You are on page 1of 14

Universal tunneling behavior in technologically relevant P/N junction diodes

Paul M. Solomon, Jason Jopling, David J. Frank, Chris D’Emic, O. Dokumaci et al.

Citation: J. Appl. Phys. 95, 5800 (2004); doi: 10.1063/1.1699487


View online: http://dx.doi.org/10.1063/1.1699487
View Table of Contents: http://jap.aip.org/resource/1/JAPIAU/v95/i10
Published by the AIP Publishing LLC.

Additional information on J. Appl. Phys.


Journal Homepage: http://jap.aip.org/
Journal Information: http://jap.aip.org/about/about_the_journal
Top downloads: http://jap.aip.org/features/most_downloaded
Information for Authors: http://jap.aip.org/authors

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
JOURNAL OF APPLIED PHYSICS VOLUME 95, NUMBER 10 15 MAY 2004

Universal tunneling behavior in technologically relevant


PÕN junction diodes
Paul M. Solomona)
IBM, SRDC; T.J. Watson Research Center, Yorktown Heights, New York 10598
Jason Jopling
ECE Department, Duke University, Durham, North Carolina 27708
David J. Frank and Chris D’Emic
IBM SRDC, T.J. Watson Research Center, Yorktown Heights, New York 10598
O. Dokumaci and P. Ronsheim
IBM, Microelectronics Division; Hopewell Junction, New York 12533
W. E. Haensch
IBM SRDC; T.J. Watson Research Center, Yorktown Heights, New York 10598
共Received 21 October 2003; accepted 18 February 2004兲
Band-to-band tunneling was studied in ion-implanted P/N junction diodes with profiles
representative of present and future silicon complementary metal–oxide–silicon 共CMOS兲 field
effect transistors. Measurements were done over a wide range of temperatures and implant
parameters. Profile parameters were derived from analysis of capacitance versus voltage
characteristics, and compared to secondary-ion mass spectroscopy analysis. When the tunneling
current was plotted against the effective tunneling distance 共tunneling distance corrected for band
curvature兲 a quasi-universal exponential reduction of tunneling current versus, tunneling distance
was found with an attenuation length of 0.38 nm, corresponding to a tunneling effective mass of
0.29 times the free electron mass (m 0 ), and an extrapolated tunneling current at zero tunnel distance
of 5.3⫻107 A/cm2 at 300 K. These results are directly applicable for predicting drain to substrate
currents in CMOS transistors on bulk silicon, and body currents in CMOS transistors in
silicon-on-insulator. © 2004 American Institute of Physics. 关DOI: 10.1063/1.1699487兴

I. INTRODUCTION more difficult partially or fully depleted SOI technologies, so


being able to accurately predict these currents is of consid-
Band-to-band tunneling presents a limit to scaling of fu- erable importance.
ture complementary metal-oxide-silicon 共CMOS兲 devices,1 In contrast to the urgent requirements for evaluation of
both in bulk CMOS, where a heavily doped P/N junction scaled CMOS, band-to-band existing tunneling data are very
exists between the drain of the transistor and the substrate, sparse. Most of it is out-of-date, hard to interpret, and inap-
and in partially depleted silicon-on-insulator CMOS,2 be- plicable to the types of profiles and voltages found in modern
tween the drain and the partially depleted silicon body. The FETs. For instance much of the older data5–7 is on 共111兲
problem is exacerbated because very heavy counter-doping silicon, other data are exclusively in the forward
direction6,8 –10 while we choose the reverse direction because
共the halo兲1 in the vicinity of the source or drain is used to
it is the worst case for high leakage currents. We do not
suppress short-channel effects. In the design of Taur et al.1 a
attempt to study tunneling in FET structures,11–13 since the
halo doping of close to 1019 cm⫺3 was used, and this doping field geometry in a FET is complex and we are interested
is predicted to scale as the square of the dimensions3 to first in establishing the tunneling parameters using simpler
maintain proportionately smaller depletion layer widths as geometries. The work of Stork and Isaac,14 is relevant to our
dimensions shrink and voltages remain constant. The tunnel- study, although doping values used there only overlap our
ing leakage in the heavily doped junction is a cause for in- study at the higher end of their data and the lower end of
creased power dissipation in the bulk field-effect transistor ours. That study was used by Taur et al.1 to establish a cur-
共FET兲 whereas in the partially depleted silicon-on-insulator rent density versus electric field relationship, but in fact it is
共SOI兲 case it can cause unwanted threshold voltage shifts difficult to extract such a relationship unambiguously from
leading to extra subthreshold drain-to-source leakage. This this data. The work of Hurkx et al.15 has sufficient documen-
tunneling sets a limit to the minimum gate length of the FET tation to provide a valuable complement to our data.
subject to a constraint on standby power4 and could deter- In silicon, tunneling theory is messy because of the in-
mine the choice between a bulk-silicon technology, or the direct and complex band structure, leading to many compet-
ing tunneling paths which are difficult to quantify. The mo-
tivation for this work therefore is not so much to verify basic
a兲
Electronic mail: solomonp@us.ibm.com tunneling theory, which has been amply done before, in sim-

0021-8979/2004/95(10)/5800/13/$22.00 5800 © 2004 American Institute of Physics

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al. 5801

pler systems, but to establish a database of empirical infor-


mation which will be directly useful in evaluating scaling
trends, as well as indirectly in the evaluation of numerical
models for device simulation. While our empirical approach
is acknowledged, the tunneling theory will be used as a
framework to attempt to unify the data and to seek universal
behavior. The literature is split on the issue of universality
with some early data giving it support.6,7 Much of the later
data11,12,14 –16 have focused on trap-assisted tunneling, which,
being process dependent, is not expressible in universal form
even though intrinsic behavior may be present as well.14 –16
While some of our data shows trap-assisted type behavior,
we will show that a large subset does indeed exhibit univer-
sal characteristics and is therefore suitable for quantifying
the limits of devices. It is to be noted that our study was
restricted to 共100兲 silicon, and that tunneling paths may be in
other directions inside of a FET, as well as the fact that FETs
FIG. 1. Schematic band diagram of a P–N junction under reverse bias, and
are commonly laid out along the 具110典 directions on a 共100兲 energy-momentum picture for indirect tunneling, where V is the internal
surface, therefore future study on other crystal orientations is potential, V EXT is the external voltage, V BI is the built-in voltage, V D is the
warranted. potential across the junction, and V b is the band gap potential. E C and E V
A summary of our work, including technological impli- are the conduction and valence band edges, ⌫ and ⌾ are crystal momenta,
and q is the phonon momentum. The minimum tunneling distance is w T,min
cations, has been presented elsewhere.17
and V FWHM is the width of the energy distribution of the tunneling current.

II. THEORETICAL METHODS tunnel distance, w T ⫽x V ⫺x C , corresponding to an internal


A. Effective tunneling distance potential V I is found by solving for the conduction and
valence-band intercepts, x V and x C , i.e., V(x C )⫽V I , and
The tunneling distance is the most important parameter V(x V )⫽V I ⫹V b , where V b is the band gap potential of sili-
for unifying I – V data in the band-to-band tunneling regime, con (E G /e). Furthermore, by assuming, in advance, an ex-
over a wide range of sample types, doping, temperature, and ponential dependence of tunnel current on tunnel distance,
voltage. This is because tunneling depends exponentially on (⬀ exp关⫺wT /␭T兴), we obtain a full width at half maximum
the tunneling distance, i.e., the shortest equienergy path be- spread, V FWHM , of the tunnel current distribution in energy
tween conduction and valence bands, give or take an acous- 共see Fig. 1兲, and the average tunnel distance 具 w T 典 . The tun-
tic zone-edge phonon 共⬃17 meV for TA phonon in Si兲.18 It nel decay length, ␭ T , for the purposes of averaging, is cho-
also depends exponentially on the band gap, but the band gap sen to be consistent with the, to be determined, dependence
is not strongly dependent on the experimental variables. A of the current density on tunnel distance 共⬃0.5 nm兲. This
schematic diagram of the tunneling situation is shown in Fig. process is iterative but in practice the averaging process de-
1. The phonon-assisted models of Keldysh,19 Price and pends rather weakly on the a priori assumed ␭ T .
Radcliffe,20 and Kane21 are applicable 关see Fig. 1共b兲兴, where Rather than just the tunneling distance, it is actually the
an electron and a hole tunnel in from the valence and con- action integral, in the WKB approximation, which deter-
duction bands, respectively, recombining near midgap 共for mines the tunneling current. Denoting this integral by ␸, the
comparable masses兲 with the aid of a phonon. To the degree tunneling probability is proportional to e ⫺2 ␸ 共square of the
of approximation of our analysis we assume equal electron tunneling amplitude兲. For our phonon-assisted process the
and hole masses 共cf. light electron and hole masses of tunneling path is split into the electron part (x C ⭐x⬍x p ) and
0.19m 0 and 0.16m 0 , respectively兲,18 and we neglect the pho- the hole part (x p ⭐x⭐x V ) where x p is the most probable
non energy which is small compared to the band gap. coordinate for the phonon transition, which is taken to be the
The internal potential V(x) is derived from the profile midgap point, for equal electron and hole masses. Using the
N(x) using the depletion approximation WKB approximation5

V共 x 兲⫽
e

冕x1
x
共 u⫺x 兲 N 共 u 兲 du, 共1兲 ␸⫽
1
ប 再冕 冑xC
xp
2m e e 关 V 共 x 兲 ⫺V I 兴 dx


x


coupled with the charge neutrality condition: 兰 x 2 N(u)du xV
⫽0, where x 1 and x 2 are the near and far-side depletion layer
1
⫹ 冑2m h 关 V b ⫹V I ⫺V 共 x 兲兴 dx , 共3兲
xp
edges, e is the electronic charge, and ⑀ is the permittivity of
silicon. The additional condition where m e and m h are the electron and hole effective masses
共assumed to be equal here兲, and ប is the reduced Planck’s
V 共 x 2 兲 ⫽V EXT⫹V BI , 共2兲
constant. Denoting ␸ F for the case where the field is uniform
where V EXT and V BI are the external 共applied兲 and built-in 关 V(x)⫽Fx 兴 , ␸ F ⫽(2/3ប) 冑emV b w T , where w T is as defined
voltages, respectively, allows the system to be solved. The before. For curved bands, as in Fig. 1, ␸ ⬍ ␸ F for the same

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
5802 J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al.

w T , and we compensate for this by plotting our data against three parameters to be determined, V b , N b0 , and ␭. This
an effective tunneling distance w TE , where w TE⫽w T ␸ / ␸ F . function has the virtue of requiring few fitting parameters,
We call the factor ␸ / ␸ F the curvature correction. To simplify having analytic solutions, and being able to emulate profiles
calculations the curvature correction is evaluated only at the varying from step to linearly graded. This functional form is
V I corresponding to the maximum tunneling probability. justified when the junction is on the exponentially falling
The mean effective mass for tunneling, m, may now be slope of the top dopant, and the depletion layer is narrow
derived from the slope of the ln J vs w TE plot enough that the more slowly varying bottom dopant profile

冋 册
approximates a constant. We shall see later that even under
9ប 2 d 共 ln J 兲 2
m⫽ . 共4兲 conditions where this function is not a good approximation
16eV b dw TE to the actual profile, it still determines tunneling distances
with remarkable accuracy.
B. Profile extraction
While chemical profiles can be obtained using
secondary-ion mass spectroscopy 共SIMS兲, as we have done C. Built-in voltage, band gap, and the depletion
in some cases, it is actually very difficult to obtain the pro- approximation
files to the accuracy required 共⬃1 nm兲 for extracting tunnel-
ing distances. Furthermore, SIMS gives the chemical rather The built in voltage, V BI⫽V b ⫺⌬V BI , for a graded junc-
than the electrically active profile, so that cross calibration tion is a function of applied voltage through the dependence
with an electrical profiling method is necessary. In this study of the doping at the depletion layer edge, N D and N A , on the
we use a capacitance–voltage (C – V) technique to extract depletion-layer width. For a nondegenerately doped junction

冋 冉 冊册
the profile information.
kT N CN V
The well known C – V method for extracting a profile of ⌬V BI⫽ • 2⫹ln , 共8兲
a one-sided junction gives the information e N DN A

CdV where N C and N V are the conduction and valence-band den-


N ⬘共 w 兲 ⫽ , 共5兲 sities of states and, in the depletion approximation, the elec-
⑀ Adw
tron and hole concentrations are assumed to be the same as
where C is the capacitance, A is the area, and, w⫽ ⑀ A/C, the the doping densities at the depletion-layer edge. The leading
depletion layer width 共see also Sec. II C, later兲 and N ⬘ (w) ‘‘2’’ in this equation comes from the well known 2kT/e
the doping profile assuming one-sidedness. For a two-sided correction18 applied to the depletion approximation to ac-
junction with w⫽x 2 ⫺x 1 , and doping concentration N(x 1 ) count for the non abrupt carrier falloff at the depletion-layer
and N(x 2 ) at the near and far sides of the junction, the fol- edges. While ⌬V BI is quite small in the doping range used
lowing relationship holds: for these experiments, 共⬃50 meV兲 it nevertheless leads to
1 1 1 large errors in determining the band gap if one neglects it.
⫽ ⫹ . 共6兲 The main error arises because ⌬V BI becomes quite large in
N ⬘共 w 兲 N 共 x 1 兲 N 共 x 2 兲
the forward bias regime, where the doping is low, and which
The C – V measurements alone are therefore insufficient to has a predominant weight in determining the built-in
extract the profile of a two-sided junction without additional voltage.22 For a linearly graded junction, Chawla and
assumptions. By fitting the doping concentration to some pa- Gummel18,23 investigated this effect and found that it re-
rameterized function one could find the parameters by fitting sulted in a lowering of about 0.1 V of the ‘‘intercept’’ voltage
measured and simulated C – V curves, but solution of these (V-axis intercept of the C ⫺3 vs V plot兲.
equations would not lead to a unique solution. For instance, For analysis of our experiments we use a variable ⌬V BI
one is unable to unambiguously determine the degree of so the Chawla–Gummel effect is taken into account auto-
asymmetry of the profile this way, so that a physically justi- matically. For a particular C – V sweep we assume that V b is
fiable trial function has to be used as the starting point. Hav- the constant parameter, to be extracted from the optimization
ing determined the parameters, the fitted profile could be routine, and that V⫽V ext⫹V b ⫺⌬V BI , where V ext is the ap-
used beyond the leakage current limited range of the C – V plied 共external兲 voltage, ⌬V BI is determined using Eq. 共8兲
measurements. and the varying values of N D and N A are derived from the
In this work we adopt two trial functions for determina- profile being fitted. As a consequence of the variable ⌬V BI ,
tion of the profile. When SIMS data are available they are the capacitance measured externally, C ext , is different from
scaled by a fitting factor A E 共electrical activation factor兲 and the internal capacitance, C, because some of the ac voltage
used, along with the built-in voltage, which is related to the goes into changing ⌬V BI . This results in a correction to the
bandgap potential, V b , 共see next subsection兲, to fit the C – V depletion-layer width
data.
The other trial function is ⑀ ⑀
w D⫽ ⫽ , 共9兲
N 共 x 兲 ⫽N b0 关 1⫺e ⫺x/␭ 兴 , 共7兲 C C ext共 1⫹ ␰ 兲

where the parameter ␭ represents the exponential falloff rate where ␰ ⫽d⌬V BI /dV. In a positively graded junction ␰ is
of the top dopant layer and N b0 is the concentration in the negative, which has the counter-intuitive result that the inter-
bottom layer far away from the junction. Thus, there are nal potential varies faster that the external potential so that

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al. 5803

C ext⬎C. In the same spirit, a correction term of (1⫹ ␰


⫹w D d ␰ /dw D ) must be applied to the doping derived from
Eqs. 共5兲 and 共6兲.
For the profile fitting we assumed non-degenerate dop-
ing for the ⌬V BI corrections, yet the carrier concentrations at
the depletion layer edge enter into the degenerate regime for
our higher doped samples. Degeneracy has two opposing ef-
fects on the built-in voltage. The direct effect of higher Fermi
levels in conduction and valence bands increases the built-in
voltage; yet the indirect effect, which is larger in our case,
decreases the built-in voltage because of a softening of the
carrier falloff rate 关the 2kT term in Eq. 共8兲兴 at the depletion FIG. 2. Sample structure for the case of a P/N diode. Opposite dopant type,
layer edges. While retaining the simpler nondegenerate including substrate, would be used for N/P diode.
analysis for our optimization procedure, we make an accom-
modation for the indirect effect of degeneracy by replacing
the temperature, T in Eq. 共8兲, with an effective temperature, ⬃7⫻1019 cm⫺3 in eight steps. Measurements were done ex-
T e , which is given by the inverse logarithmic slope of the clusively in a low voltage range 共1 V forward to 1.5 V re-
doping density versus Fermi level verse兲 which represents the range of interest for scaled
T 1 CMOS. A temperature range from 100 to 350 K was studied.
⫽ F 关 F⫺1 共 y 兲兴 , 共10兲 Lower temperatures suffer from the complications of freeze-
T e 冑2 ␲ y ⫺1/2 1/2 out, yet a moderately low temperature aids in analysis of the
where Fn is the nth order Fermi integral, and a mean doping data in removing overt effects of the thermal distribution of
factor y⫽(N D N A /N C N V ) 1/2 is used. This procedure, while the carriers, as well as reducing the leakage for the C – V
plausible, is a heuristic, i.e., lacks a rigorous proof. measurements. Temperatures of 150 and 300 K were chosen
As a check on the above methods, the results of our for presentation of the data, although data were measured
C – V analysis are compared against numerical simulations and analyzed at all 共six兲 temperatures and for all samples.
using the FIELDAY24 program, which includes full, degenerate A. Sample preparation
statistics. C – V curves are simulated using FIELDAY and the
profiles derived from our earlier analysis, and the two sets of The sample structure is shown in Fig. 2. N- or P-type
C – V curves compared 共see Sec. IV A兲. substrates were used, having resistivities of 0.017–0.018 and
0.04 –0.05 ⍀ cm, respectively. This corresponds to doping
levels of 1.4⫻1018 and 1⫻1018 cm⫺3 , respectively, which is
III. EXPERIMENTAL METHODOLOGY
below the minimum junction doping used in this study. Ox-
Our approach is to use simple geometries and structures, ide isolated areas were created using local oxidation, with
which are easy to quantify, yet use sophisticated, modern, many different sized areas being available from the mask set,
ion-implanted structures, over a wide range of dose, which while the present experiments used mainly 10 and 50 ␮m
make our study relevant to present, and hopefully future, sized squares. The N- and P-type implants were then done
CMOS integrated circuit technology. Various types of im- through a 3.5 nm screen oxide. In all cases a germanium
plants were studied to investigate to what extent our data amorphizing implant of 3⫻1014 at 15 kev was done prior to
might be process independent. Planar, oxide-isolated, P/N the junction implants. Implant parameters were as given in
junctions were fabricated over a wide area range, on low- Table I. The devices were fabricated on six, 8 in., silicon
resistivity substrates, in order to reduce series resistance. wafers divided into three junction types with shallow As, B,
This study emphasizes the reverse leakage regime, which is or BF2 implants and deep As or B implants as given in Table
dominated by band-to-band tunneling. Only samples with I. Two wafers were allocated to each type and each wafer
strictly area independent reverse current densities were in- was divided into four quadrants giving eight quadrants 共Q1–
vestigated, and small sized samples were chosen to allow Q8兲 per junction type. Following the implants the wafers
large current densities to be measured without incurring ex- were given rapid thermal anneals for 5 s, to 1050 °C for the
cessive voltage drops due to series resistance. Capacitance N on P, and to 1000 °C for the P on N wafers. The lower
measurements were the key to obtaining information on the temperature for the P on N wafers was to minimize B diffu-
doping profile and evaluating internal fields and tunneling sion. Then a premetallization forming-gas anneal was done
distances. The measurements were done on small samples at 400 °C for 30 min. The screen oxide was then etched off
and at high frequencies to extend the measurement range to using a short HF dip, and Ti/Al was deposited followed by a
higher forward and reverse voltages and higher doping lev- 200 °C 30 min sinter. The temperature was chosen to reduce
els. SIMS measurements were done on selected samples to silicide spiking. The contact resistance was tested by mea-
bolster the C – V measurements. To obtain the highest accu- suring N/N and P/P samples of different areas, where the
racy, separate ion bombardment species, Cs vs O2 , were sample resistance was found to be limited by the spreading
used on some wafers to obtain the As vs B information. resistance to the substrate.
Samples were fabricated with a wide range of implant Results of SIMS analysis are shown in Fig. 3. A very
doses to obtain junction doping ranging from ⬃2⫻1018 to pronounced boron dip is seen in the N on P wafer. The reso-

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
5804 J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al.

TABLE I. Implant splits.

Top Bottom 共B or As兲

Dose (1014 cm⫺2 )


Energy Dose Energy
Type 共keV兲 (1015 cm⫺2 ) 共keV兲 Q1 Q2 Q3 Q4 Q5 Q6 Q7 Q8

NP 3 As:2.5 20 0.6 1.2 2 2.6 3.5 4 6.3 7.5


PN–B 2 B:1.0 100 0.21 0.43 0.71 0.93 85 1.32 2.31 3.3
PN–BF2 5 BF2 :1.0 100 0.21 0.43 0.71 0.93 85 1.32 2.31 3.3

lution of the dip, to this extent, is a measure of the quality of technique was used to minimize lead resistance. It was also
the SIMS technique. Since the N/P junction is formed in the found during measurements that the resistance of the back-
vicinity of the dip it is essential to obtain the highest accu- interface was significant, because of the small area of the
racy in the analysis on the N/P samples 关P/N are less critical cleaved, header-mounted piece. To eliminate this resistance,
as seen in Fig. 3共b兲兴. To achieve this the As was sputtered a spare sample was used as a probe of the substrate potential
using Cs and the B using oxygen for these samples. It is also in the four-point technique. Even though this was a P/N junc-
seen from Fig. 3共b兲 that the junctions are shallower and con- tion, the junction was sufficiently leaky to serve as a voltage
siderably sharper when BF2 is used as the implant species. contact for the higher doped samples. Measurements were
This reflects the fact that the kinetic energy of the B in BF2 done at temperatures from 100 to 350 K in ⬃50 K incre-
is only 1.1 keV. ments. The temperature accuracy was ⫾⬃2 K. Absolute cur-
Following processing, samples were screened to make rent versus voltage characteristics are shown for all three
sure the currents were proportional to area, and those se- doping types in Figs. 4共a兲– 4共c兲, for quadrants Q2, Q5, and
lected were mounted on eight-pin TO-five headers for further Q8, for temperatures 150 and 300 K and for square samples
measurement. of 10 and 50 ␮m size. In reverse bias, which is the focus of
this study, the curves for the two areas overlay except at very
B. I – V measurements high currents where substrate series resistance is significant,
and at very low currents where other sources of leakage may
I – V measurements on the header-mounted samples were dominate. The forward characteristics are very variable, and
made using the HP 4145B parameter analyzer. A four-point gross departures from area scaling occur in some instances.
This is because the band-to-band tunneling, which dominates
the reverse characteristics, is suppressed in the forward di-
rection, and trap dominated leakage mechanisms take over.
The lack of the negative resistance region in the forward
characteristics 共except for a hint of one for the PN–BF2
sample at low temperatures, attests to the difficulty of
achieving P/N junctions in silicon which are strongly degen-
erate on both sides, by ion implantation, due to the strong
interaction between the B and As dopants.25 This contrasts
with epitaxial studies,8 –10 at considerably higher doping lev-
els, where a well-developed negative resistance region is ob-
tained.

C. C – V measurements
C – V measurements were made with the Agilent 4294A
precision impedance analyzer. To minimize parasitics, two
leads were bonded onto each pad, enabling the four-probe
technique to be extended right onto the wafer. Special cali-
bration headers containing open, short and load 共100 ⍀兲
standards were provided, and calibration was done right
down to the header. Calibration files were prepared for each
temperature. Capacitance was measured at a frequency of 18
MHz, and at signal amplitude of 25 mV. For the highest
doped quadrants the smaller, 10 ␮m sized sample was mea-
sured to reduce the effects of series resistance. Parasitic ca-
pacitance was extracted by comparing the zero-bias capaci-
FIG. 3. SIMS profiles of 共a兲 N/P and 共b兲 P/N diodes for the indicated
quadrants. For the N/P profiles the As was sputtered with Cs. For all other
tance for 10 vs 50 ␮m samples. dc voltage drops across the
profiles O2 was used the boron and BF2 implanted profiles are compared in backside contact were accounted for by subtracting the volt-
共b兲. age drops determined during the I – V measurements. This

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al. 5805

FIG. 4. Current–voltage characteristics for the N/P 共a兲 and P/N diodes using
B 共b兲 and BF2 共c兲 implants, for quadrants Q2, Q5, and Q8, for temperatures FIG. 5. C – V characteristics for the N/P and P/N diodes using B and BF2
150 and 300 K and for square samples of 10 and 50 ␮m size. implants, for quadrants Q2, Q5, and Q8, for temperatures 150 共dashed兲 and
300 K 共solid兲 compared to simulated curves based on an exponential profile
共dotted兲.

IV. ANALYSIS AND RESULTS


resistance, being due to a Schottky-like contact to the metal-
lized backside of the chip, does not affect the capacitance A. Profile extraction
共apart from the debiasing兲 because of the large parallel ca- The doping profile is extracted from the C – V curves
pacitance of that contact. These techniques enabled the ca- using the methods discussed in Sec. II B, using the exponen-
pacitance 共⬃2 pF兲 to be measured accurately in the presence tial trial function for all quadrants, and the SIMS data for the
of up to ⬃1 mS of leakage conductance. upper four quadrants. The profiles for selected quadrants
C – V characteristics are shown for all three doping types 共Q2, Q5, and Q8兲 are shown in Fig. 6 共all quadrants were
in Figs. 5共a兲–5共c兲, for quadrants Q2, Q5, and Q8, and for measured but only selected data are shown here to avoid
temperatures 150 and 300 K. The turn-up of the experimental clutter兲. At high doping the exponential fit deviated markedly
curves at large reverse bias is due to leakage. Lower tem- from the SIMS fit, especially for the PN–B sample, yet all
peratures permit a somewhat larger range for significant give excellent fits to the C – V data. This is because of the
measurements to be achieved. inherent ambiguity in the profile symmetry, as discussed in

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
5806 J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al.

FIG. 7. Band gap, determined from the C – V fits, as a function of tempera-


ture. Line styles for wafer types are: N/P 共solid兲, P/N–B 共dashed兲 and
P/N–BF2 共dotted兲. Open symbols are fits to the exponential profiles and
filled symbols to the SIMS profiles. The band gap formula is from Sze 共see
Ref. 18兲.

esting to note that the temperature dependence of V b is sub-


stantially independent of the doping in this range, while V b
itself decreases with doping. There is close agreement be-
tween the two fitting techniques, notwithstanding the ⬃20
mV higher V b for the highest-doped quadrants on th N/P and
PN–BF2 wafers when using the SIMS based fitting function.
Some systematic differences between the two methods could
be due to perturbation of the doping near the junction, caused
for instance by the boron dip, which is not included in the
exponential fitting function.
C – V curves were simulated with the FIELDAY program
共see Sec. II B兲 and the SIMS profiles, using the previously
derived activation factors and band gaps as an initial guess.
C – V curves are shown in Fig. 8 for a particular sample
共NP,Q6兲 and fitting parameters for this sample are compared
in Table II. For all of the samples the two methods give
results which are almost identical except that FIELDAY re-
quires an ⬃20 mV lower band gap. No significant extra cor-
rections are required for the more degenerate cases 共low tem-
FIG. 6. Doping profiles derived from fitting the exponential trial function perature, high doping兲 indicating that our heuristic procedure
共䊉兲 and by scaling the SIMS data 共⫹兲 to C – V curves measured at 150 K.
Quadrants Q2, Q5, and Q8 are shown for 共a兲 the N/P, 共b兲 the P/N–B, and 共c兲
the P/N–BF2 wafers.

Sec. II B. This may be seen clearly in Fig. 6共b兲, at the highest


doping, where the strong compensation actually results in a
retrograde profile 共steeper on the substrate side兲 compared to
the assumed exponential profile. At the other extreme, for the
lowest doped quadrant, Q1 关also seen for Q2 in Fig. 6共b兲兴,
the depletion edge scans past the doping peak, causing some
difficulties in the use of the exponential profile, yet the peak
is broad, so that the profile does not deviate greatly from the
exponential form.
The band gap, derived from these fits, is shown in Fig. 7,
as a function of temperature. The temperature dependence is
less than the literature value,18 especially at lower tempera-
FIG. 8. C – V curves for a 10⫻10 ␮m2 N/P diode comparing FIELDAY 共see
tures, yet gives good agreement with the trend at higher tem-
Ref. 24兲 simulations with our model at temperatures of 150 and 300 K, for
peratures. The magnitude of the disagreement 共⬃50 meV兲 is a N/P sample using a SIMS derived profile. The band gap and activation, for
comparable with the uncertainty of our technique. It is inter- best fit, are compared in Table II for the two models.

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al. 5807

TABLE II. Simulation vs model parameters. TABLE IV. Doping parametersa for P/N–B wafers.

Band gap 共eV兲 Activation N b0 (1019 cm⫺3 ) ␭ 共nm兲 N b0 /␭ (1026 cm⫺4 )


T
共K兲 Our model FIELDAY Our model FIELDAY Q 153 K 296 K 153 K 296 K 153 K 296 K

150 0.968 0.943 0.986 0.977 1 0.23 0.24 2.69 2.93 0.09 0.08
300 0.961 0.935 0.966 0.967 2 0.42 0.42 2.95 3.89 0.14 0.14
3 0.77 0.80 4.75 4.96 0.16 0.16
4 0.97 1.02 4.81 5.01 0.20 0.20
5 1.50 1.45 5.37 4.94 0.28 0.29
6 8.89 10.07 24.19 26.55 0.37 0.38
is reasonable. It was hoped that the FIELDAY fits would in- 7 9.39 15.93 21.87 36.46 0.43 0.44
crease the band gap at lower temperatures to better agree 8 14.80 13.3 12.51 28.39 0.46 0.47
with the literature, but this was not the case.
Doping parameters for all quadrants, by C – V technique,
a
For trial function N(x)⫽N b0 关 1⫺exp(⫺x/␭)兴.
at two temperatures 共150 and 300 K兲, are shown in Tables
III–V. For most part the parameters are temperature indepen-
rived using the two techniques were very similar, as shown
dent 共within the accuracy of the technique兲 lending support
by plotting the difference between them in Fig. 9. The dif-
to our extraction method. Exceptions are in the case of high
ference diminishes for the lower-doped quadrants, as ex-
doping levels 共Q6 –Q8兲, where the profile approaches linear
pected, justifying our reliance on the exponential profile fit
grading, especially for the P/N cases. In such cases the pa-
for the lower-doped quadrants.
rameters, N b0 and ␭, are not separable, i.e., neither the as-
sumptions of the model, nor the accuracy of the data are
enough to distinguish between an almost linear-graded and B. Current density versus field
the assumed exponentially graded junction. However, the Our experimental I – V data were analyzed using the
gradient, N b0 /␭, is still a good parameter, and the only one C – V extracted profiles. For the lower doped wafers 共Q1–
needed here. Q4兲 where no SIMS data were available, the exponential
Activation parameters, and doping gradient at the junc- fitting function was used, and otherwise the SIMS derived
tion, for the SIMS profiles, are shown in Tables VI and VII. fitting function was used. Our results were also compared to
The accuracy of both the C – V extraction technique and the the results of Hurkx et al.15 using his zero-bias depletion
SIMS measurements is attested to by the reasonable values layer widths and substrate doping, along with our values for
of the activation parameters. An error of ⬃10%, in activa- V b , to derive profile parameters. Other work was not used
tion, in either technique would be quite apparent since acti- because either not enough information was available,14 or the
vations of greater than 100% are not possible and activations results were on 具111典 silicon.5,7
of less than 80%, for most of the conditions, are not likely. It The logarithm of the conductivity 共current–density/
is not clear what causes the lower activations for Q8 of the voltage兲 as a function of the inverse maximum junction field
PN–B and – BF2 wafers, but this is consistent with the lower is shown in Fig. 10. According to the most simple tunneling
annealing temperature 共1000 °C兲 for these wafers and the theory18 this should give an approximately linear relationship
fact that Q8 of these wafers corresponds to the highest As for a junction with uniform internal field, under reverse bias
doping condition. As seen in Fig. 3共b兲, these quadrants are since the tunneling distance is inversely proportional to the
the most highly compensated, which could easily lead to field, and the energy range for tunneling is proportional to
sizeable errors in the extraction of the net doping from the the voltage. Both our data and that of Hurkx et al.15 deviate
SIMS data. considerably from this relationship, and there is considerable
The profile data were used to extract junction fields and spread between the data from different wafers, dopings, and
effective tunneling distance, according to the methodology of sources. Furthermore, the trend is different from that used by
Sec. II A. It was remarkable to find that despite the obvious Taur et al.1 to predict future device behavior. Plotting just the
differences in the profile shape, the tunneling distances de- current density 关Fig. 10共b兲兴 brings our results closer to Taur’s

TABLE III. Doping parametersa for N/P wafers. TABLE V. Doping parametersa for P/N–BF2 wafers.

N b0 (1019 cm⫺3 ) ␭ 共nm兲 N b0 /␭ (1026 cm⫺4 ) N b0 (1019 cm⫺3 ) ␭ 共nm兲 N b0 /␭ (1026 cm⫺4 )

Q 153 K 296 K 153 K 296 K 153 K 296 K Q 153 K 296 K 153 K 296 K 153 K 296 K

1 0.50 0.52 2.29 2.39 0.22 0.22 1 0.31 0.34 3.94 3.78 0.08 0.09
2 1.01 1.03 2.80 2.79 0.36 0.37 2 0.60 0.62 4.40 4.35 0.14 0.14
3 1.39 1.38 2.31 2.15 0.60 0.65 3 1.05 1.08 5.50 5.46 0.19 0.20
4 1.68 1.70 2.27 2.18 0.74 0.78 4 1.27 1.35 5.30 5.57 0.24 0.24
5 2.22 2.24 3.37 3.25 0.66 0.69 5 1.81 2.21 4.32 5.34 0.42 0.41
6 1.97 1.88 2.28 2.19 0.86 0.86 6 6.81 6.65 13.27 12.33 0.51 0.54
7 2.74 2.84 2.38 2.45 1.15 1.16 7 15.05 15.39 22.70 23.07 0.66 0.67
8 8.17 11.56 7.16 9.97 1.14 1.16 8 6.09 9.05 8.64 12.72 0.70 0.71
a
For trial function N(x)⫽N b0 关 1⫺exp(⫺x/␭)兴. a
For trial function N(x)⫽N b0 关 1⫺exp(⫺x/␭)兴.

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
5808 J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al.

TABLE VI. Activation parameter for all wafers.

N/P P/N–B P/N–BF2

Q 153 K 296 K 153 K 296 K 153 K 296 K

5 1.04 1.07 0.87 0.90 0.79 0.83


6 0.96 0.94 0.83 0.86 0.91 0.95
7 0.90 0.91 0.93 0.96 0.90 0.91
8 0.92 0.92 0.63 0.65 0.53 0.54

trend line, yet the functional dependence of the individual


curves is obviously different. The differences can be ex-
plained by the lower voltages used in our study, which vio-
late the uniform field assumption, by the fact that Taur’s
curve is partly based of 具111典 data, and also to possible dif- FIG. 9. Difference in effective tunnel distance calculated from profiles ex-
ferences in the physics of the tunneling process, as we will tracted using SIMS and exponential trial functions.
discuss later.

C. Current density versus effective tunneling considerable differences in profile shape from wafer to wafer
distance
and quadrant to quadrant, and is anyway not very large, be-
Tunneling current is expected to be more closely corre- ing between 0.9 and 1.0 over most of the voltage range.
lated with the tunneling distance than the electric field, espe- Data measured for all three implant types, for the four
cially at small voltages where the field varies along the tun- highest doped quadrants 共Q5–Q8兲 and for temperatures of
neling path. Therefore, following Sec. II A, the current 150 and 300 K, are shown in Fig. 13. Fitting this data, at the
density was plotted as a function of the effective tunnel dis- high current end, to straight lines given by the expression
tance, which includes our curvature correction. Better unifi-
J⫽A 共 T 兲 exp共 w TE /␭ T 兲 , 共11兲
cation of the data, and more linear plots were obtained when
choosing to plot current density rather than conductance, or
even the current normalized to the half width of the tunnel-
ing energy range 共see Fig. 1兲, so that this method of presen-
tation was chosen. Results at 300 and 150 K are shown in
Fig. 11, where a very satisfying unification of the data is
achieved especially for the higher currents, showing a rate of
falloff in tunneling current with tunneling distance of ap-
proximately one decade per nanometer. Comparison of our
data with Hurkx et al.,15 as shown in the figure, is also very
reasonable, where deviations could well be due to our lack of
knowledge of their precise doping profile. Obvious sample-
to-sample deviations from the linear trend are seen at the
lower current densities, where other leakage mechanisms
may be operative, especially for some of the lower doped
quadrants of the PN–B and PN–BF2 wafers where the leak-
age is anomalously high. At higher currents the distribution
for these wafers is very tight, with the curves essentially
overlaying each other for the upper four quadrants, giving
two sets of curves for the two temperatures. The correction
of the tunneling distance for the nonuniform field 共curvature
correction兲, is shown in Fig. 12, and appears to be an almost
universal function of applied voltage at 300 K, in spite of

TABLE VII. Doping gradient from SIMS fit.

dN/dx (1026 cm⫺4 )

Q N/P P/N–B P/N–BF2

5 0.74 0.28 0.43


6 0.82 0.36 0.54 FIG. 10. Conductivity 共a兲 and current density 共b兲 vs inverse maximum junc-
7 1.14 0.50 0.72 tion fields for all samples and all quadrants at 300 K. The fitting curve of
8 1.34 0.41 0.82 Taur et al. 共see Ref. 1兲, and the data of Hurkx et al. 共see Ref. 15兲 are shown
for comparison.

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al. 5809

FIG. 12. Curvature correction, ␸ / ␸ F , for nonuniform field, vs applied volt-


age. Data for quadrants Q2 共lower doped兲 and Q4 共highest doped兲, and for
all three wafer types NP, PN–B, and PN–BF2 , and at two temperatures, 150
and 300 K are superimposed.

it this way reveals that it is almost independent of the wafer


parameters, except for the anomalous dependence at higher
temperatures of Q6 of the NP wafer.

V. DISCUSSION
As we saw in Fig. 13, a remarkable unification of the
data has been achieved when plotting the current density

FIG. 11. Current density vs effective tunneling distance, and comparison


with the data of Hurkx et al. 共see Ref. 15兲. Data are from all quadrants of
the N/P wafers 共a兲, the P/N–B wafers 共b兲, and the P/N–BF2 wafers 共c兲.

one obtains ␭ T ⫽0.38 nm and A⫽2.0⫻107 and 5.3


⫻107 A/cm2 at 150 and 300 K, respectively. The value of
␭ T , from Eq. 共4兲, gives a tunneling effective mass of
0.29m 0 , assuming a band gap of ⬃1 eV.
D. Current density versus temperature
While a weak temperature dependence is evident in the
original data, this dependence will be distorted by the change
of the tunnel width with temperature, which decreases, at a
constant voltage, at low doping levels and increases at high
doping. To clarify this, the current was plotted as a function FIG. 13. Current density vs effective tunneling distance for the highest
doped quadrants 共Q4 –Q8兲 of all three wafer types and at 共a兲 300 and 共b兲 150
of temperature at a constant tunnel distance in Fig. 14 for
K. Each curve, for a particular quadrant and temperature, covers a voltage
Q5–Q8 of the three wafer types. The temperature depen- range as indicated in the figure, although the internal voltages for the most
dence is similar to that reported by others,5,7,14,15 yet plotting heavily doped quadrants are reduced by series resistance.

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
5810 J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al.

FIG. 14. Current density vs temperature, at a constant effective tunneling


distance, as indicated in the figure, for Q5–Q8 of all three wafers types.

versus effective tunnel distance. While from the theoretical


perspective this leaves some difficult issues to be explained,
from the empirical standpoint it is a boon for those wishing
to estimate tunneling current in silicon devices. The near
independence of the tunneling current on process and im-
plant conditions, bearing in mind that our processes were
engineered to produce high quality junctions, points to an
intrinsic process where process induced traps do not play a
major role in the tunneling process. It is worth noting, how-
ever, that a tunneling process via a midgap trap would give FIG. 15. Current density vs voltage at a constant effective tunneling dis-
tance, at 150 共a兲 and 300 K 共b兲, with each curve plotted across quadrants of
the same decay length as in the Keldysh model, where a the same wafer. Open symbols indicate quadrants 1– 4 共wafer NP only兲, and
phonon mediates the midgap transfer. Some of our higher filled symbols indicate quadrants 5– 8 of all wafers as indicated.
lying data at lower currents, as well as some of the refer-
enced data15 could be tunneling via traps. It should be noted
that our tunneling currents for different wafers, in the higher
complexities of the silicon band structure and phonon spec-
current range, agree closely with each other in spite of the
trum. The theoretical voltage dependence of the prefactor is
fact that the compensation ratios for the different wafers
dominated by the F maxVR term in Eq. 共12兲 for both direct and
were very different, ⬃30% for the highest doped quadrant of
indirect tunneling, which is clearly in contrast to our as-
the NP wafer, ⬃65% for the PN–B and ⬃45% for the
sumed lack of voltage dependence.
PN–BF2 wafers. The tunnel current densities do not seem to
The main theoretical question concerning our results is
be affected by this.
the weak voltage dependence of the tunneling prefactor. The
The effective mass, as derived in Sec. II A, of 0.29m 0 , is
voltage dependence of the current density, at a constant ef-
very reasonable considering the values of the light electron
fective tunneling distance, is shown in Fig. 15, where data
and hole masses in silicon.18 Given the complex band struc-
from different wafers quadrants are combined. None of the
ture of silicon, and our crude methods of analysis, this agree-
data show the superlinear behavior 共power law of 5/3 for a
ment is considered to be good, and awaits detailed numerical
graded junction at high voltages兲 predicted by Eq. 共12兲. For
modeling for further refinement. The magnitude of the inter-
most part the data are sublinear, approaching almost a con-
cept, for direct tunneling is approximated, under reverse bias
stant at 300 K for the PN–B and PN–BF2 wafers, while
conditions by
being closer to linear at 150 K 关This is seen also by compar-

J 0⫽
Ce 3 F maxV R
共2␲ប兲 2
冑 2m
eV b
, 共12兲
ing Figs. 13共a兲 and 13共b兲兴. This conclusion is somewhat de-
pendent on the value of the bandgap used in the extraction of
w TE , with the voltage dependence increasing for a higher
where F max is the peak junction field, an V R is the reverse assumed band gap, and is also dependent on the form of the
voltage, and C is a constant of the order of unity. The value curvature correction 共see Fig. 12兲, yet we believe this anoma-
of C and of the effective mass m depend on the model lous voltage dependence is significant, and needs to be veri-
details.7,19,18,21 Applying Eq. 共12兲 to our results, using the fied against more sophisticated models.
tunneling effective mass for m, gives an intercept which is As a check against more sophisticated theory, we com-
larger by 2–10⫻ than the experimental value. On the other pare our data with recent tunneling theory of Rivas et al.26
hand, assuming phonon-mediated, indirect processes19,20 based on the data of Thompson et al.8 While these data are in
gives an intercept which is too small, although an attempt to the forward bias regime, the very heavy doping leads to con-
obtain a simple numeric estimate was confounded by the siderable overlap of the bands making this situation appli-

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al. 5811

FIG. 17. Advanced 25 nm gate-length FET structure proposed by Taur et al.


FIG. 16. Silicon band gap, at 300 K, as a function of junction doping. For 共see Ref. 1兲, reproduced with permission, with minor changes. Drain voltage
Q5–Q8 the junction doping was extracted from the SIMS data, and for is 1.2 V. Solid contours are electrostatic potential, in 0.2 V intervals, and
Q1–Q4 the doping was extracted from the C – V using the exponential fit- dashed contours indicate doping, in 5⫻1018 cm⫺3 intervals with open con-
ting function. Reference data are from Van Overstraeten et al. 共see Ref. 29兲 tours being N type and closed contours P type. Region of shortest tunneling
and Swirhun et al. 共see Ref. 30兲. distance is shaded.

cable to our case, especially in light of the weak voltage tance, being primarily a property of the depletion region,
dependence of the current density at room temperature. Us- should, to first order, be unaffected by free-carrier band gap
ing their band diagram,26 which corresponds to the peak cur- lowering. One might argue that the potential in the contacts,
rent condition, we estimate a minimum tunneling distance of hence, the built-in voltage will be directly affected by free-
4.3 nm, which along with a curvature factor of ⬃0.9 leads to carrier band gap lowering in the contacts, but this is not so;
an estimate of the current density, based on our data, of 1500 the situation is analogous to a contact potential which is
A/cm2. This is in good agreement with the theoretical cancelled by the potential developed by the heterobarrier
estimate,26 even though their experimental value,9 was about formed between regions of higher and lower band gap
an order of magnitude higher. We suggest that the discrep- 共Second-order effects caused by band bending and by nonlo-
ancy lies in their doping profile,9 which is at the limits of the cal interactions31 may be important but are beyond the scope
SIMS technique. An article by the same group,27 published of our analysis.兲. Our data of Fig. 16 support this contention.
after submission of this work, presents a theoretical depen- Some of the lowest band gaps are found in higher-doped
dence of peak current–density versus tunnel distance with an quadrants the PN–B wafer, which has a greater compensa-
exponential decay length of 0.42 nm. This is very close to tion, and lower free-carrier concentration, than the other two
our 0.38 nm and could even be closer once the curvature wafers. Our data agrees well with that of Ref. 29 although
correction factor 共⬃0.9兲 factor is included. Now the surpris- there is a systematic shift of ⬃50 meV, with a slightly
ing thing about this is that the imaginary dispersion relations steeper trend-line, which is at the borderline of the uncer-
used by Rivas et al.27 for electrons and holes in the gap tainty in the data, and the assumptions of our analysis.
correspond to effective masses less than the 0.19m 0 and As an illustration of the engineering utility of our data
0.16m 0 band edge masses for electrons and holes, respec- and method of interpretation let’s examine the advanced FET
tively, and certainly less than the 0.29m 0 derived from our structure of Taur et al.1 and estimate the tunneling leakage
attenuation length. This discrepancy has been noted by current. The tunneling region, shaded in Fig. 17, is the region
Lake,28 and shows that the simple analytic formulas for the of shortest distance between contours separated by ⬃1 V 共the
exponential decay length do not hold up very well in the band gap兲 and is 5.4 nm in length by ⬃10 nm in width. A
complex situation presented by silicon. curvature factor of ⬃0.97 is inferred from Fig. 12 by noting
A critical aspect in our analysis was the extraction of that the total potential 共1.6 V兲 is 0.6 V greater than the band
doping profiles and tunneling distances from C – V data. It gap, thus the effective tunneling length becomes 5.25 nm.
was seen in Fig. 6, and in Tables III–VI that the C – V analy- The tunnel current density, from Fig. 13, is ⬃40 A/cm2, or
sis yields very credible results. Our results are also a source ⬃4 nA/␮m. While low, in device terms, this current in-
of band gap data, as a function of doping, for silicon. The creases exponentially with further scaling so will quickly
reason that this seemingly obvious method has not been used become important. For instance, if the whole geometry is
much to date is perhaps the sensitivity of the extracted band shrunk by 2⫻, tunneling region of 2.6⫻5 nm2, the current
gap to the modeling assumptions, and the difficulty in ob- will increase to 2.5 ␮A/␮m, which is prohibitively large for
taining highly accurate C – V data from heavily-doped P/N most applications.
junctions. In Fig. 16 the band gap is plotted against the dop- A proviso, concerning the earlier analysis is that our ex-
ing 共P or N兲 at the junction, and compared to data in the perimental results are for tunneling in the 具100典 direction
review paper of van Overstraeten and Mertens,29 and while devices are generally aligned along the 具110典 direction,
Swirhun et al.30 The band gap may be affected29 by free- in addition to the angle out of the plane 共see Fig. 17兲. This
carriers, or by the total dopant concentration, yet the capaci- might necessitate further experiments using different crystal

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions
5812 J. Appl. Phys., Vol. 95, No. 10, 15 May 2004 Solomon et al.

5
orientations. Alternatively, if our present results verify exist- A. G. Chyenoweth, W. L. Feldmann, C. A. Lee, R. A. Logan, and G. L.
ing, detailed, models, these models could be used with con- Pearson, Phys. Rev. 118, 425 共1960兲.
6
R. A. Logan and A. G. Chynoweth, Phys. Rev. 131, 89 共1963兲.
fidence to simulate future devices. 7
R. B. Fair, IEEE Trans. Electron Devices ED–23, 512 共1976兲.
8
P. E. Thompson et al., Appl. Phys. Lett. 75, 1308 共1999兲.
9
VI. CONCLUSIONS M. W. Dashiell, R. T. Troeger, S. L. Rommel, T. N. Adam, P. R. Berger, J.
Kolodzey, and A. C. Seabaugh, IEEE Trans. Electron Devices 47, 1707
Using technologically relevant ion-implanted profiles, 共2000兲.
we have shown that tunneling perpendicular to the 共100兲 10
R. Duschl, O. G. Schmidt, G. Reitemann, E. Kaspar, and K. Eberl, Elec-
surface, at high currents, is independent of the implant dose tron. Lett. 35, 1111 共1999兲.
11
D.-S. Wen, S. H. Goodwin-Johansson, and C. M. Osburn, IEEE Trans.
and type, for the conditions studied. The current density, in Electron Devices 35, 1107 共1988兲.
this regime, decreases exponentially with tunneling distance 12
B. G. Park, J. Bokor, C. S. Rafferty, and M. R. Pinto, IEEE Electron
with an attenuation length of 0.38 nm, corresponding to a Device Lett. 13, 507 共1992兲.
13
J. Koga and A. Toriumi, IEEE Electron Device Lett. 20, 529 共1999兲.
tunneling effective mass of 0.29m 0 , and a prefactor of 5.3 14
J. M. C. Stork and R. D. Isaac, IEEE Trans. Electron Devices 39, 331
⫻107 A/cm2 at 300 K and 2.0⫻107 A/cm2 at 150 K. The 共1992兲.
magnitude of the prefactor is consistent with a phonon- 15
G. A. M. Hurkx, D. B. M. Klaasen, and M. P. G. Knuvers, IEEE Trans.
assisted tunneling process, although its dependence of the Electron Devices ED–30, 1527 共1983兲.
16
E. Hackbarth and D. D. Tang, IEEE Trans. Electron Devices ED–35,
applied voltage is much weaker than expected. The doping 2109 共1988兲.
dependence of the band gap, extracted from the C – V data, 17
P. M. Solomon, D. J. Frank, J. Jopling, C. D’Emic, O. Dokumaci, P.
agrees with published data derived from transport in Si bi- Ronsheim and W. E. Haensch, IEEE Electron. Dev. Meeting, Washington,
polar transistors. Our data were shown to be useful in ana- D.C., Dec. 2003, Digest of Technical Papers, Session 9.3.
18
S. M. Sze, Physics of Semiconductor Devices 共Wiley, New York, 1967兲.
lyzing ultrascaled MOS transistors and should be useful in 19
L. V. Keldysh, Sov. Phys. JETP 6, 763 共1958兲; 7, 665 共1958兲.
quantifying the limits of silicon CMOS, especially when in- 20
P. J. Price and J. M. Radcliffe, IBM J. Res. Dev. 3, 364 共1959兲.
corporated into two-dimensional or three-dimensional device 21
E. O. Kane, J. Appl. Phys. 32, 83 共1961兲.
22
simulators. The need to extend this work to other crystal S. C. Jain, R. P. Mertens, P. van Mieghem, M. G. Mauk, M. Ghannam, G.
Borghs, and R. van Overstraeten, IEEE 1988 Circuits and Technology
orientations, was emphasized. Meeting, Paper 9.3, Digest p. 195.
23
B. R. Chawla and H. K. Gummel, IEEE Trans. Electron Devices ED–18,
ACKNOWLEDGMENTS 178 共1971兲.
24
E. M. Buturla, P. E. Cottrell, B. M. Grossman, and K. A. Salsburg, IBM J.
The authors wish first to acknowledge the support of the Res. Dev. 25, 218 共1981兲.
25
ASTL/S team under J. O. Dukovic, and the support of J. M. R. Kim, T. Aoki, T. Hirose, Y. Furuta, S. Hayashi, T. Shano, and K.
Taniguchi, IEEE Electron. Dev. Meeting, San Francisco, CA, Dec. 2000,
Warlaumont. They also wish to acknowledge the help of
Digest of Technical Papers, p. 523.
Philip Saunders for the ion implants, of Xinlin Wang with the 26
C. Rivas, R. Lake, G. Klimneck, W. R. Frensley, M. V. Fischetti, P. E.
FIELDAY code, and fruitful discussions with Max Fischetti, Thompson, S. L. Rommel, and P. R. Berger, Appl. Phys. Lett. 78, 814
Phil Oldiges, and Peter Price. They also thank R. Lake for a 共2001兲.
27
C. Rivas, R. Lake, W. R. Frensley, G. Klimeck, P. E. Thompson, K. D.
critical reading of our manuscript and for useful discussions. Hobart, S. L. Rommel, and P. R. Berger, J. Appl. Phys. 94, 5005 共2003兲.
28
R. Lake 共private communication兲.
1 29
Y. Taur, C. H. Wann and D. J. Frank, IEEE Electron. Dev. Meeting, San R. J. van Overstraeten and R. P. Mertens, Solid-State Electron. 30, 1077
Francisco, CA., Dec. 1998, Digest of Technical Papers, p. 789. 共1987兲.
2
G. G. Shahidi, IBM J. Res. Dev. 46, 121 共2002兲. 30
S. E. Swirhun, D. E. Kane, and R. M. Swanson, IEEE Electron. Dev.
3
P. M. Solomon, Proc. IEEE 70, 489 共1982兲. Meeting, San Francisco, CA, Dec. 1998, Digest of Technical Papers, p.
4
D. J. Frank, R. H. Dennard, E. Nook, P. M. Solomon, Y. Taur, and H.-S. 298.
Philip Wong, Proc. IEEE 89, 259 共2001兲. 31
M. V. Fischetti and S. E. Laux, J. Appl. Phys. 89, 1205 共2001兲.

Downloaded 25 Jul 2013 to 132.174.255.116. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jap.aip.org/about/rights_and_permissions

You might also like