). For DG-Dmg and SG-Dmg,
since Qb = 0 Eqn. (3.3) was used to estimate Eeff, independent of channel length. For
bulk architectures, peak Eeff along channel was determined via 2D simulation, as Esurf Qj/2Es.
the inversion layer), leading to little net gain in current drive [44].
However, the
magnitude of the mobility advantage indicated in our study suggests significant gains in
current drive may in fact be possible for FDSOI MOSFETs, particularly as Leff
approaches 25 nm. This depends critically upon the relation between low-field mobility
pLeff and carrier velocity at high lateral fields, in deep-sub-100-nm MOSFETs.
relation is explored in Chapter 5.
63
This
3.5 Conclusion
The most significant benefit of SG- and DG-FDSOI over bulk may be the ability to
maintain high mobilities with deep scaling. However, to fully realize this benefit it is
essential to use gates with near-mid-gap workfunctions. SG-mg, despite a poorer
prospective scaling limit compared to bulk, has superior transport properties at the 50 nm
Leff generation. Furthermore DG-Dmg has a very significant mobility advantage over
bulk near the prospective bulk scaling limit of 20 nm.
64
Chapter 4
Mobility and Scaling II: Channel Length
Dependence in Super-Halo NMOS
4.1 Introduction
In Chapter 3 we examined via simulation the interrelation between scaling and universal
low-field mobility for bulk and FDSOI MOSFETs. Measurements of mobility in deeply
scaled MOSFETs are difficult, because of sensitivity of extraction techniques to
uncertainty in channel length, inversion-layer charge, and series resistance.
In this
chapter we study experimentally the relation between channel-length and mobility in
NMOS devices in a deep-sub-100-nm CMOS technology. This is made possible by
careful determination of key parameters by inverse modeling and other techniques.
We
find evidence that in the very shortest modem MOSFETs, mobility is less than would be
expected from "universal" mobility, and independent of transverse field.
This may be
indicative of a transition in the dominant scattering mechanism, from surface- to
Coulomb- scattering.
65
Scattering Mechanisms in Deeply Scaled Bulk MOSFETs
Considering a wide range of temperature and transverse field conditions, the dominant
scattering mechanisms for carriers in MOSFET inversion layers are Coulomb, phonon,
and surface-roughness scattering [45][46][47]; we can therefore approximate by Matthiesson's rule [48]:
1
p
_
1
+
Pcoulomb
1
tphonon
1
+
(4.1)
9sr
As discussed in Chapter 3, modem MOSFETs operate with effective transverse field Eeff
on the order of 1 MV /cm and inversion layer density Ni on the order of 1x1013 cm-2 . In
this regime at room temperature the "universal" mobility behavior (Section 3.1) is thought
to be surface-roughness dominated [49].
2.4x10
18
Even for uniform channel dopings as high as
cm-3 , gff for long-channel NMOS at high Ni and Eeff appears to be independent
of doping, suggesting 9sr «< gcoulomb [49]. From Figure 4.1 we can see why this can be
expected to remain true, even for much higher channel dopings. As doping is increased
the corresponding Eeff is increased, increasing surface-roughness scattering and decreasing 9sr and g. Making the assumption that pcoulomb = 9bulk (minority carrier value) we
find that universal mobility (and thus presumably psr) is less than peoulomb for Ni up to
Ix 1019 cm-3 and beyond.
For deeply-scaled MOSFETs this may not be the case. The electric-field configuration is
dominated by 2D lateral charge-sharing effects: much of the channel depletion charge Qb
is imaged in the source and drain depletion regions instead of in the gate. This means that
in short channels, for high channel doping density Nb, Eeff is significantly less than for a
long device with the same Nb. In this case it may no longer be true that gsr << lcoulombAdvanced modem MOSFETs have highly non-uniform "super-halo" dopings to suppress
threshold-voltage rolloff. Doping is highly peaked at the source and drain ends of the
channel, so that as channel length decreases Nb increases.
In light of the 2D consider-
ations mentioned above, it is possible therefore that Coulomb scattering will become dominant in deep-sub-100-nm devices, in which case p would be less than the psr-determined
66
800.
Long-channel limit
600Bulk Limited
400-
Universal
E
0
200
-
1017
1018
1019
Nb (cm-3)
Figure 4.1: Universal vs. bulk-limited mobility as a function of channel doping. NMOS
Universal mobility is from Eqn (3.2) using coefficients from [31], and is based on Ni =
lx1013 cm-3 . Calculation of Eeff is one-dimensional from Eqn. (3.3): lateral charge sharing effects are ignored. Bulk-limited mobility is for electrons in p-region.
universal-curve mobility at the corresponding Eeff. Examination of this issue, through
measurement of the peff-Leff relationship under different conditions, is the primary goal of
this chapter.
4.2 On Experimental Determination of Low-Field
Mobility
Drain current for a bulk MOSFET can be approximated by [50]:
Id = peffC'O
[(Vgs
-
- 2
V t)Vds - Vds
(4.2)
where Lef is as defined in Eqn. (3.1). Eqn. (4.2) is valid for Vds << (Vsb + (Ds) where Vsb
is substrate bias and (s
= 2 0F [51]. For Vds << Vgs, this reduces to:
67
Id = peffC'x
Using the approximation
Qi
(4.3)
(Vgs-vt)Vds
~ C'ox(Vgs-Vt), which is accurate in strong-inversion,
effective mobility can be determined according to:
IdLeff
(44)
V*dsWQi
where
Qi is the
loW-Vds inversion charge areal density, assumed uniform throughout the
channel. V*ds (= Vds - Id Rsd) is the effective or intrinsic drain bias, reduced from the
applied value because of source/drain parasitic resistance Rsd. For a long-channel device,
where uncertainty in Left is negligible and channel series resistance (Rch) is much larger
than Rsd, it is straightforward to evaluate Eqn. (4.4).
For accurate determination of
inversion charge, CV measurements are preferred [52]:
Qi= (1/W Leff )
Cs
s
(4.5)
For advanced bulk MOSFETs with highly non-uniform lateral channel doping profiles,
accurate determination of Rsd and Leff is difficult [23].
In a short device, because Rsd
may be of the same order as Rch, there may be a significant drop of potential across the
source and drain resistance, so that V*ds may be much less than the applied value. In
addition, as mentioned in Section 2.3, determining
Qi
in a short device is difficult,
primarily because of uncertainty in Left. Without accurate Left and Rsd determination,
Eqs. (4.4-4.5) (which together have an Left2 dependency), are essentially useless.
For our subsequent investigation in Section 4.4 we rely on novel inverse modeling
techniques* [25][53] to determine Leff and Rsd; effects of error in these estimates are
carefully considered. In order to reduce the dependency of Eqs. (4.4-4.5) from Leff2 to
Left, we use a relation similar to Eqn. (2.6) to determine short-device
Qi from long-device
CV measurements:
Qi (short-chan.)=(1 / W Left JgrS*Cgsd IVds=O (long-chan.)
*
Inverse modeling results in this and subsequent chapters are from the work of Ihsan Djomehri.
68
(4.6)
where Vgs* = Vgs + AVgs with AVgs accounting for differences between long and short
devices, and also for the potential drop across the source resistance. AVgs = Vt(long-chan) Vt(short-chan)
-
IdRs; for this purpose Vt is the linear-extracted value (Vds=50 mV)
Furthermore, for Vds << Vgs, Qi is not a strong function of position along the channel, so
there is no x dependency in Eqn. (4.6).
Leff in Eqn. (4.6) is for the long-channel device,
so it does not introduce significant uncertainty.
Verification of Equation (4.6)for short channels
As discussed in Section 2.5 (there, for the case of high-Vds), it may be asked whether
long- and short- device Qi can be so simply related. Eqn. (4.6) assumes the strong 2D
field effects in short devices will not modify the inversion-charge density, except through
modulation of Vt. Based on our 2D simulation results, for low-Vds this appears to a wellfounded assumption, even for devices that are too short to be electrostatically sound. We
compare NMOS devices with uniformly-doped channels, comparing mid-channel Qi vs.
Vgs for short (50 nm) vs. long (500 nm) channels at same Tox and junction depth (Figure
4.2).
The simulation structure is as shown in Figure 3.2a. Doping is adjusted to
approximately match threshold voltages for the 50 and 500 nm device (in effect, this is
what halo doping does in real devices: Nb rises with scaling, to maintain approximately
constant Vt). The near-identical Vgs vs. Qi characteristics suggests the approach of Eqn.
(4.6) is valid.
However, in real devices at the shortest channel lengths Vt rolls off and
DIBL >> 100 mV/V despite high doping. This is the case in technology A (as shown in
Figure 2.12) which will be studied via Eqs. (4.4-4.6) in Section 4.4.
To see what to
expect for the very shortest experimental devices, in Figure 4.2 a 35 nm device is also
simulated: Vt is lower but the slope dQi /dVgs remains the same. Therefore as assumed
by Eqn. (4.6) a Vt-shift is the only correction required to make Qi for the 35 nm device
correspond to the long-channel value.
69
1012
1
8 -----.
E
6Leff = 3 5 nm
0
--
Short Channels
Long Channel
44Leff= 5 0nm
2
0
0
Leff=
0.2
0.4
0.6
5 00
0.8
-
nm
1
vgs (V)
Figure 4.2: Simulation results: comparison of mid-channel Qi-Vgs relationship in short
vs. long NMOS devices, with T0 'elec= 1.9 nm and uniform channel doping: Na =
1.25x1018 cm-3 for short devices and 0.6x1018 cm-3 for long. Abrupt junctions are 21 nm
deep. DIBL > 300 mV/V for 35 nm device, 100 mV/V for 50 nm device, and < 10 mV/V
for
the
500
nm
device
4.3 Experimental Results: Long Channel Mobility
We examine the long-channel universal mobility behavior in large (lOxlO gm) NMOS
devices from technologies A and B from Chapter 2 (see Figure 2.11-2.12 for specifications).
Mobility versus Vgs
For technology A, Toxphys is estimated at -1.5 nm. For large devices, this results in significant gate-to-drain and gate-to-source leakage (Ig) for Vgs > 0.5 V, as seen in Figure 4.3a.
To correct Id for this effect, so that it is useful for mobility determination via Eqn. (4.4),
we make the assumption that half of the gate leakage current goes to the drain, and half to
the source. The results of applying this correction can be seen in Figure 4.3b: with corrected -Is and 'd overlaid, a slight divergence is only apparent for Vgs ~ 1.5V. Substrate
70
current was measured, and found to be negligible for the whole range of Vgs. For subsequent measurements of large devices in technology A, the drain current is corrected in this
manner. Technology B does not show significant gate leakage. Applying Eqs. (4.4-4.5)
with Vds=50 mV, results for "A" and "B" are compared in Figure 4.4. Note that using Vds
> 10 mV can result in significant error in mobility extraction unless Vgs >> Vds [38], due
to overestimation of channel charge by Eqn. (4.5). However, in this study we are interested in mobility for Vgs = Vdd, so this error can be neglected.
Mobility versus Effective Transverse Field
Experimental results reported in the literature typically use Eqn. (3.3) to determine Eeff,
the effective or average transverse field seen by electrons in the inversion layer. Channel
depletion charge Qb in this case is determined by:
V
(4.7)
Qb= fy'Cg Vds=O
where Cgb is gate-to-body capacitance, and Vfb is the flat-band voltage [52]. In practice
Vfb may be determined from simulation. Due to lateral charge-sharing effects in short
channels Eqn. (4.7) is only valid for long channels.
We examine an alternative
experimental method for Eeff determination that is applicable to both long and short
devices. Assuming that for NMOSFETs with n+ poly gates there is only a small step (at
high Vgs and low Vds) between the source and channel potentials, we determine Eeif by:
E
S
eff = 3Telec
_
(4.8)
2s
where Qj is a function of Vgs, determined via Eqn. (4.5). Toxelec is also a function of Vgs,
measured from Eox/C'gsd(Vgs). Comparing Eeff determined by Eqs. (3.3) and (4.7) with
the method of Eqn. (4.8), we see in Figure 4.5 that except for low-Vgs (where it is no
longer valid to neglect the source-to-channel contact potential) the measured results are
almost identical for long-channel devices (technology B).
plot of universal mobility for technology B (Figure 4.6).
71
This result is reflected in the
(a)
x 10-
W%
Tech A
5
Vds = 50 mV
Leff= 10 m
4
3
-is
.
s
,'/
S
2
.
-- -~''
- --
1
/
'
0
IL
-6.5
0
0.5
1
1.5
1
1.5
vgs (V)
x 10
(b)
-5
2 ~.,I
2-
K-
-Is+Ig/2 = Id + I92
1.510.51-
-8. 5
0
0.5
vgs (V)
Figure 4.3: (a) Gate leakage current I9 for technology A, with effect on source and drain
currents. (b) Correction to allow Id determination for mobility measurements.
72
Vt
400 -
E""
I~
300
.
(n09000
00
0
>l
Tech. A
C
E
B
"0
0000Tech.
00
200-
0
.0
0
100og
0'
Vds =50 mV
VtLfflm
060~Lef
= 10 pM
0
0
0.5
1
1.5
2
2.5
Vgs (V)
Figure 4.4: Measured effective mobility vs. Vgs for technologies A and B.
1.2
(Qb+Q I i2)!,
1
-
--
Vgs / 3 TOxelec -Q
2Es
E 0.80.6LU.
0.2-
= 50 mV
''Vds
Leff = 10 pm
0
0.5
1
1.5
Vgs (V)
Figure 4.5: Two experimental methods of Eeff determination.
73
2
2.5
500
Vds = 50 mV
'
400-
.
Leff = 10 p
300-
E
2-
200100-
Eeff= (Qb + Q/ 2) /Es
-
1
8.2
-
0.4
-
Eeff = Vgs/ 3Toxec -Qi
1
0.6
-
0.8
1
2Es
1
1.2
Eeff (MV/cm)
Figure 4.6: Measured effective mobility compared for two methods of Eeff determination.
Technology B. Qj and Qb are from split-CV measurements [52]. TXelec = OI
C'gsd(Vgs).
Applying this method as well to technology A, the combined results are presented in Figure 4.7. Also shown is the universal mobility curve for the Si0
2
gate dielectric, from [31],
and an alternate universal mobility curve fitted to results of technology A.
This fit
matches experimental mobility with mobility extracted from simulation (using inversemodeled doping profiles) according to Eqn. (4.4); fitting is for a long-channel device at
low Vds. The experimental technologies have nitrided gate oxides, so are expected to be
superior to Si-Si0 2 universal mobility for high Eeff. Significantly, for high Eeff the mobility in "A" is in fact not inferior to "B", contrary to speculation made in Section 2.4 to
account for the poorer electron velocity in saturation. This apparent contraction is further
motivation to investigate mobility behavior in short devices, and is resolved in this chapter.
74
500,
500
3Universal
400
Cl)
(calibrated to "A")
0000000
300
0
000
0
0
E
0
200
0
Tech. A
Tech. B
100- Vds = 50 mV
Universal
Leff = 10 jim
0'
0.4
0.6
0.8
1
1.2
Eeff (MV/cm)
Figure 4.7: Measured effective mobilities for technologies A and B, plotted along with
universal curve mobilities. For measured data, Eeff is determined via Eqn. (4.8).
4.4 Experimental Results: Mobility versus Channel
Length
Effective Mobility versus Channel Length at Constant Gate Bias
We apply Eqs. (4.4, 4.6) to measure the pe vs. Leff relationship for technology A. For
mobility measurements we use the inverse-modeled set of devices (same wafer and die)
from which the parameters Rsd = 190 Qipm and AL = 65 nm were determined. In Figure
4.8, results are shown versus channel length, at two gate biases. Our Rsd estimate may be
considered on the low end, compared to other reported experimental estimates [54][55].
However, if we assume a higher value in Eqs. (4.4, 4.6) we no longer see the expected
result of geff converging toward the long-channel (10 jim) value as Leff is increased: for
example, assuming Rsd = 220 ipm there is significant "overshoot": peif at 100 nm is
approximately 10% higher than the long-channel peff.
75
In the absence of any physical
explanation for why this would occur, we consider the convergence seen in Figure 4.8 as
some confirmation of the inverse-modeled Rsd. Figure 4.8 shows that 9ef at the shortest
channel length (35 nm) is almost independent of gate bias and hence transverse field Eeff,
suggesting that the universal behavior of Ieff with Eeff has been lifted. It appears that psr
is not the limiting component of gep
for the shortest devices. This will be explored fur-
ther, below.
300-
-etf-ongchan (Vgs = 1V)
Vgs = 1 .0V
250-
E
1.5V
C>--Vs=
200-
F eff-Iong.chan (Vgs = 1.5 )
150-
1001
40
-
60
Vds = 50 mV
80
-
100
Effective Channel Length (nm)
Figure 4.8: Measured peff vs. Leff for different gate biasing. Technology A. Toxelec=2.4
nm. DIBL at Leff = 45nm = 120 mVIV.
On V*ds as a source of error
Since V*ds is sensitive to Id, it is not constant across the range of Leff and Vgs investigated,
but varies by about a factor of two, as shown in Figure 4.9.
For high Eeff and for long
channels the dependence of pgef on V*ds should not be significant, but for shorter channels
this is unclear. In Figure 4.10 we explore this dependency, comparing peff measured at
Vds = 20 mV and 50 mV (a range on the order of the range of V*ds in our eff vs. Leff
experiment). We find that the error introduced is less than 0.3% for the range of Leff
investigated.
76
40
VgS= 1.0v
30-
E 20F-
V gs
=1.5V
Wo
10
Tech. A
C'
40
80
60
100
Effective Channel Length (nm)
Figure 4.9: Intrinsic drain-to-source bias in (= Vds - IdRs) channel for pe measurements
of Figure 4.8. Applied Vds is 50 mV.
1. 0 6 51
1
1
Tech. A
1.061E
0
S
1.055F
a)
1.051.00*
E
1.045-
C).
(,
1.041.0c, '
1%J
5
50
100
150
Effective Channel Length (nm)
Figure 4.10: Measured dependence of short-device peff extraction on applied Vds, for a
range of Leff. Vgs = IV-
77
Effective Mobility versus Channel Length at Constant Overdrive
In order to explore the geff-Leff relationship for several values of constant inversion charge
Qj, we measure geff via Eqs. (4.4, 4.6) at three gate overdrives Vod=Vgs-Vt (where Vt is
the linear-extracted value at Vds= 5 0 mV). This is shown in Figure 4.11. Although constant overdrive does not correspond exactly to constant Qj, in fact it provides a very good
estimate, as can be seen from Figure 4.12, which gives the Qi/q (as determined from long
channel CV measurements, interpreted for short-channels according to Eqn. (4.6).
300 -
'
'od
(V) = 0.8
1.0
250-
200-
Leff (nm~
E
1.451.3
-..
tml.2
150 Vs = 50 mV
100
40
09
60
80
100
Effective Channel Length (nm)
Figure 4.11: Measured geff vs. Leff for different gate overdrive. Technology A. Applied
Vgs is indicated in inset.
Effective mobility at short channels appears to be independent not only of gate bias but of
inversion-charge density as well. This would not be expected in the presense of a strong
screening effect (invesion-charge screening of ionized channel impurities), which has
been invoked in interpreting MOSFET peff data at lower transverse fields [56] [57]. However, for firmer conclusions as to the relevance of carrier screening in deep-sub-100-nm
devices, it will be necessary to look at a technology with stronger halos and/or shorter
78
012
12
1.2
10 81
C
1.0
8-Vod(V)= 0.8
0
p
__r
0
0
6 --
Tech. A
VdS = 50 mV
4
40
60
80
1
100
Effective Channel Length (nm)
Figure 4.12:
Range of Qj corresponding to Figure 4.11.
channel lengths: it is possible that for a shorter device peff would "cross over" and in fact
be higher for higher inversion layer density, as would be expected if (1) Coulomb scattering is dominant and (2) strong Coulomb screening exists (see Figure 4.15, below).
Effective Mobility versus Channel Length at Constant Effective Field
Using Eqn. (4.4) for determination of Eeff, mobility measured for two constants Eeff is
shown in Figure 4.13. This clearly demonstrates the disappearance of universal mobility
behavior at short channels. This behavior, as well as the trend of Jeff degradation (up to
about 30% for the lower Eeff value) are interpreted as evidence that with deep-sub-100-nm
scaling, with heavy halo doping, the dominant scattering mechanism changes: pcoulomb <
psr
However, another scattering mechanism has been proposed for deep-sub-100-nm
NMOSFETs: electron mobility may suffer from long-range Coulomb interactions, not
with ionized channel impurities but rather with electrons in the heavily doped source/drain
and gate regions [58]. In this case the heavy channel doping due to halos might not in fact
be the cause of the observed geff degradation. This is a very significant point, for consid-
79
eration of alternative device architectures as discussed in Chapter 3: if such long-range
Coulomb interactions are dominant, undoped channels will not be a significant benefit.
300 - - III-
0 e0ff-ong chan
(Eeff =
0.8
- - _
MV/cm)
Eeff (MV/cm)
=
0.8
=1.1
250
W eff-long-chan
c
E
200
4
(Eeff= 1.1 MV/cm)
Lef (nm
1.5 -
1.401.2
100
Vds
=
50 mV
40
0.9
60
80
100
Effective Channel Length (nm)
Figure 4.13: Measured gf vs. Leff for different Eeff. Technology A. Eeif is determined
via Eqn. (4.8). Applied Vgs is indicated in inset.
For Iteff at constant Vgs (Figure 4.8) we observe close convergence to the long-channel
mobility limits with increasing Leff, this is not the case at constant Eeff. This suggests our
method of estimating Eeff for short channels (Eqn. 4.8) needs refinement; ignoring the step
in potential between source and channel may be introducing significant error. In addition,
we are ignoring any dependence of Rsd on gate bias.
Technology B
The behavior of technology B with channel-length scaling is not as well understood at this
point, in part because the corresponding inverse modeling results are not as refined. Very
preliminary results suggest that with Leff scaling there is a decrease of dependence of peff
on Eeff--as observed with technology A, but significantly less pronounced. Regarding the
80
work of Chapter 2, it tentatively appears that at the short channels for which velocity was
compared ("A" to "B"), Iteff degradation for technology B is much less significant than for
technology A. Further investigation is warranted.
In Section 2.4 a 14% decrease in electron velocity in saturation was shown for technology
A relative to technology B, comparing at constant Qj and DIBL. For the same gate bias
conditions, and making the simplifying assumption of no Jteff degradation at shorter channel lengths in technology B, the corresponding mobility degradation moving from "B" to
"A" is approximately 25%.
4.5 Investigation of Sources of Error
The results of gef versus channel length presented in this chapter must be considered preliminary, for two primary reasons: sensitivity of experimental peff to error in Rsd and Leff
estimations, and irregularities in results at longer channel lengths.
Effects of Error in Rsd and Leff Estimates
Figure 4.14 demonstrates that error in the Leff used for Eqn. (4.4) affects significantly the
degree of degradation seen at the shortest channel, but does not affect the general form of
the gef
-
Leff relationship: the high- and low- Eeif data still converges, and g
independent of Eeff at the shortest channel length.
remains
However, in Figure 4.15 we see that
using incorrect Rsd in Eqs. (4.4,4.6) leads to more significant error. If we are overestimating Rsd, then the observed convergence at low Leff could be merely an artifact of the measurement technique. We think this unlikely: as mentioned the estimate for Rsd is already
considered on the low end.
Underestimation is more likely: in this case the high- and
low- Eeif lines cross over, as would be the case if carrier screening were a significant
effect.
To be more confident of the Rsd and Leff numbers from inverse modeling, the mobility
model used must be further refined: our inverse modeling up to the time of this work
assumed universal mobility (fitted to technology A for high Eeff [as in Figure 4.7]) for all
81
channel lengths.
The results presented in this chapter suggest that this assumption is
invalid.
(a) Underestimating Leff
3jUU
Eeff(M V/cm)
(b) Overestimating Leff
300
0.8
1.1
- - - - - - 0
250
E
200
-
-
-
U
Long Channel Limits
-
150
---
-
250
.
......
..
..
..
..
200
1501Vds=
100
Tech. A
50 mV
60
40
80
40
100
60
80
100
Effective Channel Length (nm)
Figure 4.14: Effects of error in Leff estimation, i-/- 5nm. Actual estimated AL is 65 nm.
(a) assumes AL = 60 nm. (b) assumes AL = 70 nm.
(a) Underestimating Rsd
3uuTeMV/c-)O.
-
-
--
(b) Overestimating Rsd
-
(0
250-
E
0
200-
-
-
-
- - -
- - -
200 -
Long Channel Limits
150
-
Tech. A
Vds= 50 mV
40
- - -
%f)0
1501
100
.........
250-
60
80
100'
100
40
60
80
100
Effective Channel Length (nm)
Figure 4.15: Effects of error in S/D series resistance estimation, +/- 10%. Actual estimated Rsd is 190 Q-gm. (a) assumes Rsd = 209 9i-gm.
82
(b) assumes Rsd: = 171 nm.
Results for Longer Channels
At medium channel lengths, measured Reff values depart from the long-channel (L = 10
gm) values by up to -10%, as shown in Figure 4.16. In light of this, the assumption that
AL and Rsd are constant across a broad range of channel lengths must be questioned.
Alternately, it is possible that isolation-induced compressive stress is degrading mobility
in some devices, if it can be assumed that this effect varies with differing L/W ratios. Further investigation is necessary to understand this issue.
300
_eff-Io'ng chan (Vg
_
=
1V)
-- -
25----
1.OV
Vg
eff-Iong han (Vg = 1.5V)
0
150 --
100
100
200
300
400
Effective Channel Length (nm)
Figure 4.16: Measured oyff vs. Leff for different gate biasing. Technology A. Txeelec=2.4
nm. Results for points of Leff < 150 nm same as in Figure 4.8.
4.6 Conclusion
We have shown that mobility in the shortest NMOSFETs from a deep-sub-100-nm technology does not behave according to a traditional universal relationship with Eeff. We
interpret this as evidence that Coulomb scattering (perhaps from ionized channel impurities, but possibly from electrons in the source, drain and gate) is limiting the mobility.
However, it is crucial to have an Rsd estimate that is accurate to within a few percent, to
83
make firm conclusions about mobility behavior at these channel lengths. In addition it
would be valuable to repeat these observations in another technology with shorter Leff.
84
Chapter 5
Electron Velocity Dependence on Mobility
in Deep-sub-100-nm bulk NMOS
5.1 Introduction
In Chapters 3 and 4 we have shown mobility degradation with scaling in deep-sub-100-nm
bulk NMOS devices, and have suggested architectural alternatives (FDSOI-MOS) for
improved mobility. However, the relevance of low-field mobility to the performance of
deep-sub-100-nm MOSFETs is not well understood. In this chapter we investigate experimentally, for electrons in short (45 nm) NMOS devices, the relation between mobility at
low lateral electric fields (peff) and velocity in the MOSFET saturation regime (veff),
where peak lateral fields in the channel are high.
Mobility is modified by externally
applying uniaxial stress to devices. The corresponding shifts in electron velocity are
found to be significant, and are accounted for in energy balance modeling. Understanding
the geff-veff relationship is essential for interpreting trends of velocity with scaling
observed in Chapter 2.
Previous experimental investigations [33][34][59] of the geff-veff (or gewf-gm) relationship
have employed alternative NMOS structures with lightly-doped channels and gate oxides
85
in the 4-6 nm (physical) range; the observed geff dependencies may not apply to deepsub-100-nm NMOS devices. We know of no such investigations performed on devices
that are "well-tempered" for channel lengths well below 100 nm.
5.2 Experiment
Mobility with Uniaxial Stress
We investigate NMOS transistors in the 1V CMOS technology A, with Leff for
electrostatically sound devices (DIBL <= 120 mV/V) down to -45 nm (see Figure 2.112.12).
Using a four-point bending apparatus, schematically illustrated in Figure 5.1,
compressive and tensile uniaxial stress parallel to the direction of electron transport is
applied to a silicon strip containing several processed die. Surface strains of up to 0.12%
are achieved, as estimated from the sample thickness T and radius of curvature R
according to the relation: strain = T/2R. This method allows electrical characterization of
the same set of devices with and without strain, reducing sources of experimental error.
Fractional change in low-field effective mobility (6 eff
Age/
eff) corresponding to the
induced strain is measured in long (10 gm) and short (45 nm) devices, as shown in Figure
5.2. With Vds low, for linear-region operation, and for Vgs = Vdd, effective mobility is
determined from Eqn. (4.4):
IdLeff
geff = V*dsWQi
(5.1)
where V*ds (= Vds - Id Rsd) is the intrinsic drain bias. As in Chapter 4, inversion charge
density Qi is determined from measurements of capacitance of the long device, but
adjusted for the short device via Eqn. (4.6). This long-channel capacitance was measured
without strain. Since Vth shifts with strain E, it is necessary to add an additional term
AVt-
Vt(E=0) - Vt(E#0) to the integration limit in Eqn. (4.6). This approach assumes
that introducing strain does not change the shape of the CV characteristic, but only causes
a shift. Experiment indicates that this assumption introduces an error of less than 0.1%.
86
Some of the problems associated with mobility measurement in short devices--difficulty
in accurately determining Leff and Qi--are not a significant problem for this experiment,
because only the ratio of strained to unstrained mobility is required. By measuring geffstrained/geff-unstrained the uncertainties in Leff and
O
tt 1. t
0
mint tt
tt
Qi cancel
t
out.
t It
ett t
t
t
t
t
I-V, C-V
wi
a
V TC
~w
Figure 5.1: Schematic illustration of strain experiment.
This "cancellation of uncertainties" does not apply to the drain-bias term. For short
channels and low Vds, channel resistance (Rch) is of the same order as Rsd. Because of
this the use of V*ds instead of Vds is found to be critical in the determination of 8peff (short
device): the large relative contribution of Rsd to total series resistance means that AV*ds
with strain is significant (AV*ds = Vds - AMdRsd)-
Effect of uncertaintiesin Rsd on 8geff measurement
For our analysis we assume that Rsd does not change with strain; this is not strictly true,
because of the piezoresistance effect in the source/drain. Note that:
89ef
/ Id0 V*dsE
IdFV*dso
d
87
-
1
(5.2)
+10
Vgs = 1V
I
I
I
Vds = 50 mV
S+5-
10 pmr
U)Leff
4--
0
Leff
00
45 nm
00
-5
-0.05
Tech. A
I
0
I
+0.05
+0.1
% Uniaxial Strain
Figure 5.2: Experimental results: normalized mobility shift in long and short (Leff = 10
gm, -45 nm) devices vs. uniaxial strain. Each point represents a fractional shift relative
to unstrained value (which is represented by *).
where I', V*dsE are values with strain, and Id, V*dsO unstrained. From this it can be seen
that the piezoresistance effect on Rsd may lead to an overestimation of 89eff
For
example, with tensile strain the piezoresistance effect will increase Id' and decrease VdSE,
so that the 8ggff measured is larger than the "intrinsic" value. The effects of this will be
discussed further in Section 5.3.
The determination of
ggff in short devices is also sensitive to misestimation of the
nominal (unstrained) value of Rsd; this too will be discussed in Section 5.3.
Determination of
pg9ff in the long channel device is not significantly affected by this.
This sensitivity to Rsd is evident in the greater scatter among data points for 8geff-short:
for each measurement at a new strain value, the probe-tip-to-pad contact resistance
varies, changing slightly the total S/D series resistance. This is also why the straight-line
fit to the data does not pass through (0,0) for the short devices.
88
On long- versus short- channel behavior
The difference between
6 peff-long
and kteff-short--a 40% reduction of dependence on
strain for geff in the short devices--may be indicative of a transition in the dominant
scattering mechanism with device scaling, as discussed in Chapter 4. peff-ong is limited
by interface scattering (Section 4.1). However, for deep-sub-100-nm channels alternate
scattering mechanisms may dominate (Section 4.4). If this is in fact the case, a different
dependence of gerf on strain is possible.
Vgs = 1V
2
.
1
Vds=1V
10-
0
o -1 -2
,*
* *-
Tech. A
-0.05
0
0.05
0.1
0.15
% Uniaxial Strain
Figure 5.3: Experimental results: normalized velocity shift (near the source) in short (Leff
-45 nm) device vs. uniaxial strain. Each point represents a fractional shift relative to
unstrained value (which is represented by *).
Velocity versus Mobility
For this work, velocity is extracted via two methods described in Chapter 2: either ve =
vgmi from peak intrinsic saturated transconductance (Eqn. 2.5), or ve = viad from saturated
drain current (Eqn. 2.8). Figure 5.3 gives the fractional change in short-device electron
velocity (ve
- Ave/ve), plotted versus strain; here ve=vidi. As with short-device gf,
measurements are sensitive to probe-tip-to-pad contact resistance variation. In Figure 5.4
89
short-device
8
ve is correlated with long-device
6
igff.
(Note that ve, vgr
and vidi always
refer in this chapter to velocity in the short devices). The measured 8ve dependence on
strain is essentially the same by either method, appearing to be independent of velocity
extraction method. A least-squares linear fit gives 5ve/geff= 0.3.
S
C,,
a!)
CD
0
I
I
I
. +2
+4.
00
-C
I
Vds = 1V
Vgmni"
+20
Cl)
_
0
-2
-
Vidi
0
>
Tech. A
-5
0
+5
+10
% Mobility Shift, Long Device
Figure 5.4: Experimental results: 8 ve extracted by two methods, versus long-device 8 9eff.
Mobilities are measured at (Vgs,Vds) = (IV, 50 mV). Velocity vidi is measured at
(Vgs,Vds) = (IV, lV). Velocity vgm is from peak saturated intrinsic transconductance,
measured at Vds= lV; the peak occurs for Vgs somewhat less than IV. For this data 6ve /
8 peff = 0.31 when using vgn, and 0.30 for vidi. Each point represents a fractional shift
relative to unstrained value (which is represented by *)
It is more relevant to examine the relation between velocity shift and short-device
mobility shift (Figure 5.5) so that mobility and velocity are measured in the same device.
Again the results do not depend strongly on the velocity extraction method employed.
This smaller mobility shift for short devices, noted previously, translates to a 55% larger
value of 8ve/pg9ff than seen in Figure 5.4. Clearly, for understanding transport in deepsub-100-nm MOS devices, it is not valid to infer short-device mobility behavior from
90
measurements of long devices from the same technology. We denote the ratio 8 ve/sgeff
(with both values from the short devices) as RVp, and interpret it as a measure of the
dependence of source-end electron velocity in the MOSFET saturation regime on lowfield mobility. From Figure 5.5, Rv9 = 0.46-0.48.
CD
0
+4
+2
C.)
CD
V
+2
0
0
-0
-2
0
4-**
e
O*
-2
,
Vidi
-4
-4
-2
0
+2
+4
% Mobility Shift, Short Devices
Figure 5.5: Experimental results: 6 ve extracted by two methods, versus short-device
6 pgff. Gate and drain biases are as described for Figure 5.4. For this data, the ratio Rvg =
6 ve / peff = 0.48 when using vgmi, and 0.46 for vidi. Each point represents a fractional
shift relative to unstrained value (which is represented by *).
5.3 On Sources of Error in R jt Estimation
Series Resistance Effects
As mentioned, neglecting the piezoresistance effect in the source and drain may lead to an
overestimate of 89eff, producing an artificially low result for Rvg.
To assess the signifi-
cance of this effect, we measure a square resistance test structure for this technology (technology A): an n-well in a p-type substrate, with doping in the low-10
91
17
cm-3 range, to a
depth of 1.5 gm. The test configuration is illustrated in Figure 5.6.
We find that test-
structure resistance Rt varies linearly with applied strain E. This is characterized by a
gaugefactor:
G = (ARt/Rt) IE
For our test structure G = 50.
To estimate from this result how Rsd is changing with
strain, we make use of two observations: (1) from inverse modeling results for technology
A we find that polysilicon resistivity contributes approximately 40% of the total Rsd, and
(2) the piezoresistance effect decreases with increasing doping, by a factor of 2 for an
increase of Nd from low-10
17
to mid-10 1 9 cm- 3 [61].
This latter doping level is more
appropriate for the source and drain of the real MOSFET. We define an effective gauge
factor Geff
(ARsd/Rsd) /E which takes these two corrections into account: Geff = (G/ 2 )*
40% = 10. For our maximum tensile strain of 0.12% this means that Rsd may decrease by
1.2%. This means that we are underestimating V*ds (= Vds - Id Rsd) at strain by up to
1.2%, since Vds = 2 *IdRsd (from Figure 4.9) for these bias conditions. This is very significant, given that the maximum measured shift in peff-short in Section 5.2 was 5%.
To
recalculate Rv taking S/D piezoresistance into account, in Eqn. (5.1) to calculate peff we
let:
V*ds = Vds - Id Rsdo (1- GeffE)
where Rsdo is unstrained S/D series resistance. For velocity calculation, in Eqn. (2.8) we
replace the integration limit IdRs with IdRs(1-GeffE).
Re-solving, we find Rv is signifi-
cantly increased: from 0.47 to 0.59.
In addition we check the sensitivity of our measurement of Ryg to an misestimation of the
unstrained valued Rsd by 10%: with all parameters extracted assuming Rsd = 190Qpm *
1.1, we find that Ryg is 0.42-0.43: approximately 10% reduced. As we found previously
in Chapter 4, it is essential to have an accurate Rad estimate for determination of shortdevice transport quantities. This sensitivity of Ry, to Rsd is mainly through the 8peff
measurement: this is evident from the fact that, making the same assumption of Rsd =
190Qgm * 1.1 we find that 8ve shifts by only about 1% (an increase).
92
Axis of stress
V2
V3
44
114
200 gm
114
Figure 5.6: Resistance test structure setup. Current was forced through terminals 1-4,
Test strucand the corresponding voltage drop measured between terminals 2 and 3.
ture resistance Rt was measured for strain = 0, 0.04, and 0.08%. Rt
V3 -V 2 / 114-
Drain Bias for Low-Field Measurements
For all mobility measurements in this chapter, Vds=50 mV. To verify that this in fact corresponds to low-field conditions (i.e. far from velocity saturation) for a 45 nm MOSFET,
we estimate the corresponding intrinsic value V*ds from Figure 4.9 as approximately 25
mV. Making the simple approximation that (for Vds << Vgs-Vt) lateral field Ex~ V*dsLeff,
we find EX = 5.6 KV/cm. We evaluate this via the Caughey-Thomas expression [64]:
(EX) = F
i0
1+
(5.3)
(
vsat
7
If we estimate go = peff = 250 cm 2 /Vs (from Figure 4.8) and vsat = 1xi0 cm/s, then we
find g(Ey) / go = 0.99. This suggests a sufficiently low applied Vds was used. This is
further supported by Monte-Carlo simulations of the drift-velocity / electric field relation
3
17
in doped bulk silicon [65], where for bulk doping density Nb = 4x10 cm- velocity-
93
saturation effects do not begin to appear until Ex = 7 KV/cm. For Nb = 4x10 18 cm- 3 , this
critical EX value increases to
-
15 KV/cm. These simulations, however, were performed
for the <111> crystallographic direction.
5.4 Energy Balance Modeling
We find that drift-diffusion (DD) simulations severely under-predict RW.
However,
closer agreement is found in earlier theoretical work with energy transport models [32],
with a value of RVg = 0.5 predicted at the 50-nm channel length node, compared to our
experimental value (with S/D piezoresistance corrections) of RV = 0.6. We explore this
further with energy balance (EB) modeling of a realistic simulated super-halo NMOSFET
with 2D
doping profiles
carefully designed to match
measured subthreshold
characteristics (DIBL, subthreshold slope, Ioff) of the short (45 nm) devices. To model
drain current, a "universal" 9gef vs. Eeff relation is used [9][31], but modified to fit our
measured values of peff-ong (as described in Section 4.3).
The EB model requires
specification of an energy relaxation time (Tw) characterizing the decay of carriers with
kinetic energy values out of equilibrium with the semiconductor lattice: this cannot be
measured directly, and must be determined from theoretical calculations or from
calibrating simulations to experimental results. With our calibrated universal mobility,
and with r, = 0.11 ps and an electron saturation velocity vsat = 8.3x10 6 cm/s, we match
closely the unstrained Iof vs. Ion characteristic of NMOS devices over a broad range of
gate-length in the experimental technology A.
This value of vsat agrees with
experimental values found for silicon inversion layers in long-channel devices [62]. To
simulate the introduction of strain, we use a simple multiplier to increase Reff by 10% for
all Eeff. This assumes that both t, and vsat are independent of strain; for vsat this is
supported by theoretical studies [63].
The resulting shift in Ion (2.7%) corresponds
directly to carrier velocity shift: simulated Rv = 8Ion/keff = 0.27. However, it has been
pointed out that tw may not be independent of strain: experimental results for Si/SiGe
strained-silicon NMOSFETs have been successfully modeled by assuming Tw and geff
94
have approximately equivalent fractional change with strain [34].
Monte Carlo
calculations in [63] support this assumption of t,-increase with strain. Assuming 8'r =
8
geff (where 8,
= ATw/Tw), we find by simulation Rv
= 0.55, much closer to the
experimental value.
On Energy- and Momentum- Relaxation Interdependence
In [66] an expression is derived relating t, with Tm (momentum relaxation time) where p
= qTm/m* (m* = effective mass). Starting with expressions for conservation of electrons,
momentum, and energy as derived from the Boltzmann transport equation, if (1) steadystate homogeneous conditions exist, (2) the Caughey-Thomas expression relates chordal
electron p. to lateral field Ex, and (3) the electron diffusivity is only weakly dependent
upon electron temperature Te (as suggested by Monte Carlo and experimental techniques,
summarized in [67]), then the following relation holds:
T
3kBTO 110
2
q
1
V2 1+(To/Te)
(5.4)
sat
where To is the lattice temperature.
In this case, since IoocTm, tw = A tm where the
proportionality constant A depends only upon Te. This suggests a physical basis for
letting 8T, = 8pef with strain. This simple picture is complicated by the fact that Te will
increase with
TW:
as Te increases A increases. This suggests 8Tw could be greater than
peff, unless Te is significantly higher than To at the channel location where electron
velocity is extracted.
The above derivation makes no assumptions regarding the dominant scattering
mechanism. However, Monte-Carlo simulations for bulk silicon (for the <111> direction)
do not suggest
Tw
and Tm are so strongly interrelated, for the case of heavy Nb. For 104
EX < 105 V/cm, almost no dependence of Tw on Nb is indicated [65].
<
This is in apparent
contradiction with the above conclusion. The quantity 'r, may have different physical
interpretation depending upon the implementation of different models.
95
5.5 Conclusion
By corroborating measured velocity and mobility dependence on strain, we have
demonstrated experimentally for the first time the importance of low-field effective
inversion-layer mobility in deep-sub-100-nm bulk NMOS.
Crucial to this work is the
determination of mobility shift directly from the shortest device, for which accurate
estimates of Rsd are essential. Energy balance models are found to agree with the
observed behavior, when it is assumed energy relaxation time shifts with strain. The
experimental value Ry, ~ 0.6 suggest that the 14% velocity degradation seen in Chapter 2
(moving from technology B to the shorter technology A [Section 2.4]) should correspond
to a 24% mobility degradation. This agrees well with the results of Section 4.4, where a
mobility decrease of up to 25% was estimated moving from "B" to "A".
96
Chapter 6
FDSOI Design-Space
6.1 Introduction
In Chapter 3 it was shown that FDSOI (Fully-Depleted Silicon-On-Insulator) MOSFETs
with mid-gap workfunction gates should have a significant mobility advantage over bulk,
in the deep-sub-100-nm regime.
In this chapter we examine via 2D simulation the
design-space for three basic FDSOI alternatives: single-gate (SG), and double-gate (DG)
either with symmetrical workfunction mid-gap gates (DG-CDmg) or asymmetrical workfunction n+p+ gates (DG-n+p+).
We focus primarily on scalability and drive current.
Methodology
Investigations have shown that DG FDSOI MOSFETs may ultimately scale to -10 nm
channel length [68] [69].
In this chapter we focus on nearer-term possibilities, consider-
ing scaling constraints and performance for the same four design nodes (based approximately on the 1999 ITRS) discussed in Chapter 3, with Leff from 25 to 70 nm (Table 6.1).
As we will show, for this regime it is not required to scale silicon film thickness (Tsi) for
FDSOI below 5 nm for most cases. For Tsi > 5 nm, quantum-mechanical effects on
threshold voltage are not significant [39], suggesting that a classical distribution of inver-
97
sion-layer electrons is still an appropriate approximation. This allows use of a 2D coupled
Poisson's / drift-diffusion solver [13] for investigating the sub-threshold region of device
operation, to determine scalability and threshold-voltage control. For Tsi < 5 nm, Vt
depends critically upon film thickness due to quantum effects [39]; even slight variations
could make such ultra-thin films impractical. For modeling saturated drive-current (Ion)
we use an energy balance model; the physical justification for this is not clear for deepsub-100-nm devices. However, as noted in Section 5.4, when model parameters are set
with proper care the EB model gives realistic results, for Left down to at least 40 nm.
TABLE 6.1
FDSOI DESIGN NODE PARAMETERS
Approximate ITRS
Year
Lef
(nm)
T xphyst
Vdd
loff
(nm)
(V)
(A/gm)
2000-1
2002
2005
2008
70
50
35
25
2.4
1.8
1.4
1.1
1.35
1x10- 8
1.1
0.9
0.75
1x10- 8
1x10~8
1x10-8
tTOXelec is 0.1-0.2 nm greater than TOxPhys
To retain the mobility advantage discussed in Chapter 3, all films are assumed intrinsically
doped: threshold voltage Vt is determined by the gate workfunctions and device geometry.
High mobility is an additional reason for considering only films with Tsi > 5 nm: thinner
films may have degraded mobility due to the proximity of a second scattering interface
near the inversion layer [44].
Gate oxide thickness considerationsfor FDSOI
In Chapter 3 (Figure 3.7) it was shown that, for inversion-layer areal density Qi/q= 1x10 1 3
cm-2 an FDSOI device with mid-gap gate will operate with effective transverse field Eeff
= 0.75 MV/cm, while for a bulk MOSFET Eeff will be approximately 35% to almost 100%
higher, depending upon the scaling node. From these numbers, we can infer from 1-D
coupled Poisson-Schroedinger solutions in [36] that Tinv (inversion layer centroid depth)
in an FDSOI device will be -1.0 nm deeper compared to a heavily-doped bulk device at
98
the same scaling node. This translates to a difference in T0 xelec of -0.3 nm. Because of
this it may not seem appropriate to use the same roadmap for TOelec for FDSOI. However, if the lower limit to To, is direct tunneling leakage, FDSOI may permit thinner physical To, by virtue of its substantially reduced transverse surface electric field.
Unfortunately, direct tunneling has no simple dependence on VgS or Eeff [72], so it is difficult to evaluate this benefit here.
For our analysis in this chapter, we assume the same
TOXphys, per generation, as was used for bulk-device simulation in Chapter 3.
Solving Poisson's equation assuming a classical distribution for the electron inversion
layer results in a very modest increase in Tinv with reduced channel doping and Eeff:
TeX lec for FDSOI in our simulations is 0.1-0.2 nm larger than TOxphys (no gate-depletion
effects), and up to 0.1 nm larger than Toxelec for the simulated bulk devices of Chapter 3.
6.2 Single Gate Scaling Results
Single-gate FDSOI scalability has been examined in [39] with a similar approach (2D
Poisson/drift-diffusion solver, and a classical distribution of electrons): Vt-rolloff (with
Vds=Vdd) was examined versus Leff for various Tsi and Tox combinations. We take an
approach more suited for comparison to roadmap nodes: several parameters (Leff, T0 X,
Vdd) are scaled together. For each scaling node we vary Tsi to meet two criteria. As in
Section 3.2 we require DIBL ; 100 mV/V. Regarding the second criterion: in Section 3.2
scaling was examined for two structures with similar subthreshold-slope (S) characteristics; a minimum-worst-case-Vt requirement was enough to insure that a device could be
sufficiently turned off.
However, S can vary widely between SG and DG structures: a
maximum-I0 ff requirement (1x 108 A/pm) is more suitable here. In Figure 6.1 we see the
effect of varying Tsi on DIBL and Iof for the four scaling nodes. Each point represents a
simulated structure from which DIBL and Ioff were extracted. Solid lines represent the fit
to this data, from which the critical Tsi was interpolated. The results are shown in summa-
99
requirements below Leff = 35 nm will require Toxelec <
rized in Figure 6.2. Meeting I
1.3 nm and/or Tsi < Snm.
Gate:G =4.6 V
x
TL
y
Tbox = 100 nm
Lef
15 nm
(b) lof vs. Tsi
(a) DIBL vs. Tsi
2001
Leff (nm):
Leff (nm):
150-
25
35
25 35
50
70
,-7
E
50
70
100-
S-8
-J
M
- --
-
- --
- -
- -- - -
-9
501F
Vds=Vdd; Vgs=0
-1
II
0
10
20
0
10
20
Silicon Film Thickness (nm)
Figure 6.1: Scaling simulations for SG-mg device. (Avant! TMA Medici [13]). Drift3
20
diffusion with Lombardi mobility model. No gate depletion. S/D are n+ 1.5x10 cm ,
abrupt junctions. Na = 1x10 15 cm-3 in channel. In (a) and (b) each point corresponds to
a set of IV transfer characteristics from device with specified Tsi, from which DJIBL
(according to Eqn. 2.10) and Ioff were extracted. Tox and Vdd from Table 6.1.
On choice of D
For both DG- and SG-(mg structures, we choose a gate workfunction (4.6 V) -100 mV
below mid-gap, in order to give an appropriate long-channel Vt (300-350 mV, for SG-Gmg
and DG-qmg at all four scaling nodes; see [73] for a discussion of optimal Vt). This is
expected to result in a PMOS Vt that is too negative, which can be corrected by counter-
100
20
Criteria:
o DIBL = 100 mVN
C
15 C/
U,CD
l
off
= 1x10~8 A/pm
10-
E
C
0
5
SG-CDmg
0
30
40
50
60
70
Effective Channel Length (nm)
Figure 6.2: Interpolated results from Figure 6.1: Tsi required per scaling node of Table
6.1. To, and Vdd vary with Leff according to Table 6.1.
doping of the channel, if mobility in the PMOS device is not considered as critical as in
NMOS. Alternatively, there has been some progress in modifying the workfunction for a
mid-gap material by controlling deposition parameters [74]: one gate material could have
cDg optimized for both N- and PMOS. Our choice of 09=4.6 V corresponds to Schottky
barrier measurements of tantalum [71]. SG FDSOI MOSFETs have been fabricated successfully with Ta gates, but with Vt -100 mV lower than expected for both N- and PMOS;
the authors attribute this to surface state charge [70].
6.3 Double Gate Scaling Results
DG Alternatives
The two alternatives for DG FDSOI design are illustrated schematically in Figure 6.3.
Symmetrical DG-Dmg has two inversion layers and DG-n+p+ one. In the long-channel
limit Qj for the two cases can be simply related if it is assumed that the inversion charge is
a 2D sheet at the Si-SiO 2 interface. In this case the silicon film can be treated as a dielec-
101
tric layer, and the capacitor model shown is appropriate.
The corresponding ratio of
charge between DG-n+p+ and DG-CDmg is shown in Figure 6.4 for a range of Tsi and Tox.
For thin silicon films this charge-sheet assumption is not appropriate, in which case the
Qj
relationship must be found from a Poisson solution (or coupled Poisson-Schroedinger
solution, for very thin films). In addition, 2D effects at short channels significantly modify
Qj for
a DG-n+p+ device, so that no simple relationship exists between
Qj in the
two
alternative architectures. In this case Figure 6.4 is only illustrative of qualitative trends.
FSzED
77~
T
T
(a) DG-CDmg
(b) DG-n+p+
Figure 6.3: Alternative double-gate FDSOI structures.
E
0
CL
1
0.9
Increasing Tox
0.8
0
........
0.7
0.6
Al.
50
-O Te= 1.0 - 2.5 nm
5
10
15
20
Silicon Film Thickness (nm)
Figure 6.4: Ratio in inversion charge in DG-n+p+ vs. DG-(Dmg MOS, according to longchannel charge-sheet model.
102
Leff
15 nm
(a) DIBL vs. Tsi
(b) lof vs. Tsi
-6
200
Leff (nm):
Leff (nm):
25
150
5
50
100 - - - - - -
- -
70
- - - - - - -
25 35
50
70
E -7- - - -
-
-8 -
--
50-
-
-
70
- - - - - - - - - - - - -
9.
Vds=Vdd; Vgs=O
n
0
-
20
40
1n1
60
0
10
20
30
Silicon Film Thickness (nm)
Figure 6.5: Scaling simulations for DG-n+p+ device. Model and doping specifications
from Figure 6.1 apply. To, and Vdd from Table 6.1.
Simulation Results
As with the single-gate device, we vary Tsi and interpolate between DIBL and Iof results
from many simulated structures (Figures 6.5 - 6.6) to find the critical Tsi to meet the two
short-channel criteria.
Results are summarized in Figure 6.7.
As we find consistently
for FDSOI, the IOf criterion is the more difficult to meet. In [8] it is pointed out that the
DG-n+p+ should have better electrostatic confinement and hence relaxed Tsi requirements. From our simulations, this is true if we only consider the DIBL criterion. Considering as well maintaining Ioff : 1 x10~8 A/gm, DG-n+p+ actually requires thinner Tsi. In
this sense DG-CDmg scalability is found to be superior.
103
Le)G = 4.6 V
',Ieff
T"; - -
D
-4-(b) loft vs. Tsi
(a) DIBL vs. Tsi
200
Leff (nm):
Leff (nm):
150
E
35
50
70
-
-
- - 7- --
E
1001-
0
35
7
--
25
-
-8
m
_j
--
50
25
-9
'
'
'
)
'
40
20
-10-
20
10
0
60
30
Silicon Film Thickness (nm)
Figure 6.6: Scaling simulations for DG-(Dmg device. Model and doping specifications
from Figure 6.1 apply. To, and Vdd from Table 6.1.
I
50
a
I
I
50
60
Criteria:
C
40
A O DIBL = 100 mVN
= 1x10-8 A/pm
AO If
0
C:
30
DG- +p+
C.
E
20
C
10
0
0)
'0
30
40
70
Effective Channel Length (nm)
Figure 6.7: Interpolated results from Figure 6.5-6.6: Tsi required to meet short-channel
criteria for each scaling node of Table 6.1. To, and Vdd vary with Leff according to Table
6.1.
104
For these device architectures, DIBL and Vt-rolloff behaviors are decoupled: large DIBL
does not fully correlate with large Vt-rolloff, and vice versa. This can be seen in Figure
6.8, where the DIBL and rolloff components are separated for both architectures. (In this
context, Vt-rolloff is the lowering of low-Vds threshold voltage by 2D effects; high Vds
effects are captured in the DIBL metric).
(a) DG-n+p+
E
80
0 Vt-rolloff
C3 DIBL
60-
0
El
60-
Vt-rolloff
DIBL
40-
40-
E
20-
0
(b) DG-(mg
80
30
40
50
60
70
0
20-
0
30
40
50
60
70
Effective Channel Length (nm)
Figure 6.8: Simulation results: DIBL and Vt-rolloff for DG structures.
Vt-rolloff =
5
Vt(long-device) - Vt(short device), where Vt is linear extracted at Vds= 0 mV. Tsi values used
are from Figure 6.7, meeting the Ioff requirement. Other parameters as per Table 6.1.
DIBL is via Eqn. (2.10).
On threshold voltage dependencies in DG-n+p+ versus DG-(mg
We can understand the much-greater Vt-rolloff behavior of the DG-n+p+ device by investigating further the long-channel Vt dependencies. From both simulation and analytical
results presented in Figure 6.9 it is clear that the bottom-gate proximity has a strong influence on Vt. Threshold voltage is decreased dramatically as distance to the p+ bottom-gate
is increased, either by thickening the bottom-gate oxide (Tbox) or increasing Tsi. Considering the situation in 1-D, the proximity of the p+ gate tends to lower potential cD(y) in the
silicon film, and consequently (Ds at the top Si-SiO 2 interface as well. This in turn requires
105
a higher Vgate to raise (IS to the point of inverting the surface: raising Vt. This bottom-gate
influence is more pronounced as Tj and T,, are thinned.
For the case of short-channel
devices, it is conceptually useful to consider the channel (in sub-threshold, with no inversion layer) as a 2D-box where 1(x,y) is set only by the four boundary conditions: source,
drain, gate and bottom-gate. As the ratio Tsi / Leff becomes larger, the boundary conditions at the ends of the box (the source and drain) increase in influence over c1(x,y) and
hence Fs,
This means that gate control over cIS is decreased relatively; this affects the
bottom-gate much more, because of its greater distance to the top interface.
Vt-rolloff in
a DG-n+p+ device can thus be thought of as the bottom-gate progressively, with decreasing L, losing influence on Is and hence Vt
TOXphys= 1.5 nm
0
0.3-
Tbox =
S
1.5 nm
0
>
_
l
3
>, .ft
Simulated
Analytical
>0
3nm
3
Z
(p
= 0.2-
0)
6 nm
0.1.
0
5
10
15
20
25
Silicon Film Thickness (nm)
Figure 6.9: DG-n+p+ long-channel Vt: dependency on Tsi and Tbox, with corresponding
range of DG-(mg long-channel Vt indicated, upper right. Analytical are ID Laplace
solutions (appropriate in the limit of zero channel depletion charge) for Qi/q=1x1O 11
cm-3
Conversely, our simulations show that Vt in a long-channel DG-(mg device is nearly independent of Tsi and Tbox, as also indicated in Figure 6.9. The bottom-gate does not signifi-
106
cantly influence Vt, so the decrease of bottom-gate influence with scaling does not result
in a significant lowering of threshold voltage.
Importance of lateral SID profiles
The simulation structures for the scaling studies in Sections 6.2 - 6.3 have abrupt doping
transitions from source/drain to channel. In order to understand the importance of the lateral S/D gradient, the simulations in Figure 6.6 (b) are repeated with Gaussian profiles.
The Gaussian characteristic length a,, corresponds loosely to the lateral rolloff in nm/
decade-of-doping. As cy is varied, effective and physical gate lengths are kept fixed
(with Leff determined by the points at which the S/D doping fall to 2x10
the metallurgical junction length decreases with increasing a,.
19
cm-3 ). Only
We focus on the Ieff crite-
rion only, since for all architectures investigated in this chapter it is the more stringent.
The results are summarized in Figure 6.10. We see that for a DG-CDmg device their is no
significant deleterious effect on scalability for ax 5 4 nm.
This is similar to results
reported in [2] for a simulated bulk super-halo NMOSFET designed for Leff =25 nm.
25
E
20CD,
5
15.
E
10
C:
Q Abrupt
A ax=2nm
.
*
o0*
9o
30
40
50
ax=4nm
Tax=8nm
60
70
Effective Channel Length (nm)
Figure 6.10: Simulation results: effect of lateral S/D doping gradient on Tsi required to
meet Ioff 1 x 10-8 A/pim. Shown versus scaling nodes of Table 6.1. Simulation grid
spacing is 0.6 nm at transitions from S/D to channel. To, and Vdd vary with Leff according to Table 6.1.
107
6.4 Relaxing Ioff Criteria
In this section we consider the consequences of relaxing off-current constraints for future
scaling generations, for both SG- and DG-MOSFET architectures.
Innovative techniques
to manage static power dissipation may permit off>> 1x10-8 A/gm for the fastest devices
of a technology generation.
This is considered for SG-Dmg devices in Figure 6.11; mod-
erately larger values of DIBL (120 mV/V) are allowed as well. We see that this substantially relaxes constraints on Tsi. However, for 1x10- 6 < Ioff < 1x10
channel length) DIBL becomes the limiting criterion.
7
A/gm (depending on
Significantly, if Ioff = 1x10~7 A/gm
is permissible, SG-(mg devices can scale to 25 nm for Tsi > 5nm: this is near the prospective bulk scaling limit [2]. Figures 6.12 and 6.13 repeat this analysis for DG-n+p+ and
DG-Dmg devices.
20
Criteria:
0 DIBL = 120 mVN
Wc15 - 0 'off, A4"m
~1X1
a)
0-7
0'
-c
10-8
'
10-
E
50
0
io
30
40
50
60
70
Effective Channel Length (nm)
Figure 6.11: SG-@mg MOSFET, interpolated results from Figure 6.1: Tsi required per
scaling node of Table 6.1. T0 , and Vdd vary with Leff according to Table 6.1.
108
6( I
50,
*
C,)
C
Criteria:
o DIBL = 120 mVN
Ioff, A/m
40-
-o-
30-
E
0
1x10-7
o'
2010F
30
40
50
60
70
Effective Channel Length (nm)
Figure 6.12: DG-n+p+ MOSFET, interpolated results from Figure 6.5: Tsi required per
scaling node of Table 6.1. To, and Vdd vary with Leff according to Table 6.1.
6( 'I
C 50.
Criteria:
o DIBL = 120 mVN
*
CLL
CD
lof, A/Lm
E 4030-
E
C
0
1 X10-7
20-
'.
-x0-8X1
-~~
......
10-
0)
2)
30
40
50
60
70
Effective Channel Length (nm)
Figure 6.13: DG-CDmg MOSFET, interpolated results from Figure 6.6: Tsi required per
scaling node of Table 6.1. To, and Vdd vary with Leff according to Table 6.1.
109
6.5 Energy Balance Simulation of Io
As noted in Section 5.4, EB simulations have been found to match well the experimental
Ion/loff characteristics for a family of bulk devices with Leff down to -45 nm. For this reason we choose here to focus on the Leff= 50 nm scaling node of Table 6.1, for our use of
EB simulations to compare saturated drive current for the three FDSOI architectures studied in this chapter. Using the Tsi values found from Figures 6.2 and 6.7 to meet Ioff =
1xi 0-8 A/gm, we find Ion using the Lombardi mobility model [75]. In [76] it was shown
that the Lombardi model closely approximates universal mobility behavior for the case of
low channel doping. Energy relaxation time tw was 0.15 ps. This is reasonable from the
perspective of Eqn. (5.4), assuming vsat=9x10 6 cm/s and go of 300-350 cm 2/Vs as determined from Figure 3.9 for SG with Qi/q = 1x10 13 cm-3 or DG with Qi/q = 2x10
(reasonable values assuming Vdd and Tox chosen from Table 6.1).
13
cm-3
The results are shown
in Figure 6.14. The EB model gives moderately larger values for Ioff, depending on the
-7.
R.
Leff = 50 nm
Vds=Vgs=Vdd
E
-6
0
0
0
E
0 E
0
-8
CM
loff from DD model
0
0 SG-Dmg
E
-8
Tsi
DG-Dmg :Tsi
7.5 nm
18.5 nm
A DG-n+p+: T,,= 11.5 nm
'0
0.5
1
1.5
2
2.5
Ion (MA/9)
Figure 6.14: Energy balance simulation results for Ion and Ioff for the three FDSOI architectures. For solid symbols, lumped resistances were added to the source and drain (80
Qgm per side). Total Rsd for solid symbols is 190-200 Qgm; for open symbols, 30-40
Qpm. Tsi values chosen to meet Ioff criterion(Figure 6.2-6.7); other parameters are for the
50 nm node in Table 6.1. r,=0.15 ps, Lombardi mobility model.
110
device architecture.
We see that as expected, for the case of low source/drain resistance
(Rsd) the Ion for DG-(cmg is approximately twice that of SG-(cmg. DG-n+p+ falls in
between.
The total gate capacitance is significantly different for each architecture; from
the simple model of Figure 6.3 we can see that Csg < Cdg-n+p+ < Cdg-cDmg- We can normalize for this difference, by comparing carrier velocity as determined from gm / WC'oX
(Eqn. 2.4) where gm is the peak saturated transconductance. Cox is determined in simulation from the maximum slope of the dQi/dVgs curve (taken for a short device with Vds=O,
extracted from mid-channel; the validity of this approach for a short device was demonstrated in Section 4.2). The results are given in Table 6.2; the lower velocity of DG-n+p+
can be attributed to lower mobility with higher Eeff, in turn due to the fact it has inversion
charge density
Qj
higher than the SG device, and higher than the
Qj per channel for
the
DG-dmg device. Note that the degradation of Ion and gm / WC'x due to Rsd appears to be
greatest for the DG-Dmg device: with higher current the potential drop across Rsd is larger,
so the intrinsic Vds is degraded further.
TABLE 6.2
NORMALIZED TRANSCONDUCTANCE FOR THREE FDSOI ARCHITECTURES; LEFF = 50 NM
gm/WC'ox (cm/s)
ipm
Rsd=190-200
gm/WC'ox (cm/s)
Rsd= 3Q- 4 0 Qgm
SG-Dmg
8.6x10 6
9.9x10 6
DG-n+p+
6.7x10 6
8.3x10 6
DG-Dmg
7.9x10 6
10.1x10 6
111
6.6 Conclusion
We have explored the design-space for the three basic FDSOI NMOS architectures, which
were shown in Chapter 3 to have the potential for significant mobility advantage to bulk at
deeply scaled channel lengths. For the single-gate device, it has been shown that scaling
below Leff = 35 nm may not be achievable for practical silicon film thickness, unless Ioff
requirements are relaxed. For double-gate devices we have shown that hypothetical midgap top- and bottom- gates are superior to n+/p+ poly gates, in terms of both scalability
and drive current.
112
Chapter 7
Double-Gate Fabrication Technology
7.1 Introduction
Realization of the ideal double-gate device structure involves three major technical challenges: formation of gates above and below a thin single-crystalline silicon layer, achievement of fine alignment between top- and bottom-gates, and achieving low source/drain
resistance for the thin silicon film. We address these issues through integration of three
primary technologies: wafer bonding with pre-patterned features, interferometric alignment, and selective epitaxy for raised source/drains.
7.2 Formation of Gates
Three basic DG fabrication approaches are illustrated schematically in Figure 7.1.
Approaches (b) and (c) place extremely stringent demands upon minimum lithographically-defined feature size (Lmin), and hence may not be practical for fully-depleted double-gate devices [8].
From Figure 6.7 we can see that for the 50 nm channel-length
generation, Tsi < 20 nm (given Toxelec = 2 nm). For (b) and (c) Ti is defined by lithogra-
phy and etch: Tsi=Lmin.
This is approximately four times smaller than the Lmi. required
113
for a bulk MOSFET at the same Leff generation (Lgate = 70-80 nm, assuming 10-15 nm
gate-to-S/D overlap per side). Although there are various approaches for laterally shrinking features beyond the limits of the photolithography technology (e.g. resist-ashing [78],
or trench-narrowing via spacer formation [79]), barring a fundamental breakthrough these
do not solve the problem because they do not reduce the variation AL .n- Typically the
required tolerance has been +/- 10% [80]; this is also the ITRS requirement through 2014
[1]. Lithography sufficient for bulk at Leff = 50 nm may have Lgate ~ 80 nm and thus
ALmn on the order of +/- 8 nm. For (b) and (c) this translates to equivalent ATsi: from Figure 6.6b, it is apparent this at this would result in totally unacceptable 10: in the worst8
case about 20X the nominal value of 1x10- A/gm.
z
rrent-Carry ng
tiZ
Plane
x
Y
z
urrent-
Y
X
ying
oBottom
Bott
Gater
X
Icon Wafer
(a)
Illcon
(b)
G
RMcon Wafer
Water
(c)
Figure 7.1: Three fundamental structural possibilities for DG-FDSOI MOSFETs; illustration from [77].
Option (a), the planar geometry, appears more feasible: technologies for defining vertical
dimensions (implantation, diffusion, epitaxy) can be very finely controlled. Planar double-gate devices have been fabricated by pre-patterned bonding [81][82]. We have developed a variation based on the SOIAS process (Silicon-On-Insulator-with-ActiveSubstrate) [83]. This process is illustrated in Figure 7.2; further details of the SOIAS process can be found in [84]. The CMP planarization technology required for pre-patterned
bonding is described in [82].
114
0
-t
0
Grow bottom-gate oxide on SIMOX
Pattern bottom-gates
.
SIMOX Buried Oxide
Etch SIMOX wafer and buried oxide
Polysilicon
Bottom-gate
Oxide
Silicon
0
Cn,
0
l.
LTO
Oxide
Bonded
Interface
(a)
(c)
Deposit LTO and planarize
Flip and bond to handle wafer
Front-End SOI Process
0
0
LTO
CD
LTO
Oxide
Handle
Wafer
0
Bonded
Interface
0
(d)
Cr
CD
(b)
7.3 Alignment of Gates
Several schemes for self-alignment of gates have been proposed [85] and demonstrated
[77][86], but have not been shown to be practical for deeply scaled devices with thin Tsi.
Models predict that the tolerance in aligning top and bottom gates has to be within Lgate/4
in order to avoid performance deterioration due to bottom-gate overlap capacitance [87].
In order to meet this challenge we apply the IBBI (Interferometric Broad Band Imaging)
alignment technique* [88]. A pattern of interference fringes is formed by diffraction from
a grating on the mask and a complementary grating on the substrate: the technique relies
on the extreme sensitivity of this pattern to the relative position of mask to substrate, to
achieve detectivity of sub-nanometer misalignment. We integrate this with proximity Xray lithography: the configuration of X-ray mask and pattern is shown schematically in
Figure 7.3. Fabricated test devices, involving pre-patterned bonding for bottom-gates and
IBBI alignment of top- to bottom-gates, is shown in Figure 7.4. We have obtained alignment detectivity on the order of several nanometers. However, final alignment results are
equally a function of precise pattern placement in the fabrication of top and bottom X-ray
masks. We are now developing a process to obtain this precision through close proximity
X-ray mask replication. Patterns from the top-gate mask are transferred directly to the
bottom-gate mask, eliminating errors in gate-to-gate relative placement due to e-beam
field-stitching and other sources.
7.4 Reduction of S/D Parasitic Resistance
Silicidation of thin-film SOI devices for low parasitic series resistance Rsd is difficult for
Tsi << 50 nm. It is crucial that the silicide does not completely consume the silicon film,
in order to maintain a large silicide/silicon interface area [89][55]. This is necessary for
traditional silicides with relatively high silicide-to-silicon barrier height Db. One possible
solution is to use alternative silicides with Db << Eg/ 2 (where Eg is the Si bandgap). In
[90] this was done for Schottky S/D PMOS with PtSi and NMOS with ErSi; consumption
* This work was performed in collaboration with Mitch Meinhold and Euclid Moon.
116
X-ray Flux
IBBI Signal
X-ray Mask
Resist / Hardmask
---
Poly-Si (Top gate material)
Channel
Buried gate &
Alignment marks
Van _Si
Si0
2
Si Wafer
Figure 7.3: Illustration of the IBBI alignment scheme, applied to double-gate fabrication
via X-ray lithography.
DG-test
devices
Figure 7.4: Test double-gate structures fabricated via pre-patterned flip-and-bond, with
IBBI alignment.
117
of the entire Si film did not lead to unacceptable Rsd. However, since CDbn = Eg - CDbp two
silicides are required for low barriers for CMOS. An alternative approach is to increase
the thickness of the S/D regions via selective epitaxial growth of silicon. This has been
previously demonstrated in single-gate FDSOI devices fabricated with a traditional silicon
nitride (SiNx) spacer process [91]. Epitaxial selectivity to SiNx is more difficult to
achieve than to Si0
2
[92]; in [91] this is achieved by introducing HCl during growth,
which etches polysilicon nuclei as they form on the SiNx.
LTO
LTO
Si0
SiO 2
2
(a)
(b)
LTO
Lateral
overgrowth
Si-epitaxy
2
SiO 2
(c)
(d)
Si0
Figure 7.5: Sacrificial SiNx spacer / selective epitaxial raised S/D process. (a) Gate stack
is etched using LTO (low-temperature oxide) hard-mask, followed by re-ox (not shown)
and S/D arsenic implants. (b) LTO liner and SiNx LPCVD deposition (13 and 90 nm
respectively), followed by anisotropic spacer RIE etch in NF 3 :0 2 , stopping on LTO liner.
(c) Pre-epitaxial wet clean (H2 0 2 :H 2 SO 4 followed by HF). (d) Selective removal of
SiN, spacers in H3PO 4 , followed immediately by UHVCVD selective growth.
118
Raised Source and Drain Process
We have developed a selective-epitaxial raised-S/D process (with sacrificial SiNx spacer)
which does not require SiNx selectivity*.
Figure 7.5 outlines this process.
Epitaxial
growth is in a vertical-flow hot-wall UHVCVD reactor with SiH 2 Cl 2/H 2/PH 3 , at 750C.
Growth is preceded by an H2 desorb step at 850C.
Figure 7.6 shows the results of this
process: high-quality S/D regions are formed, to be followed by silicidation.
Figure 7.6: Selective epitaxial raised S/D results, shown for a bulk MOSFET. Epitaxial
layer is 75 nm thick. The 13 nm LTO liner between epi and n+ poly is not visible here,
but its presence can be inferred from the lack of epi-growth on the n+ poly side-wall.
(SEM image taken at 100000X, uncoated).
Facetingat epitaxy/insulatorinterface
A significant problem encountered in the development of this process was the formation
of facets at the epitaxy-SiO 2 boundary. This phenomenon is well-known in selective epitaxy [92]. For the process of Figure 7.5 faceting was found to prevent lateral overgrowth
* This work was performed in collaboration with Tom Langdo.
119
(a)
Si Epi
Gate S'02
Si Substrate
50 nmn
(b)
Figure 7.7: XTEM images. (a) Faceting at Si / LTO interface, with 750C UHVCVD
growth via SiH 2 C2/H2. (b) Faceting suppressed, with 750C growth via SiH 2 C12 /H2/
PH 3 -
120
of the LTO liner, leaving the Si film thin at the liner edge: this could be consumed entirely
by silicidation, leading to high series resistance. We found faceting could be entirely suppressed by the introduction of moderate amounts of phosphorus in-situ doping (Na ~
1x10
18
cm-3 ) during the growth process, via PH 3 . Figure 7.7 demonstrates this with an
LTO spacers on bulk substrates.
The n-doping introduced into the source and drain by this process is low compared to
CMOS S/D doping levels: it is not an impediment in forming S/D regions for PMOS
devices.
7.5 Conclusion
We have demonstrated the key process modules required to build a deeply-scaled doublegate MOSFET: bottom-gate formation via pre-patterned planarization and bonding, direct
alignment of top and bottom gates via IBBI, and epitaxial raised source and drain for
reducing parasitic resistance associated with very thin-film FDSOI devices. Technical
issues still remain in the integration of these models, but there appear to be no fundamental roadblocks to this approach.
121
Chapter 8
Conclusion
8.1 Summary of Results
Experimental Results
Our study of transport in deep-sub-100-nm NMOS technologies yielded several new
observations, which together may be key to understanding drive-current performance
trends as bulk MOSFET scaling is continued.
In Chapter 2 it was shown that NMOS
devices from an advanced CMOS technology operate well below the thermal or ballistic
limit of source-to-channel carrier injection, and that continued scaling may actually be
decreasing both the ballistic efficiency and the carrier velocity. Degradation of electron
low-field mobility with channel length scaling, observed in Chapter 4, is a possible explanation for this. For this explanation to be tenable, velocity in deep-sub-100-nm MOSFETs
in saturation must have a strong dependence on low-field mobility: this was demonstrated
via the strain measurements of Chapter 5.
Simulation Results
Whereas the experimental work of Chapter 4 linked the mobility degradation in short Sibulk MOSFETs to a transition in the dominant scattering mechanism, in Chapter 3 via 2D
122
simulations it was demonstrated that even without this transition, future Si-bulk MOSFET
scaling is expected to lead to significant mobility degradation.
Single- or double-gate
Fully-Depleted (FD)SOI MOSFETs, if manufacturable with mid-gap workfunction gates,
should not suffer from this drawback. In Chapter 6 it was further shown that, from a scalability standpoint, mid-gap workfunction gates could be a significant enabler for FDSOI
MOSFETs, allowing single-gate devices to scale down to 25-35 nm channel lengths with
appropriate threshold voltage, and double-gate devices to scale with relaxed silicon film
thickness requirements. Changing MOSFET architecture from Si-bulk to FDSOI is one
viable route to improving mobility and velocity while maintaining scalability.
On the need for enhanced carriervelocity
The NMOS carrier velocities measured in Chapter 2 are sufficient to meet the roadmap
targets for future generations (Figure 1.1) only if aggressive Toxelec scaling is possible and
Vt requirements are relaxed. Whether or not these targets are met, velocity enhancement
leads to better performance or equivalent performance at lower power, and may open
applications to Si-based MOSFETs that are currently dominated by more expensive III-V
devices.
8.2 Future Work
Experimental Analysis of Deep-Sub-100-nm Devices
As discussed throughout this work, interpretation of experimental measurements of deeply
scaled devices are subject to uncertainties in series resistance Rsd and channel length Left.
This is especially critical for the results of Chapter 4, the relation between mobility and
Left in super-halo bulk devices: errors in Rsd estimation affect not just the quantitative
results but the physical interpretation. We have relied heavily on inverse modeling for our
Rsd and Left values; these numbers depend on the mobility model chosen, and are at the
present time not definitive. Therefore it is preferable to derive an alternative mobility
123
extraction method for deeply scaled MOSFETs that does not require precise knowledge of
Rsd and Leff. In [93] it is shown that the slope of the Rtot (total MOSFET series resistance)
versus Lmask relationship, within one technology, corresponds to low-field mobility effective mobility. This mobility is not constant with channel length, but it should nonetheless
be possible to extract a mobility-vs.-length relationship from the local slope of a fit to a
sufficient number of Rtot vs. Lmask data points. This approach is illustrated for a small
data set in Figure 8.1.
(a)
(b)
300
800 - Vds =50 mV
000
0
0
0
0
C
600 -
a
400 -
-
s0op
200-
E
0
250 1200
--
$150-r
: 100 .
e ffW Q i
50--
0501
100
150
200
250
100
150
200
250
Lmask (nm)
Figure 8.1: Illustration of alternative method for egff determination in short NMOS
devices, adapted from [93]. For given Lmask, mobility is found from dRtot / dLmaskInversion charge density Qj is determined from long-device CV integration, corrected for
short device according to Eqn. (4.6). Results are extremely sensitive to the polynomial
fit (here, to degree 3) to the data, suggests that more points are required for reliable
results. Technology A.
In addition, the exploration in NMOS devices of electron mobility, velocity, and their
interrelation should be repeated for holes in deep-sub-100-nm PMOS devices, for a complete understanding of complimentary device behavior.
124
Fabrication
Chapter 7 outlined the development of key process modules for the fabrication of thinfilm aligned double-gate FDSOI devices. Integration of these modules is a significant
challenge, but to date there has been demonstrated no alternative scheme for double-gate
MOSFETs that does not have serious drawbacks. The IBBI-alignment approach, with prepatterned bonding for bottom-gate fabrication and selective-epitaxial raised source and
drain, may be the most attractive alternative and should be pursued.
125
References
[1]
Semiconductor Industry Association "International Technology Roadmap for Semiconductors," 1999 Edition.
[2]
Y. Taur, C. Wann, D. Frank, "25 nm CMOS Design Considerations," IEEE IEDM
Technical Digest, p789 (1998).
[3]
R. Chau, J. Kavalieros, B. Roberds, R. Schenker, D. Lionberger, D. Barlage, B.
Doyle, R. Arghavani, A. Murthy, G. Dewey, "30nm Physical Gate Length CMOS
Transistors with 1.0 ps n-MOS and 1.7 ps p-MOS Gate Delays," IEEE IEDM Tecnical. Digest, p45 (2000).
[4]
H. Wakabayashi, M. Ueki, M. Narihiro, T. Fukai, N. Ikezawa, T. Matsuda, K.
Yoshida, K. Takeuchi, Y. Ochiai, T. Mogami, T. Kunio, "45-nm Gate Length CMOS
Technology and Beyond using Steep Halo," IEEE IEDM Tecnical Digest, p49
(2000).
[5]
M. Armstrong, "Technology for SiGe Heterostructure-Based CMOS Devices",
Ph.D. Dissertation, MIT (1999).
[6]
S. Lo, D. Buchanan, Y. Taur, W. Wang, "Quantum-Mechanical Modeling of Electron
Tunneling Current from the Inversion Layer of Ultra-Thin-Oxide nMOFETSs,"
IEEE Electron Device Letters 18, p2 0 9 (1997).
[7]
M. Koh, W. Mizubayashi, K. Iwamoto, H. Murakami, T. Ono, M. Tsuno, T. Mihara,
K. Shibahara, S. Miyazaki, M. Hirose, "Limit of Gate Oxide Thickness Scaling in
MOSFETs Due to Apparent Threshold Voltage Fluctuation Induced by Tunnel Leakage Current," IEEE Transactions on Electron Devices 48, p2 5 9 (2001).
[8]
H. Wong, D. Frank, P. Solomon, C. Wann, J Welser, "Nanoscale CMOS," Proceedings of the IEEE 87, p537 (1999).
[9]
A. Sabnis and J. Clemens, "Characterization of the Electron Mobility in the Inverted
<100> Si Surface," IEEE IEDM Technical Digest, p18 (1979).
[10] K. Banoo, J. Rhew, M. Lundstrom, C. Shu, J. Jerome, "Simulating Quasi Ballistic
Transport in Si Nanotransistors," presented at the International Workshop on Computational Electronics, Glasgow, May 2000. To be published in VLSI Design, 2001.
126
[11] G. Timp, J. Bude, K. Bourdelle, J. Garno, A. Ghetti, H. Gossmann, M. Green, G. Forsyth, Y Kim, R. Kleiman, F. Klemens, A. Kornblit, C. Lochstampfor, W. Mansfield,
S. Moccio, T. Sorsch, D.M. Tennant, W. Timp, R. Tung, "The Ballistic Nano-Transistor," IEEE IEDM Technical Digest, p5 5 (1999).
[12] F. Assad, Z. Ren, D. Vasileska, S. Datta, M. Lundstrom, "On the Performance Limits
for Si MOSFETs: A Theoretical Study," IEEE Transactions on Electron Devices 47,
p2 3 2 (2000).
[13] Avant! Medici, User's Manual, Version 1999.2 (1999).
[14] M. Lundstrom, "Scattering Theory of the Short Channel MOSFET," IEEE IEDM
Technical Digest, p3 8 7 (1996).
[15] G. Shahidi, D. Antoniadis, H. Smith. "Electron Velocity Overshoot at Room and Liquid Nitrogen Temperatures in Silicon Inversion Layers," IEEE Electron Device Letters 9, p9 4 (1988).
[16] S. Chou and D. Antoniadis, "Relationship Between Measured and Intrinsic
Transconductance of FETs," IEEE Transactions on Electron Devices ED-34, p448
(1987).
[17] S. Laux, M. Fischetti, D. Frank, "Monte Carlo Analysis of Semiconductor Devices:
The DAMOCLES Program," IBM Journal of Research and Development 34 (4),
p4 6 6 (1990).
[18] F. Assad, Z. Ren; S. Datta, M. Lundstrom, P. Bendix. "Performance Limits of Silicon MOSFET's," IEEE IEDM Technical Digest, p5 4 7 (1999).
[19] Y. Taur, Y. Mii, R. Logan, H.S. Wong. "On Effective Channel Length in 0.1 pm
MOSFETs," IEEE Electron Device Letters 16, p13 6 (1995).
[20] Y. Taur and E. Nowak, "CMOS Devices below 0.1 gm: How High Will Performance
Go?" IEEE IEDM Technical Digest, p215 (1997).
[21] D. Antoniadis. "Physics and Technology for MOSFETs at 0.1 micron and Below,"
ESSDERC Proceedings, p3 (1993).
[22] T. Mizuno, R. Ohba, "Experimental Study of Nonstationary Electron Transport in
Sub-0.1 gm Metal-Oxide-Silicon Devices: Velocity Overshoot and it's Degradation
Mechanism," Journal of Applied Physics 82 (10), 5235 (1997).
[23] Y Taur, D. Zicherman, D. Lombardi, P. Restle, C. Hsu, H. Hanafi, M. Wordeman,
Matthew R. B. Davari, G. Shahidi, "A New 'Shift and Ratio' Method for MOSFET
Channel-Length Extraction," IEEE Electron Device Letters 13, p2 6 7 (1992).
127
[24] L. Su, S. Subbanna, E. Crabbe, P Agnello, E. Nowak, R. Schulz, S. Rauch, H. Ng, T.
Newmann, A. Ray, M. Hargrove, A. Acovic, J. Snare, S. Crowder, B. Chen, J. Sun,
B. Davari, "A High-Performance 0.08 pm CMOS," VLSI Technology Symposium,
Digest of Technical Papers, p12 (1996).
[25] Z. Lee, "A New Inverse-Modeling-Based Technique for Sub-100-nm MOSFET
Characterization," Ph. D. Dissertation, MIT (1998).
[26] H. Hu, J Jacobs, L. Su, D. Antoniadis, "A Study of Deep-Submicron MOSFET Scaling Based on Experiment and Simulation," IEEE Transactions on Electron Devices
42, p6 6 9 (1995).
[27] N. Yang, W. Kirklen Henson, J. Hauser, J. Wortman, "Estimation of the Effects of
Remote Charge Scattering on Electron Mobility on n-MOSFET's with Ultrathin
Gate Oxides," IEEE Transactions on Electron Devices 47, p 4 4 0 (2000).
[28] S. Datta, F. Assad, M. Lundstrom, "The Silicon MOSFET from a Transmission
Viewpoint," Superlattices and Microstructures 23, p7 7 1 (1998).
[29] F. Assad, Z. Ren, D. Vasileska, S. Datta, M. Lundstrom, "Modeling On-Currents for
n-MOSFETs: Ultimate Limits vs. the NTRS", Proceedings of the 1999 International
Conf. on Modeling and Simulation of Microsystems, San Juan, Puerto Rico, p388
(1999).
[30] Mark Lundstrom, personal communication, March 2001.
[31] M. Liang, J. Choi, P. Ko, C. Hu, "Inversion-Layer Capacitance and Mobility of Very
Thin Gate-Oxide MOSFET's," IEEE Transactions on Electron Devices ED-33, p409
(1986).
[32] M. Pinto, E. Sangiorgi, J. Bude, "Silicon MOS Transconductance Scaling into the
Overshoot Regime," IEEE Electron Device Letters 14, p3 7 5 (1993).
[33] T. Mizuno and R. Ohba, "Physical Limitations and Design for Sub-0.1gm MOS
Devices: Carrier Velocity Overshoot and Performance Fluctuations," Electronics
and Communications in Japan, Part2, 81 (8), p18 (1998).
[34] K. Rim, J. Hoyt, J. Gibbons, "Fabrication and Analysis of Deep Submicron StrainedSi N-MOSFETs," IEEE Transactions on Electron Devices 47, p 14 0 6 (2000).
[35] J. Brews, W. Fichtner, E. Nicollian, S. Sze, "Generalized Guide for MOSFET Miniaturization", IEEE Electron Device Letters 1, p2 (1980).
[36] Y. Ohkura, "Quantum Effects in Si n-MOS Inversion Layer at High Substrate Concentration," Solid State Electronics 33, p1581 (1990).
128
[37] M. Sherony, L. Su, J. Chung, D. Antoniadis, "SOI MOSFET Effective Channel
Mobility," IEEE Transactions on Electron Devices 41, p276 (1994).
[38] C. Huang, J. Faricelli, N. Arora, "A New Technique for Measuring MOSFET Inversion Layer Mobility," IEEE Transactions on Electron Devices 40, p1 13 4 (1993).
[39] H. Wong, D. Frank, P. Solomon, "Device Design Considerations for Double-Gate,
Ground-Plane, and Single-Gated Ultra-Thin SOI MOSFET's at the 25 nm Channel
Length Generation," IEEE IEDM Technical Digest, p407 (1998).
[40] A. Wu, T. Chan, V. Murali, S. Lee, J. Nulman, M. Garner, "Nitridation Induced Surface Donor Layer in Silicon and its Impact on the Characteristics of n- and p-Channel MOSFET's," IEEE IEDM Technical Digest, p271 (1989).
[41] H. Momose, S. Kitagawa, K. Yamabe, H. Iwai, "Hot Carrier Related Phenomena
For n- and p- MOSFETs with Nitrided Gate Oxide by RTP," IEEE IEDM Technical
Digest, p267 (1989).
[42] T. Ghani, K. Mistry, P. Packan, S. Thompson, M. Stettler, S. Tyagi, M. Bohr, "Scaling Challenges and Device Design Requirements for High Performance Sub-50 nm
Gate Length Planar CMOS Transistors," VLSI Technology Symposium, Digest of
Technical Papers, p174 (2000).
[43] L. Su, J. Jacobs, J. Chung, D. Antoniadis, "Deep-Submicrometer Channel Design in
Silicon-on-Insulator (SOI) MOSFET's," IEEE Electron Device Letters 15, p3 6 6
(1994).
[44] D. Esseni, M. Mastrapasqua, G. Celler, F. Baumann, C. Fiegna, L. Selmi, E. Sangiorgi, "Low Field Mobility of Ultra-Thin SOI N- and P-MOSFETs: Measurements
and Implications on the Performance on Ultra-Short MOSFETs," IEEE IEDM Technical Digest, p671 (2000).
[45] C. Sah, T. Ning, L. Tschopp, "The Scattering of Electrons by Surface Oxide Charges
and by the Lattice Vibrations at the Si-Si02 Interface," Surface Science 32, p561
(1972).
[46] S. Sun and J. Plummer, "Electron Mobility in Inversion and Accumulation Layers on
Thermally Oxidized Silicon Surfaces," IEEE Transactions on Electron Devices ED27, p1497 (1980).
[47] T. Ando, A. Fowler, F. Stern, "Electronic Properties of Two-Dimensional Systems,"
Reviews of Modern Physics 54, p4 3 7 (1982).
[48] D. Jeon and D. Burk, "MOSFET Electron Inversion Layer Mobilities--A Physically
Based Semi-Empirical Model for a Wide Temperature Range," IEEE Transactions
on Electron Devices 36, p14 5 6 (1989).
129
[49] S. Takagi, A. Toriumi, M. Iwase, H Tango, "On the Universality of Inversion Layer
Mobility in Si MOSFETs: Part I--Effects of Substrate Impurity Concentration,"
IEEE Transactions on Electron Devices 41, p2357 (1994).
[50] R. Pierret, FieldEffect Devices, Reading, Massachusetts: Addison-Wesley (1990).
[51] Y. Tsividis, Operation and Modeling of the MOS Transistor,New York: McGrawHill (1987).
[52] C. Sodini, T. Ekstedt, J. Moll, "Charge Accumulation and Mobility in Thin Dielectric MOS Transistors," Solid-State Electronics 25, p8 3 3 (1982).
[53] Z. Lee, M. McIlrath, D. Antoniadis. "Two-Dimensional Doping Profile Characterization of MOSFETs by Inverse Modeling using I-V Characteristics in the Subthreshold Region," IEEE Transactions on Electron Devices 46, p 16 4 0 (1999).
[54] M. Rodder, Q. Hong, M. Nandakumar, S. Aur, J. Hu, I. Chen, "A Sub-0.18 gm Gate
Length CMOS Technology for High Performance (1.5V) and Low Power (IV),"
IEEE IEDM Technical Digest, p563 (1996).
[55] H. Liu, J. Bums, C. Keast, P. Wyatt, "Thin Silicide Development for Fully-Depleted
SOI CMOS Technology," IEEE Transactions on Electron Devices 45, p1099 (1998).
[56] S. Takagi, M. Iwase, A. Toriumi, "On the Universality of Inversion-Layer Mobility
in N- and P-Channel MOSFETs," IEEE IEDM Technical Digest, p3 9 8 (1988).
[57] M. Kondo and H. Tanimoto, "An Accurate Coulomb Mobility Model for MOS
Inversion Layer and its Application to NO-Oxynitride Devices," IEEE Transactions
on Electron Devices 48, p2 6 5 (2001).
[58] M. Fischetti and S. Laux, "Performance Degradation of Small Silicon Devices
Caused by Long-Range Coulomb Interactions," Applied Physics Letters 76, p2 2 7 7
(2000).
[59] R. Ohba and T. Mizuno, "Nonstationary Electron/Hole Transport in Sub-0.1 gm
MOS Devices: Correlation with Mobility and Low-Power CMOS Applications,"
IEEE Transactions on Electron Devices 48, p3 3 8 (2001).
[60] 0. Tufte and E. Stelzer, "Piezoresistieve Properties of Heavily Doped n-Type Silicon", Physical Review 133, pA1705 (1965).
[61] Y Kanda, "A Graphical Representation of the Piezoresistive Coefficients in Silicon," IEEE Transactions on Electron Devices, ED-29 p6 4 (1982).
[62] W. Coen and R. Muller. "Velocity of Surface Carriers in Inversion Layers on Silicon," Solid-state Electronics 23, p3 5 (1980).
130
[63] T. Yamada, J. Zhou, H. Miyata, D. Ferry, "In-Plane Transport Properties of Si/Sij_
XGeX Structure and its FET Performance by Computer Simulation," IEEE Transactions on Electron Devices 41, p15 13 (1994).
[64] D Caughey and R. Thomas, IEEE Proceedings 55, p2 19 2 (1967).
[65] L. Zanfi, A. Losi, C. Jacoboni, C. Canali, "Impurity Effect on High-Field Transport
Properties of Electrons in Silicon," IEEE Transactions on Electron Devices ED-24,
p2 8 1 (1977).
[66] P. Bagwell, D. Antoniadis, T. Orlando, "Quantum Mechanical and Nonstationary
Transport Phenomena in Nanostructured Silicon Inversion Layers," in Advanced
MOS Device Physics, San Diego: Academic Press (1989).
[67] C. Jacoboni, C. Canali, G. Ottaviani, A. Quaranta, "A Review of Some Charge
Transport Properties of Silicon," Solid-State Electronics 20, p7 7 (1977).
[68] L. Chang, S. Tang, T. King, J. Bokor, C. Hu, "Gate Length Scaling and Threshold
Voltage Control of Double-Gate MOSFETs," IEEE IEDM Technical Digest, p719
(2000).
[69] Z. Ren, R. Venugopal, S. Datta, M. Lundstrom, D. Jovanovic, J. Fossum, "The Ballistic Nanotransistor: A Simulation Study," IEEE IEDM Technical Digest, p715
(2000).
[70] H. Shimada, T. Ushiki, Y. Hirano, T. Ohmi, "Tantalum-Gate SOI MOSFETs Featuring Excellent Threshold Voltage Control in Low-Power Applications," IEEE SOI
Conference Proceedings, p9 6 (1995).
[71] K. Yamada and K. Tomita, "Formation of Metal Silicide-Silicon Contact with
Ultralow Contact Resistance by Silicon-Capping Silicidation," Applied Physics Letters 64, p 3 4 4 9 (1994).
[72] Y. Taur and T. Ning, Fundamentalsof Modern VLSI Design, Cambridge, UK: Cambridge University Press (1998).
[73] D. Frank, P. Solomon, S. Reynolds, J. Shin, "Supply and Threshold Voltage Optimization for Low Power Design," International Symposium on Low Power Electronics
and Design, Proceedings, p317 (1997).
[74] K. Nakajima, Y. Akasaka, M. Kaneko, M. Tamaoki, Y. Yamada, T. Shimizu, Y.
Ozawa, K. Suguro, "Work Function Controlled Metal Gate Electrode on Ultrathin
Gate Insulators," VLSI Technology Symposium, Digest of Technical Papers, p95
(1999).
131
[75] C. Lombardi, S. Manzini, A. Saporito, M. Vanzi, "A Physically Based Mobility Model for
Numerical Simulation of Nonplanar Devices," IEEE Transactions on Computer-Aided
Design 7, p 1 1 6 4 (1988).
[76] M. Darwish, J. Lentz, M. Pinto, P. Zeitzoff, T. Krutsick, H. Vuong, "An Improved
Electron and Hole Mobility Model for General Purpose Device Simulation," IEEE
Transactions on Electron Devices 44, p1529 (1997).
[77] H. Wong, K. Chan, Y Taur, "Self-Aligned (Top and Bottom) Double-Gate MOSFET
with a 25 nm Thick Silicon Channel," IEEE IEDM Technical Digest, p427 (1997).
[78] J. Chung, M. Jeng, J. Moon, A. Wu, T. Chan, P. Ko, C. Hu, "Deep-Submicrometer
MOS Device Fabrication using a Photoresist-Ashing Technique," IEEE Electron
Device Letters 9, p 1 8 6 (1988).
[79] D. Hisamoto, W. Lee, J. Kedzierski, E. Anderson, H. Takeuchi, K. Asano,
T. King, J. Bokor, C. Hu, "A Folded-Channel MOSFET for Deep-Sub-Tenth Micron
Era," IEEE IEDM Technical Digest, p1032 (1998).
[80] S. Wolf and R. Tauber, Silicon Processingfor the VLSI Era: Volume 1--Process
Technology, Sunset Beach, CA: Lattice Press (1986).
[81] T. Tanaka, H. Horie, S. Ando, S. Hijiya, "Analysis of P+ Poly Si Double-Gate ThinFilm SOI MOSFETs," IEEE IEDM Technical Digest, p 6 8 3 (1991).
[82] A. Wei, "Device Design and Process Technology for Sub-100 nm SOI MOSFETs,"
Ph.D. Dissertation, MIT (2000).
[83] I.Yang, C. Vieri, A. Chandrakasan, D. Antoniadis, "Back Gated CMOS on SOIAS
for Dynamic Threshold Voltage Control," IEEE IEDM Technical Digest, p877
(1995).
[84] I. Yang, A. Lochtefeld, D. Antoniadis, "Silicon-On-Insulator-with-Active-Substrate
(SOIAS) Technology," IEEE SOI Conference Proceedings, p106 (1996).
[85] J. Lee, G. Taraschi, A. Wei, T. Langdo, E. Fitzgerald, D. Antoniadis, "Super SelfAligned Double-Gate (SSDG) MOSFETs Utilizing Oxidation Rate Difference and
Selective Epitaxy," IEEE IEDM Technical Digest, p71 (1999).
[86] B. Roberds, E. Whang, A. Rudolph, B. Doyle, "Investigation of a Novel SelfAligned Dual Gate MOSFET Structure," IEEE SOI Conference Proceedings, p109
(1998).
[87] H. Wong, D. Frank, Y Taur, J. Stork, "Design and Performance Considerations for
Sub-0.1 Rm Double-Gate SOI MOSFETs," IEEE IEDM Technical Digest, p747
(1994).
132
[88] E. Moon, P. Everett, H. Smith, "Immunity to Signal Degradation by Overlayers
using a Novel Spatial-Phase-Matching Alignment System," Journal of Vacuum Science and Technology B 13, p2 6 4 8 (1995).
[89] L. Su, M. Sherony, H. Hu, J. Chung, D. Antoniadis, "Optimization of Series Resistance in Sub-0.2 pm SOI MOSFET's," IEEE Electron Device Letters 15, p36 3
(1994).
[90] J. Kedzierski, P. Xuan, E. Anderson, J. Bokor, T. King, C. Hu, "Complementary Silicide Source/Drain Thin-Body MOSFETs for the 20 nm Gate Length Regime," IEEE
IEDM Technical Digest, p57 (2000).
[91] M. Cao, T. Kamins, P. Vande Voorde, C. Diaz, W. Greene, "0.18 pm Fully-Depleted
Silicon-On-Insulator MOSFETs," IEEE Electron Device Letters 18, p2 5 1 (1997).
[92] M. Goulding, "The Selective Epitaxial Growth of Silicon," Materials Science and
Engineering B17, p4 7 (1993).
[93] G. Niu, J. Cressler, S. Mathew, S. Subbanna, "A Total Resistance Slope-Based Effective Channel Mobility Extraction Method for Deep Submicrometer CMOS Technology," IEEE Transactions on Electron Devices 46, p1 9 1 2 (1999).
[94] R. Pierret, Advanced Semiconductor Fundamentals,Reading, Massachusetts: Addison-Wesley (1989).
[95] S. Wang, Fundamentalsof Semiconductor Theory and Device Physics, Englewood
Cliffs, NJ: Prentice Hall (1989).
133
134
Appendix A
Simulation Techniques
In this work we use both drift-diffusion (DD) and energy balance (EB) simulations. For
the former, solutions are found numerically to Poisson's equation coupled with the Driftdiffusion equation, which for electrons can be expressed as [94]:
Jn(x) = qgnEx + qDnVn
where n is electron density and Jn, 9n, and Dn are electron current density, mobility and
diffusivity, respectively. N is related to lateral electric field Ex through the Caughey-Thomas relationship [64] which accounts for velocity saturation effects:
i-to
p(E )
=
1+
E 2
vs at
In this case velocities higher than vsat are not permitted. While DD simulations are in
many circumstances sufficient for modeling MOSFET electrostatics, in scaled devices
with high drain bias carrier velocity (and hence drain current) may be severely under-predicted. As discussed in Section 2.2 we know the restriction ve
vsat to be non-physical in
deeply-scaled MOSFETs. Electrons transitioning abruptly from low to high lateral electric field regions can be accelerated to beyond vsat for a time characterized by t',, the
energy relaxation time [66].
t, is the decay time for carriers with kinetic energy values
135
out of equilibrium with the semiconductor lattice: this situation can be modeled the
Energy balance approach. In this case EX in the Caughey-Thomas expression is replaced
by an "effective" electric field ET, which is related to the local homogeneous carrier temperature Tc according to [95]:
2
p(E
)E-
3
kTC
kTe
2T
q
q
=
where k is Boltzmann's constant and To the lattice temperature. A fuller description of the
DD and EB models, their relation to the Boltzmann transport equation, can be found in
[66]. A full description of the relations between electric field, carrier and lattice temperature, and current density can be found in [13].
Neither of these modeling approaches take into account the thermal limit to source-tochannel injection velocity discussed in Chapter 2. For the EB model this can lead to serious over-prediction of carrier velocity and drive current in MOSFETs if Ty is not properly
chosen. As noted in Section 5.4, we have found that for a wide range of channel lengths
(down to Leff
40 nm) in an advanced NMOS technology, Ion/off characteristics can be
matched closely to experimental values for T, = 0.11 ps.
136
Download