Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

This article has been accepted for inclusion in a future issue of this journal.

Content is final as presented, with the exception of pagination.

IEEE TRANSACTIONS ON ELECTRON DEVICES 1

A Review of Hot Carrier Degradation in


n-Channel MOSFETs—Part I:
Physical Mechanism
Souvik Mahapatra , Fellow, IEEE, and Uma Sharma , Student Member, IEEE

Abstract — Transistor parametric drift due to conduction- was less than the dimensional scaling, resulting in a higher
mode hot carrier degradation (HCD) in n-MOSFETs is electric field in the device. Hence, the S/D junctions of the
reviewed, for long- and short-channel length (LCH ) devices device were changed from a conventional single junction to
having different source/drain (S/D) junction structures. The
HCD magnitude and time kinetics shape are discussed
lightly doped drain (LDD)-based double junctions to reduce
for stress under different gate (VG ) and drain (VD ) biases HCD with technology scaling [8]. However, due to aggressive
with varying VG /VD ratio, and without and with substrate junction scaling to mitigate short channel effects (SCEs) with
bias (VB ). Post-dc stress kinetics is discussed. The pub- further technology scaling, the doping of the LDD regions was
lished data are qualitatively analyzed to identify the roles of increased to minimize the S/D series resistance (RSD ). Hence,
different underlying physical mechanisms. In part II of this these devices have source–drain extension (SDE)-based S/D
article, impacts of technology scaling and stress tempera-
ture (T) and comparison of dc and ac stress are discussed. double junctions. The HCD became less of a concern with the
introduction of LDD devices (∼1990s). However, it gradually
Index Terms — Anode-hole injection (AHI), channel hot showed up in SDE devices (∼2000s) and is now an important
electron (CHE), channel initiated secondary electron
(CHISEL), electron and hole trappings, hot carriers, impact
issue again in high-performance FinFETs.
ionization, parametric drift, substrate and gate current, trap It is important to remark that many key HCD features have
generation. changed over the years with technology scaling.
1) The worst case HCD condition got shifted from
I. I NTRODUCTION VG ∼ VD /2 in long L CH [1]–[3] to VG ≥VD in short L CH
[4]–[6] devices (VG : gate bias, the definition of long and
T HE drift of transistor parameters, such as threshold
voltage (VT ), linear (IDLIN ) and saturation (IDSAT )
drain current, transconductance (G M ), and subthreshold
short L CH is shown in Section III).
2) The time kinetics shape got changed from a simple
power law in single, heavily doped drain junction
slope (S), due to hot carrier effect, first reported ∼40 years devices [2] to self-saturated nonpower-law in LDD
ago [1] remains as a well-known reliability concern for devices [3] and again back to power law with soft
long and short L CH bulk [2]–[4] and fully depleted silicon- saturation at long stress time in SDE devices [4]–[6]
on-insulator (FDSOI) [5] MOSFETs and FinFETs [6]. The (self-saturation and soft-saturation of time kinetics are
parametric drift is due to localized charge buildup near the defined in Section IV).
drain junction of the device at or near the channel and gate 3) The temperature (T) activation got shifted from negative
oxide interface. The localized charges result in the higher (HCD increases at lower T) in long L CH [7], [9] to
parametric drift when measured under reverse (source-drain positive (HCD increases at higher T) in short L CH [6],
flipped) compared with the forward (normal) mode [7]. Hot [7] devices.
carrier degradation (HCD) was reported in MOSFETs with 4) Higher HCD was reported for ac compared with
different source/drain (S/D) junction structures, for the reduc- dc stress in long L CH [10], and HCD got reduced
tion in channel length (L CH ) from ∼2 μm to ∼20 nm, for ac compared with dc stress in short L CH [6]
oxide thickness (TOX ) from ∼20 to ∼1 nm, and drain bias devices.
(VD ) from ∼10 to ∼1 V [1]–[7]. Note that the bias scaling 5) The reverse substrate bias (VB ) resulted in increased
Manuscript received April 19, 2020; revised May 4, 2020; accepted
HCD in bulk, long L CH devices [7], [11] although, for
May 6, 2020. The review of this article was arranged by Editor C. Monzio FDSOI, HCD reduces with reverse back-gate bias [12].
Compagnoni. (Corresponding author: Souvik Mahapatra.)
The authors are with the Department of Electrical Engineering, The short L CH devices handle lower voltages and are used in
IIT Bombay, Mumbai 400076, India (e-mail: souvik@ee.iitb.ac.in; modern logic cores, and long L CH devices are routinely used
sharma.uma2211@gmail.com). in many circuits (e.g., memory periphery and display drivers)
Color versions of one or more of the figures in this article are available
online at http://ieeexplore.ieee.org. and handle higher voltages (the classification of low and high
Digital Object Identifier 10.1109/TED.2020.2994302 stress VD is done in Section II). Unlike core devices for logic
0018-9383 © 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See https://www.ieee.org/publications/rights/index.html for more information.

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
2 IEEE TRANSACTIONS ON ELECTRON DEVICES

circuits, HCD has always remained an important concern in (increase in carrier T) as they drift in the channel from
high-voltage devices for analog applications [13]. source to drain under a favorable E LAT . Most of the electrons
This two-part article provides a comprehensive review and undergo various forms of scattering and lose energy by the
qualitative analysis of important experimental data for HCD time they reach the drain. However, a few of them escape
in n-channel MOSFETs and FinFETs published over the last scattering and continue to gain energy, and these hot electrons
∼40 years, i.e., the features listed earlier in 1)–5), with the initiate the following phenomena near the drain end of the
following targeted goals. First, this would serve as a reference MOSFET: 1) impact ionization and generation of electron–
for graduate students and practicing engineers especially those hole pairs and holes give rise to substrate current (ISUB ) [14],
entering the field. Next, this would guide the development while electrons flow to the drain; 2) gate current (IG ) due
and benchmark of technology CAD (TCAD)-based models for to the injection of holes at low VG (VG <VD /2) and electrons
predictive simulation of HCD, to help device design. Finally, at high VG (VG > VD /2) into the gate aided by favorable
this would also help in accessing the qualification and lifetime E VRT direction [15], [16]; 3) generation of new defects at
projection methodologies for different device technologies. or near the channel/gate insulator interface, which can be
In this article, experimental data are reviewed for various probed by the charge pumping (CP) technique1 [3]–[5], [11],
types of devices (single, LDD and SDE junction-based planar [17]–[23]; and 4) hole and electron trapping in the gate oxide,
and FDSOI MOSFETs and FinFETs), for the wide variation respectively, at low and high stress VG ’s [16]–[18], [24]–[27].
in device dimensions (L CH and TOX ) and stress conditions These generated defects and trapped charges are responsible
(VG , VD , and VB ), to develop a comprehensive (although for device parametric drift. Charges in the channel region
qualitative) picture of the HCD physical mechanism. Although impact VT and G M , in the overlap and LDD/SDE regions’
no new data are shown, some of the results are reinterpreted. impact series resistance (R S ) and G M (by R S ), and
The debates or gaps in the present understanding are also charges in all the regions impact drain current (ID ). However,
highlighted. In addition, the worst case HCD stress condition IDSAT is lower than IDLIN due to charge screening near the
is identified for various types of devices. Section II provides drain end at high sense VD .
an introduction to the HCD mechanism and discusses various Fig. 1(b) shows the energy band diagram (for VG –VD > 0 V)
processes that get triggered at different stress conditions. The along with the gate to substrate direction to highlight several
impact of varying the stress VG -to-VD ratio, in long L CH energy values of interest: Si bandgap of ∼1.1 eV for impact
devices under high stress VD and short L CH devices under ionization and conduction (valence) band offset between the
low stress VD , is discussed in Section III. The impact of channel and gate oxide for electron (hole) injection of ∼3.1 eV
stress VD variation is discussed in Section IV. The impact of (∼4.7 eV)2 [16]. It is important to remark that although ISUB
anode-hole injection (AHI) during stress under high VG /VD is traditionally used as a measure of hot electrons [2], [14],
ratio and reverse VB is discussed in Section V, and its an electron resulting in impact ionization (ISUB /ID ∼ 10−3 ) is
effect on the post-dc stress kinetics is also discussed. This not the one responsible for the creation of new defect as the
article is concluded, and modeling guidelines are provided in original electron and generated electron–hole pair are of low
Section VI. energy after the impact ionization event. Few hot electrons,
In part II of this article, experimental data at worst case after redirectional collision toward the gate oxide/channel
stress conditions for different devices (L CH /TOX , and S/D interface, can either get injected into the gate and result in IG
junction) are reviewed to understand the impact of technology at high VG (IG /ID ∼ 10−9 ) or get trapped in the preexisting
scaling. The differences in the time kinetics shape of different gate insulator defects, or can collide with the defect precursors
S/D junction devices are discussed. The impact of L CH and and generate defects. Holes generated out of impact ionization
TOX scaling is analyzed. The impact of T variation in long and move toward the source due to favorable E LAT and gain energy
short L CH devices is explained. The dc and ac stress kinetics and can either result in IG at low VG (IG / ID ∼ 10−11 ), or get
are compared under high and low VD stress conditions. trapped, or generate new defects. Therefore, the electrons (or
holes) that constitute IG do not result in defect generation
and/or trapping, and vice versa. Note that the ratios shown
II. D IFFERENT H OT C ARRIER P ROCESSES
earlier are typical peak values to indicate the probability of
Fig. 1(a) shows the schematic of an LDD MOSFET. For different processes, and these would change across devices
conduction-mode HCD stress (VG > VT0 , VT0 is the prestress and stress conditions. The electrons and holes responsible for
threshold voltage, and VD > 0 V), the applied drain bias drops HCD are denoted as “lucky carriers” [28].
across the channel (primarily) and LDD (depends on doping)
regions. The lateral electric field (E LAT ) increases from the
source toward the drain, peaks at the channel/LDD junction, 1 In CP technique, the gate of the MOSFET is repetitively pulsed between
and reduces toward the LDD/drain junction; the direction of accumulation and inversion, the source and drain are tied together to ground
E LAT is from drain to source. The vertical electric field (E VRT ) (or slightly reverse biased), and the substrate current is measured, which is
coming from trap (∼defect) assisted electron–hole recombination at or near
peaks at the source end of the channel and reduces toward the the channel/gate insulator interface. Increase in CP current after HCD stress
drain. The direction of E VRT is from gate to channel although is due to generation of new traps. Besides channel region, CP can be used to
in some cases (VG <VD ), the direction can flip especially near probe defect generation inside the drain junction, especially in LDD devices,
under the application of large gate pulse [3].
the drain end of the channel. 2 In these papers, high V and low V regimes are denoted, respectively,
D D
The conventional channel hot electron (CHE) process is by whether VD is above or below the voltage (∼3 V) corresponding to the
qualitatively explained using Fig. 1(a). Electrons gain energy electron injection barrier.

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
MAHAPATRA AND SHARMA: REVIEW OF HCD IN n-CHANNEL MOSFETs—I 3

Although computationally heavy, the most accurate method


to calculate carrier heating is to solve the Boltzmann transport
equation using the kinetic Monte Carlo (KMC) method [29].
Fig. 1(c) shows the plots of the distribution of hot electrons
(E > 3.1 eV) and holes (E > 4.7 eV) along the channel near
the drain from KMC simulations in an LDD MOSFET [11].
The energy cutoffs are for carrier injection over the electron
and hole energy barriers; however, the energy threshold for
defect creation can be much lower (see Section IV-B). The
distribution of hot holes is a bit shifted from the drain
junction toward the channel center than that of hot electrons.
Fig. 1(d) shows the plots of the peak hot electron and hot hole
density (HHD) as a function of VD . Although both reduce at
lower VD , the HHD reduces much faster than hot electrons as
the energy barrier is higher for holes than electrons. Therefore,
the injection of hot holes and electrons would reduce with
device scaling due to the coreduction in stress VD . Note that
the abovementioned results are for illustrative purposes and
can change across devices and stress conditions.
Under the application of reverse VB in bulk MOSFETs, Fig. 1. (a) Schematic of an LDD MOSFET. Carrier heating process,
the holes generated out of the primary impact ionization primary and secondary impact ionization, respectively, at VB = 0 V and
undergo secondary impact ionization as they flow toward VB <0 V, and gate injection are shown. Charges due to HCD in the chan-
nel (square), gate–drain overlap (triangle), and spacer (diamond) regions
the substrate and generate additional electron–hole pairs, are shown. (b) Energy band diagram showing the energy thresholds for
as shown in Fig. 1(a) [29]. The resulting holes increase impact ionization, and electron and hole injection over their respective
ISUB (ISUB /ID ∼ 10−1 ), and the secondary electrons have channel-oxide barriers. AHI process (refer to Section V) is illustrated at
(VG –VD )> 0 V. KMC simulated (c) hot electron density (HED) and HHD
higher energy for injection into the gate and increase IG distributions in the channel and (d) peak HED and HHD as a function of
(IG /ID ∼ 10−7 ) compared with the CHE process; the ratios VD (normalized) (see [11]). HHD is negligible at VG = VD in (d).
denote typical peak probabilities. This is known as the chan-
nel initiated secondary electron or “CHISEL” process. More
energetic carriers are generated, and the hot electron/hole
distributions are also broader due to the secondary ionization
process (not explicitly shown) [11].
AHI process is routinely invoked in the time-dependent
dielectric breakdown (TDDB) studies [30], [31]. Under HCD
stress, AHI can get triggered at a higher VG /VD ratio [see
Fig. 1(b)]. In the portion of the channel where the direction
of E VRT is favorable, electrons would get injected from the
channel to the gate and initiate impact ionization to gener-
ate electron–hole pairs in the gate, and the generated holes
are injected back into the oxide and generate defects. The Fig. 2. Measured VG dependence of (a) ISUB and IG in single-junction
MOSFET at high VD [32] and (b) ISUB in different LCH FinFETs at low
favorable position is near the drain junction if VG /VD ratio VD [33].
becomes high, and this would aid the injection of hot electrons
into the gate to trigger the AHI-initiated trap generation3 [11].
AHI also becomes higher for VB <0 V stress, which results [32]. Note that ISUB peaks at VG ∼ VD /2; similar data are
in more energetic carriers due to the secondary ionization shown in other reports [2], [14]–[16], [28]. The bell-shaped
process [11], [29]. characteristics are due to two competing processes: reduction
The abovementioned processes are invoked to qualitatively in E LAT (∼carrier heating) and an increase in the drain current
explain the experimental results presented in this article. All (ID ) at higher VG . It is noteworthy that the ISUB /ID ratio,
data shown are at room temperature (RT) unless specified an indicator of carrier heating, decreases monotonically as
otherwise. VG is increased (fixed VD ) for classical HCD (not explicitly
shown).4 IG is due to hot holes at lower VG and hot electrons at
III. I MPACT OF S TRESS G ATE B IAS ON CHE P ROCESS higher VG since the ratio of VG /VD determines the magnitude
A. Substrate and Gate Currents and direction of E VRT near the drain junction of the device
where hot holes and electrons exist. E VRT direction toward the
Fig. 2(a) shows the plots of the measured ISUB and IG as a
gate is favorable for hole injection at VG <VD /2 and toward the
function of VG in a long L CH device under high stress VD
substrate favors electron injection at VG > VD /2 (refer to [16]
3 Pure AHI due to tunneling of channel electrons to gate can happen near
the source end of the channel, provided that TOX is thin enough for tunneling 4 Devices featuring this bell-shaped I
SUB versus VG curve are classified as
and VG is high enough for the impact ionization event. long L CH . All features reported in such devices are defined as classical HCD.

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
4 IEEE TRANSACTIONS ON ELECTRON DEVICES

Note that the maximum G M and VT at VG ∼ VD /2 con-


dition is primarily due to generated defects (∼ ICP ) as both
holes and electrons are injected into the gate resulting in
smaller overall trapping. This coincides with the peak of the
ISUB versus VG curve [see Fig. 2(a)], and many have used this
condition to study HCD in long L CH devices under higher
stress VD . However, although the peaks of ICP and VT are
at VG ∼ VD /2 [see Fig. 3(a)], the peak is somewhat broader
for ICP with larger spread toward lower VG compared with
Fig. 3. Measured fixed time (a) ΔVT and ΔICP (normalized to the that of VT ; similar results are shown in [19] and [22]. The
respective peak values) at 1.8 ks [18] and (b) ΔGM and ΔICP [26] versus broader ICP peak toward lower VG indicates that hot holes
stress VG . (b) Data (ΔGM ) are shown for basic HCD at 500 s (ΔGM ) and are efficient in defect generation. Hole injection and trapping
also after an electron injection phase of 10 s (ΔGM and ΔICP ). Single-
junction long LCH MOSFETs. at VG <VD /2 stress reduce VT compared with ICP due to
compensation and can explain the narrower peak or shaper
reduction of VT than ICP as VG is reduced.
for details). IG due to hot holes peaks at VG ∼ VD /4 and Indeed, many have suggested the injection of hot holes
reduces at lower VG as ID is small (fewer electrons to initiate into the gate oxide as a precondition for defect genera-
impact ionization). IG is small at VG ∼ VD /2 because of the tion [16], [18], [26]. It has been suggested that the energy
simultaneous injection of electrons and holes. IG due to hot released due to the recombination of trapped holes (following
electrons peaks at VG ∼ VD , and reduces at higher VG as carrier hot hole injection) and electrons (also injected) triggers defect
heating reduces. generation in these devices, having long L CH and stressed
Fig. 2(b) shows the plots of the measured ISUB versus VG under high VD [35]. However, it observed that the HCD defects
in different L CH devices at fixed (low) VD [33]. ISUB is higher are chemically not the same (and/or are not formed by the
at lower L CH at fixed VG and VD as expected due to higher same mechanism) as those generated by uniformly injected
hot carrier density. However, the ISUB versus VG dependence carriers into the gate oxide for VG > 0 V/VD = 0 V stress6 [36].
gets shifted from the conventional bell shape with a peak at Therefore, “injection inside the oxide bulk” is not a basic
VG ∼ VD /2 (in long L CH ) to monotonically increasing at higher requirement; energetic holes during HCD stress can collide
VG as L CH is reduced.5 It is noteworthy that the ISUB /ID with the defect precursors at or near the channel/gate oxide
ratio remains almost constant as VG is increased in short interface to generate defects.
L CH devices (not explicitly shown) [33], which is unlike the However, if present, “hole injection inside the oxide bulk”
classical behavior seen in long L CH devices. The same is also can generate additional defects. As proof, Fig. 3(b) shows
reported in [34]. Note that IG due to the hot hole and electron the plots of the measured G M versus stress VG following
injection would reduce in short L CH devices as VD is scaled. a short electron injection phase at VG = VD condition after
IG due to the quantum mechanical tunneling can be large due the basic HCD stress [26]. Note that G M increases after the
to thinner TOX (<3 nm) in these devices [7]. short electron injection phase, and the peak G M location is
shifted toward lower VG . ICP also increases by ∼20% after
the short poststress electron injection phase (data shown only
B. Classical HCD Behavior (High VD ) after the final short electron injection phase, but not right after
basic HCD). However, the electron injection phase itself, i.e.,
Fig. 3 shows the plots of the measured ICP and VT [18]
when done alone, does not generate defects in any measurable
[see Fig. 3(a)] and ICP and G M [26] versus stress VG [see
quantity. Remarkably, the difference in G M between the
Fig. 3(b)], where ICP is the increase in the CP current due
original HCD and post electron injection is larger when the
to stress-generated defects. The HCD peaks at VG ∼ VD /2 and
original stress is done at lower VG .
reduces for VG <VD /2 and VG > VD /2 conditions. Note that
It has been suggested that neutral electron traps (defects) are
G M and VT (poststress minus initial) are negative and
created in the oxide bulk by injected holes at VG <VD /2 stress
positive, respectively, and indicate mobility reduction and
(more on this in Section V) [26], and these traps get activated
negative oxide charge buildup as VG is varied, except at
during the brief electron injection phase. It is also possible that
really low VG (in Fig. 3(a) and in [16] and [24]) when the
additional defects are created during the poststress electron
exact opposite is reported. The generation of acceptor-like
injection phase since a ∼20% increase in ICP is present due
defects (negatively charged only if occupied by electrons,
to the trapped hole and electron recombination process [35].
neutral otherwise) and electron trapping are responsible for the
This explains a larger G M increase after electron injection
reduction in mobility and negative gate oxide charge buildup.
when the original stress is done at lower VG . In addition,
However, at very low VG , oxide trapped holes overcompensate
neutralization of trapped holes by the injected electrons would
for the effect of acceptor-like generated defects; GM increases
make G M primarily due to generated acceptor-like defects
due to channel shortening (not seen in Fig. 3(b) but shown
in [16] and [24]), and negative VT shift is observed. 6 The impact of deuterium (D )-based postmetallization anneal (PMA) is
2
observed for HCD stress but not much for uniform stress under VD = 0 V.
5 Devices featuring this monotonically increasing I
SUB versus VG curve are It is now well established that defect generation is lower under HCD stress in
classified as short L CH . All features reported in such devices are defined as D2 PMA compared with the conventional hydrogen (H2 )-based PMA devices
nonclassical HCD. (refer to [36] and references therein).

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
MAHAPATRA AND SHARMA: REVIEW OF HCD IN n-CHANNEL MOSFETs—I 5

TABLE I
P OWER -L AW S LOPE F ROM D IFFERENT R EPORTS (H IGH VD S TRESS )

Fig. 4. Measured time kinetics of (a) ΔGM in LDD MOSFET [37] and
(b) ΔVT in single-junction MOSFET [27] at different VG /VD ratios.

and can also explain the observation. Therefore, the net


increase in negative charges results in higher G M after the
short electron injection phase. This has important implications
for ac stress at high VD , which is discussed in part II of this
article.
Note that ICP , G M , and VT reduce similarly when VG
is increased beyond the VG ∼ VD /2 condition, which indicates
the inefficiency of defect creation by hot electron injection
(and/or collision) alone (E VRT is less conducive for hot hole
injection at VG > VD /2). As shown in [19], hot electron injec-
tion is less efficient than hot hole injection (either alone,
or recombination of trapped holes by injected electrons) in
generating defects at or near the channel/gate oxide interface.
Moreover, negligible electron trapping is observed in these
devices due to a similar reduction in ICP , G M , and VT
at VG > VD /2 stress condition. In the presence of electron
trapping, G M (and/or VT ) would reduce less compared
with ICP at VG > VD /2 stress (refer to [17] for additional
details).
Fig. 4 shows the plots of the measured time kinetics of
G M [see Fig. 4(a)] [37] and VT [see Fig. 4(b)] [27]
esis (the slopes are identical for VG = VD /2 stress). Hence,
for stress under varying VG /VD ratio. Note that G M
when stress VG is increased beyond VD /2, the reduction in
reduces (negative) and VT increases (positive) monotonically
NIT from CHE or pure HCD [however, the AHI contribution
for VG ∼ VD /2 and VG > VD /2 conditions and shows power-law
can increase in some cases (see Section V)] and increase in
time dependence with slope n. This is observed across various
electron trapping (depends on the oxide quality and thickness)
reports [2], [3], [9], [11], [17], [20], [22], [27], [28], [36]–[41],
can explain the time kinetics of parametric shift.
and the time slopes are listed in Table I, together with device
However, G M increases first and then reduces with time at
and experimental details. Note that HCD is primarily due to
VG <VD /2 stress condition (switches from positive to negative).
defect generation at VG ∼ VD /2 stress, as the same n is reported
This is because of the dominance of hole trapping at an
for NIT , the generated interface trap density measured by CP,
early time and acceptor-like generated defects at a longer time
and G M [17]. However, the slope varies between different
(as trapping saturates), visible, respectively, before and after
published reports, which is further discussed in Section IV-B.
the “dip” in the log–log plot. The long-time slope n (after
The slope reduces at VG > VD /2 stress compared from the
“dip”) appears larger due to the early hole trapping phase.
same report (although this is not always the case, e.g., [37],
Finally, the generation of neutral oxide traps, if any, has no
[41]), see Table I. Such reduction is presumably due to
impact, unless trap filling is done to expose those, as shown
AHI-initiated defect generation (impacts NIT , VT , and
in [26]. Therefore, the relative and opposing contribution of
G M ) and electron trapping (impacts VT and G M ). The
hole trapping and trap generation would explain the time
AHI process is explained in Section V. Note that electron
kinetics of parametric shift as stress VG is reduced below VD /2.
trapping is negligible at VG ∼ VD /2 but can show up under
VG > VD /2. By invoking fast saturation of the electron trapping
time kinetics (n∼0 in a log–log plot at longer stress time [42]), C. Scaling and Nonclassical HCD Behavior (Low VD )
the reduction in time slope can be explained (addition of Fig. 5 shows the plots of the measured IDLIN versus VG
power-law time kinetics (n > 0) and saturated time kinetics for the variation in L CH (at fixed TOX ) [see Fig. 5(a)] [12] and
(n = 0) would reduce the slope of the resultant time kinetics). TOX (at fixed L CH ) [see Fig. 5(b)] [43]; for a similar type of
The measured n for G M is lower compared with NIT at devices, the stress VD is low and kept fixed for both plots.
VG = VD for the same device [17], which proves this hypoth- For a fixed VG , IDLIN increases at short L CH but reduces

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
6 IEEE TRANSACTIONS ON ELECTRON DEVICES

Fig. 5. Measured fixed time ΔIDLIN from devices having variation in Fig. 6. Measured ΔIDLIN time kinetics at different stress VD in
(a) LCH (at fixed TOX ) [12] and (b) TOX (at fixed LCH ) [43] versus stress VG . (a) relatively long LCH SDE MOSFETs at VG ∼ VD /2, VG ∼ VD and only
All data are from FDSOI MOSFETs. The value of stress time is not
VG (VD = 0 V) stress [46] and (b) relatively short LCH FDSOI MOSFETs
available. at VG ∼ VD /2 and VG = VD stress [12].

at thinner TOX ; the scaling effects are discussed in detail in TABLE II


P OWER -L AW S LOPE F ROM D IFFERENT R EPORTS (L OW VD S TRESS )
part II of this article. However, of interest is the shape of
the IDLIN versus VG plot, which gradually shifts from bell-
shaped to monotonically increasing dependence, as either L CH
is reduced or TOX is increased.
Note that the original explanation of increased HCD at
VG = VD is defect generation by multi (two), hot (moder-
ately) electron interaction process [12], [43]–[45]. An alter-
native explanation is as follows. For a well-controlled, “long-
channel” device, the pinch-off region and E LAT reduce sharply
at higher VG (fixed VD ), which reduces carrier heating. This
is evident from the reduction in ISUB with VG in long L CH
devices [see Fig. 2(b)]. The same does not hold in a device
encountering relatively larger SCE (short L CH at fixed TOX or
thicker TOX at fixed L CH ), as evident from the monotonically
increasing ISUB versus VG plot in short L CH devices [see
Fig. 2(b)]. Hence, the carriers are still hot at high VG (as
E LAT is high and pinch-off region exist) and are also high
in density (∼higher ID ); hence, the CHE process does not
reduce like the long channel case and is presumably the reason
for the nonclassical VG dependence of HCD. It is remarkable
that hot holes are present in devices showing high ISUB at
VG = VD stress. Since hot holes are more efficient than
electrons in generating defects, as discussed in Section III-B,
it is possible that enhanced HCD at higher VG is linked to hot
holes. However, it is difficult to be fully certain if this is the
case, or it is just the hot electrons that generate defects in these
conditions [7]. Proper quantification is needed to resolve such this also suggests a reduction of the AHI-initiated process (see
differences, which is beyond the scope of this review article. Section V) at VG = VD condition near the drain due to low VD .
Fig. 6 plots the measured IDLIN time kinetics in relatively However, the gate bias-assisted AHI process can still happen,
long [see Fig. 6(a)] [46] and short [see Fig. 6(b)] [12] L CH especially near the source side of the channel, although note
devices for VG ∼ VD /2 and VG = VD conditions at low VD that the uniform stress (VD = 0 V) shows much lower IDLIN
(<3 V) stress (note that the L CH in Fig. 6(a) is much shorter compared with HCD stress (VD > 0 V) [see Fig. 6(a)]; refer to
than those used in Fig. 4). Note that power-law time kinetics the next section for a related discussion of Fig. 6(b). The slope
is observed [with soft-saturation at longer stress time, see n from various reports shows a widespread even for low VD
Fig. 6(a)] but with very different slope n of ∼0.5 (the early stress (n ∼ 0.2−0.5) [4]–[7], [12], [44], [46]–[51]. The slopes
part before saturation) in Fig. 6(a) and ∼0.18 in Fig. 6(b). are listed in Table II and discussed further in Section IV-B.
However, the slope does not vary with variation in VG /VD ratio.
The invariance of n as VG is varied is unlike that observed for
high VD stress (see Fig. 4). This is due to the negligible injec- D. Effect of Positive Bias Temperature Instability (PBTI)
tion of hot electrons at VG = VD condition, since VD is low, Note that PBTI became an important concern espe-
and negligible electron trapping (even though the tunneling is cially for the high-K metal gate (HKMG) gate insulator
large) due to thinner TOX and low trapping volume. Moreover, process [52], [53]. It depends on E VRT and degrades the

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
MAHAPATRA AND SHARMA: REVIEW OF HCD IN n-CHANNEL MOSFETs—I 7

Fig. 7. Correlation of ΔIDLIN and ΔVT for IDLIN sensed in (a) subthreshold
and (b) above the threshold for HCD stress at different VD /VG ratios; data
from PBTI stress are shown for comparison [56]. SDE MOSFET, TOX , not
available.

channel region (not overlap), and the defects are nonuniform Fig. 8. Measured time kinetics of (a), (c), and (d) ΔVT and (b) ΔIDLIN
(reduce toward the drain) for VD > 0 V stress. It affects at different stress VD for (a) and (b) VG ∼ VD /2 and (c) and (d) VG = VD
stress. Data are scaled along the time axis to universal relation (solid
VT and IDLIN (not G M and R S ). It can influence symbols). See text for device details. Data in (b) are at RT. Time-axis
the HCD parametric drift, especially for VG = VD stress, scaled data are shown as crossed symbols.
used in short L CH devices under lower VD stress [54], [55].
On the contrary, HCD affects VT , G M , R S , and IDLIN .
Therefore, a correlation of IDLIN to VT would help estimate IDLIN versus VT correlation is largest in this case. Note
the contribution of PBTI during HCD stress. that PBTI shows up at VG = VD and more so at VG > VD
Fig. 7 shows the plots of the measured IDLIN versus condition, making the overall (HCD+PBTI) degradation more
VT for stressing at different VG /VD conditions in a short uniform in the channel, which reduces the IDLIN versus VT
L CH device at low stress VD when IDLIN is sensed in correlation toward pure PBTI case. Since the time slope n is
subthreshold [see Fig. 7(a)] and above threshold regimes [see very similar between PBTI (n ∼ 0.17) [52], [53] and pure
Fig. 7(b)] [56]. Note that this device uses HKMG-based gate- HCD [n ∼0.18 at VG ∼ VD /2, see Fig. 6(b)] in this case, n
stack where PBTI is present. For a given VT , very large % does not change under VG = VD stress, as shown in Fig. 6(b).
IDLIN is seen for subthreshold sense as the prestress IDLIN is
small. IDLIN and VT are perfectly correlated for HCD (all IV. I MPACT OF S TRESS D RAIN B IAS ON CHE P ROCESS
VD /VG ratios) and PBTI (VD = 0 V) stress. However, IDLIN A. Time Kinetics and Universal Scaling
and VT are not correlated with HCD and PBTI when sensed
Fig. 8 shows the plots of the time kinetics of measured VT
above the threshold. IDLIN at a given VT reduces at a lower
in a long L CH planar MOSFET with a single S/D junction
VD /VG ratio, with a very similar correlation between the HCD
[see Fig. 8(a)] [9], IDLIN in a long L CH LDD MOSFET
at VD <VG and PBTI (VD = 0 V) stress.
[see Fig. 8(b)] [3], VT in a short L CH SDE MOSFET [see
The original interpretation is the HCD results in more
Fig. 8(c)] [51], and VT in a short L CH FinFET [6] as a
uniform channel degradation for multicarrier-based defect
function of VD [see Fig. 8(d)]. The stress VD is high and
generation processes at VG ≥VD compared with the single
VG ∼ VD /2 in Fig. 8(a) and (b), and the stress VD is low and
carrier process at VG <VD /2, with a larger spread for the many
VG = VD in Fig. 8(c) and (d). Power-law time kinetics is seen
carrier processes at VG > VD than the two-carrier process at
in Fig. 8(a), (c), and (d), while a self-saturating nonpower-
VG = VD [12], [43], [44]. PBTI (VD = 0 V) results in fully
law time dependence is seen in Fig. 8(b). The difference in
uniform channel degradation. Hence, the IDLIN versus VT
the time-dependent shape between the LDD device versus
correlation for PBTI and HCD (at VG > VD ) stress gets closer
the others7 is discussed in detail in part II of this article.
to each other.
Note that a comparison of HCD magnitude across devices
An alternative explanation is as follows. The subthreshold
in Fig. 8 is challenging due to the very different nature of
sense is not affected by R S , both IDLIN and VT are
gate oxide quality (more on this in Section IV-B) and data at
due to defect generation in the channel, and they perfectly
different T (more on this in part II of this article). However,
correlate for HCD stress at various VG /VD conditions and
VD dependence can be studied.
PBTI. IDLIN is affected by R S when sensed above the
Conventionally, the fixed-time degradation is plotted versus
threshold. IDLIN versus VT varies for PBTI and HCD
1/VD on a semilog scale (exponential dependence) [2], [22].
stress due to the difference in the relative degradation of the
This is done as the same holds for the VD dependence of
channel and overlap/SDE regions. The relative contribution
of channel degradation is larger than that of the overlap/SDE 7 In Fig. 6(a), the time kinetics shows power law at shorter time, and a

region for PBTI (VD = 0 V) stress, and IDLIN versus VT soft-saturation is seen at longer time only when the degradation becomes
correlation is lowest. PBTI is negligible at very low VG , and high and reaches the maximum. However, in Fig. 8(b), the rate of degra-
dation reduces continuously (self-saturation) at the beginning of stress and
VG ∼ VD /2 stress is due to HCD localized near the drain. The asymptotically settles to a final value even if the degradation is far from the
relative contribution from channel degradation is lowest, and maximum limit.

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
8 IEEE TRANSACTIONS ON ELECTRON DEVICES

energy (E C ) of ∼1.5 eV often assigned to the dissociation


of hydrogen (H)-passivated Pb centers [7], [60], [61]. One
possibility can be the carrier–carrier scattering induced carrier
energy enhancement [29]. It is also possible that the bond
strength is lower in some devices [6]. It is remarkable that
the HKMG devices have a chemical oxide-based interlayer
(between the channel and high-K layers), which is a low T
process [62]. It is reported that the oxide growth T (∼oxide
quality) can impact the bond strength and hence the magnitude
of HCD [39]. The self-heating (SH) effect [63] in FinFETs
Fig. 9. Correlation of 1/S versus VD for (a) single-junction MOSFETs (and FDSOI) can further activate HCD (see part II of this
having abrupt and graded junctions and different LCH [39] and (b) data article for T activation of HCD). More work is needed to
obtained from Fig. 8 (see text for details). The data sets are arbitrarily understand these issues.
shifted along the Y-axis for clarity, so only the slope (∼VAF) should be
considered. Except for LDD devices [see Fig. 8(b)], other devices
show power-law time dependence. In general, power-law time
kinetics can be explained either by a reaction-limited process
ISUB (∼carrier heating) [2] due to the exponential dependence (e.g., spread in the energy required to break a defect) [64]
of ISUB on 1/E LAT (see the Lucky-Electron model in [28]). or a diffusion-limited process [65]. More analysis is required
The voltage acceleration factor (VAF), i.e., the slope of the to identify the correct physical process. Moreover, the effect
HCD versus 1/VD plot, depends on L CH /TOX [2] and junction of generated charges on the potential and electric field in the
profile [38], as these quantities impact E LAT . channel needs to be considered, which can reduce the carrier
Note that the VD dependence of degradation can be found heating (∼ISUB ) with the passage of time, to estimate HCD
at a fixed time in Fig. 8(a), (c), and (d) as the time kinetics time kinetics.
shows power law with similar n across VD . This is not feasible Another intriguing factor is the wide range of n reported
in Fig. 8(b), as the nonpower law dependence would make in Tables I and II. This can be partially explained as follows.
the fixed time VD values dependent on the time when it is The HCD generated Pb centers do not recover [33] unless a
recorded. Therefore, the time kinetics at a different VD is very high T bake is used after stress [61]. The parametric
scaled by a factor S along time (X-axis) to obtain a unique drift can recover after stress, more likely for thicker TOX
behavior as shown; S is larger for higher VD , and vice versa. devices when electron trapping is present during stress [66].
The time axis scaling is explained in detail in [3], [4], [13], Note that electron trapping can be small but present even for
and [57]. Moreover, it is realized afterward that HCD is an VG ∼ VD /2 stress, as the electron and hole spatial profiles are
energy-driven process [45], [58]. Since carrier energy is ∼VD , physically separate [see Fig. 1(c)], while it can certainly be
1/S, which is a measure of a lifetime, is plotted versus VD in present for VG > VD /2 stress.
a log–log plot (power-law VD dependence) [57]. On the other hand, during CP measurements, any bulk gate
Fig. 9 shows the plots of the VD dependence of 1/S for oxide traps (generated during HCD stress if VD is high) would
abrupt and graded single-junction MOSFETs with fixed TOX get filled by holes and electrons, respectively, during the pulse
but a different L CH [see Fig. 9(a)] [38] and for different low/high phases. Note that the energy bandgap scanned by CP
devices [see Fig. 9(b)] corresponding to Fig. 8. The power-law depends on the local VFB and VT values8 (in accumulation
VAF is higher (∼38) for the abrupt compared with the graded and inversion for pulse low/high phases), and these would
(∼30) single-junction devices from the same report [38], but change (VFB becomes more negative and VT more positive
it does not vary with L CH for a given type of junction. due to hole and electron trapping, respectively) during CP
Moreover, similar VAF (∼24–30) is seen for the other devices measurements. The energy bandgap scanned by CP would
in Fig. 8 [3], [6], [9], [51] and also for other devices listed in reduce and results in a reduction in CP current (∼recovery)
Tables I and II [4], [20], [40], [46], [50] (not explicitly shown, when the measurement is done after stress.
see [57] for additional cases), covering wide variation in L CH , It is now well-known that recovery increases the power-law
TOX , and stress VD . slope n when measured using the slow method [67]. The
The time-axis scaling at a different stress VD (with VG slopes listed in Tables I and II (except in [6], [49], and [50])
chosen to represent the worst case or more realistic use are measured using slow methods, and so the actual value
condition) can be used for lifetime determination under HCD would be a bit lower than shown, especially for the thicker
stress. This can be used for all types of devices to remain TOX devices. This, however, is unlikely the case for thinner
consistent across all technologies, as the method can account TOX devices since the electron trapping is negligible. Hence,
for any soft- or self-saturation effect. if “properly” measured, we speculate that the reported n shown
in Table I would be ∼0.5 at VG ∼ VD /2 stress for most cases.
Note that n ∼ 0.5 is reported for most cases also in Table II.
B. Discussion on Defects and Time Kinetics However, the spread in n toward lower (than n ∼ 0.5) values
Note that Pb centers (Si3 –Si–) [59] are often associated
8 In the CP method, V
with defects generated during HCD (CHE stress) [7]. It is FB and VT are defined as the gate voltage needed
to achieve critical hole and electron density at the channel surface so that
indeed intriguing that appreciable HCD is observed for stress complete trap filling happens during the duration of pulse low/high phase,
at low VD of ∼1 V [6], as this is much below the critical respectively [19].

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
MAHAPATRA AND SHARMA: REVIEW OF HCD IN n-CHANNEL MOSFETs—I 9

Fig. 10. Impact of stress VG on measured (a) fixed time ΔIDLIN in a


different LCH FDSOI MOSFETs [43] and (b) 1/lifetime (from 10% reverse
ΔIDSAT ) at a different VD in LDD MOSFET [45]. Fig. 11. Measured ΔICP time kinetics (a) during and (b) after stress at
different VG /VD ratio in LDD MOSFET having a different TOX [11].

in Tables I and II is hard to explain and needs more work.


The nonpower-law time dependence, often reported in LDD
devices [3], [20], [40], is discussed in part II of this article.
Interestingly, similar n is seen from different probes in single-
junction [28] and SDE [4] devices, while this is not the case for
some LDD devices (the ones showing power-law time kinetics)
[37], [41]. This aspect is also discussed in part II of this article.
Finally, as mentioned earlier, n reduces for VG > VD /2 condi-
tion because of electron trapping and AHI. The impact of the
AHI-initiated process is discussed in the following.

V. I MPACT OF A NODE -H OLE I NJECTION


A. Stress Under High VG /VD Ratio
Fig. 10(a) shows the plots of the measured IDLIN versus Fig. 12. Measured fixed time ΔIDSAT from forward and reverse mode
sense in devices having (a) long and (b) short LCH versus stress VG , for
VG at fixed VD in devices having different L CH (but fixed TOX ) VB = 0 V and VB <0 V stress, SDE MOSFETs [7]. Stress time is 4 ks.
[43]. Fig. 10(b) shows the plots of the inverse of estimated
lifetime versus VG at various stress VD (but fixed L CH and for VG = VD stress; the reduction in n and recovery are more
TOX ); lifetime equals the time to reach 10% IDSAT , measured for the thinner TOX device. The results can be explained as
in the reverse mode [45]. These are relatively “long-channel” follows.
devices and show bell-shaped IDLIN (and 1/lifetime) versus The AHI process, illustrated using Fig. 1(b), is triggered
VG dependence for VG variation up to VD . It is interesting at VG = VD stress. By assuming power-law time kinetics
to remark that the peak IDLIN point shifts toward higher but lower slope n for the AHI-initiated defect generation,
VG as L CH is scaled (as the device slowly becomes “short- the reduction in n at higher VG stress can be explained (e.g.,
channel” type, but the “bell-shape” is present) [see Fig. 10(a)]. Fig. 11 and also in [17], see Table I). However, the chemical
However, the increase in HCD is observed for VG > VD , which nature of defects generated by the AHI-initiated process is
is more prominent at short L CH [see Fig. 10(a)] and higher likely different from those generated by the pure HCD process,
VD /VG ratio [see Fig. 10(b)]. Note that this feature has been as they show a lower power-law slope during stress and
reported in [58] first, which prompted a shift in the under- recovery after stress. The contribution due to pure HCD (CHE)
standing of the HCD mechanism from field to energy-driven would reduce, while that from the AHI-initiated process would
process. Note that the original explanation of increased HCD increase as VG is increased (for classical devices), and the
at VG > VD is the generation of defects by multi (many) cold impact will be more at thinner TOX (when everything else
electron interaction process [12], [43]–[45]. An alternative stays same), which can explain the features reported in Fig. 11.
explanation is the triggering of the AHI-initiated process for However, AHI would reduce in short L CH devices, even having
both Fig. 10(a) and (b) and electron trapping for Fig. 10(b); thinner TOX since the stress biases (VG and VD ) are reduced
trapping is less significant in Fig. 10(a) due to thinner TOX . and if the VG /VD ratio is not high. Therefore, short L CH devices
The impact of the AHI-initiated trap generation can be show the same n at various VG /VD ratio (VG ≤VD ) [see Fig. 6]
directly estimated using CP measurements. Fig. 11 shows and no poststress recovery [33], as they are governed by the
the plots of the measured ICP time kinetics during [see CHE process.
Fig. 11(a)] and after [see Fig. 11(b)] stress at VG = VD /2 and
VG = VD conditions in devices having different TOX [11].
HCD is due to the pure CHE process and shows a power-law B. Stress Under Reverse VB
slope of n ∼ 0.6 at VG = VD /2 stress and no recovery after Fig. 12 shows the plots of the measured IDSAT versus
stress. The power-law slope reduces, and recovery is observed, VG for VB = 0 V and VB <0 V stress, in devices having

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
10 IEEE TRANSACTIONS ON ELECTRON DEVICES

A potential candidate for AHI-initiated defects is hydro-


gen (H) bridge (O3 –Si–H) [70], which can convert into
Oxygen vacancy (Ov) when H is dislodged. The generated
Ov centers can repassivate if H atoms are nearby and can
be probed by CP if these are located near the channel/gate
oxide interface (depends on the CP frequency) [71]. Note that
Ov defects (∼neutral electron traps) in the oxide bulk would
result in stress-induced leakage current (SILC) during TDDB
stress [72], [73]. Also note that the power-law slope for SILC
is n∼0.35 [74], lower than that (n∼0.5) seen for most CHE
stress. Such neutral electron trap creation is also suggested
Fig. 13. Measured ΔNIT time kinetics (a) during and (b) after stress after the hole injection in Fig. 3(b) [26]. However, this is
without and with reverse VB in LDD MOSFET [11].
purely speculation at this moment, and other candidates are
indeed possible for the AHI defects.
relatively long [see Fig. 12(a)] and short [see Fig. 12(b)] L CH ,
and stressed, respectively, at relatively higher and lower VD VI. C ONCLUSION
(note that the L CH and VD values in Fig. 12(a) are smaller than The following conclusions can be drawn regarding the
the values in Fig. 3) [7]. The relatively long and short L CH HCD physical mechanism in n-channel MOSFETs and the
devices show, respectively, the bell-shaped and monotonic VG suitability of TCAD for predictive simulation.
dependence as expected. However, higher IDSAT is observed HCD can be classified whether hot carriers are injected into
for VB <0 V stress due to secondary ionization related effects. the gate oxide or not, which depends on stress VD . Generation
HCD stays localized near the drain junction for both devices, of acceptor-like defects is always present, which is augmented
without and with reverse VB stress, since the drift is larger by injection and trapping of electrons at high VG /VD ratio and
for reverse compared with the forward measurement mode. high VD above ∼3 V and holes at low VG /VD ratio and high VD
The impact of VB is more at long L CH and results in higher above ∼4.7 V, in long L CH , thicker TOX devices. Trapping is
IDSAT , more so at higher stress VG . Hence, IDSAT at VB <0 not seen in short L CH , thinner TOX devices at VD <3 V stress,
V stress does not fall off as sharply as that for the VB = 0 while trap generation is even down to VD ∼1 V stress. Hence,
V case, which can be explained using the AHI-induced trap it is imperative that the TCAD framework for different classes
generation effect. The impact of reverse VB stress is lower in of devices should contain these relevant physical processes.
the short L CH device due to reduced body effect. Defect generation under CHE stress is linked to the genera-
Fig. 13 shows the plots of the measured ICP time kinetics tion of hot holes from the impact ionization process (∼ISUB ).
during [see Fig. 13(a)] and after [see Fig. 13(b)] stress at Both peaks at mid-VG in long L CH devices at high VD stress
VG = VD /2 condition but without and with VB (reverse and in some short L CH devices under low VD stress when
bias) [11]. The power-law slope reduces during stress, and devices do not suffer from SCE. However, both peaks at
recovery is observed after stress under VB <0 V condition. VG ∼ VD stress in short L CH devices when SCE is large.
Note that the core CHE process should reduce under reverse The carrier heating follows this trend, and hence, it is not
VB due to the increase in VT0 and reduction in ID that triggers possible to ascertain if hot holes or hot electrons are the
the carrier heating process. This is indeed the case in FDSOI candidates that produce these defects. It is highly probable that
when stressed under reverse back-gate bias [12]. Therefore, the hot carriers collide with near interfacial defect precursors
Fig. 13 can be explained by again invoking the AHI-initiated to generate defects, and the injection into the oxide is not
process, as higher electron energy in the substrate, due to the needed. However, coinjection of holes and electrons into the
secondary impact ionization event, would also result in higher oxide, when present (VD > 4.7 V), generates additional defects.
electron injection and AHI at VB <0 V stress (see [11] and [29] TCAD presently can explain the ISUB versus VG dependence in
for details). long and short L CH devices. However, a connection needs to be
established between carrier heating and the defect generation
processes at different VG /VD ratio, without and with the
C. Discussion on Defects and Time Kinetics situation when hot carriers are coinjected into the gate oxide
Based on the poststress buildup of defects after HCD stress (for high-voltage devices).
at VG <VD /2 [68] and irradiation [69], the following is a likely The CHE process would reduce at higher VG /VD ratio (since
process for AHI-induced generated traps. For large VG /VD or the pinch-off region and carrier heating reduce) and at reverse
VB <0 V stress, hot electrons, upon injection from the substrate VB (as ID reduces and so the density of hot carriers). However,
to the gate near the drain region of the device, trigger impact AHI-initiated defect generation is seen in such cases, triggered
ionization in the gate. Resulting holes are injected back into by hot electron injection from the channel for VD > 3 V stress.
the oxide, and they break hydrogen (H)-passivated defects to AHI triggered by tunneling of channel electrons and PBTI can
produce atoms (H) and/or ions (H+ ). This is followed by the also generate defects, especially in thinner TOX devices and
diffusion (or drift) of H (or H+ ) toward the channel/gate oxide higher VG /VD stress. To the best of our knowledge, AHI and
interface and subsequent generation of defects at or near the related processes are not available in TCAD at present.
channel/gate oxide interface, which can be measured using the Both electron and hole trappings, when present, saturate at
CP method. longer stress time. However, defect generation shows power-

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
MAHAPATRA AND SHARMA: REVIEW OF HCD IN n-CHANNEL MOSFETs—I 11

law time dependence although with a wide range of power-law [17] E. Takeda, A. Shimizu, and T. Hagiwara, “Role of hot-hole injection in
slope. The reason for this is not fully understood although a hot-carrier effects and the small degraded channel region in MOSFET’s,”
IEEE Electron Device Lett., vol. EDL-4, no. 9, pp. 329–331, Sep. 1983,
few possibilities have been discussed. A suitable framework doi: 10.1109/EDL.1983.25751.
for the time kinetics of trap generation and trapping needs to [18] P. Heremans, R. Bellens, G. Groeseneken, and H. E. Maes, “Consis-
be adopted in TCAD. In particular, the setup should be able to tent model for the hot-carrier degradation in n-channel and p-channel
explain the wide range of measured time slopes for the CHE MOSFETs,” IEEE Trans. Electron Devices, vol. ED-35, no. 12,
pp. 2194–2209, Dec. 1988, doi: 10.1109/16.8794.
trap generation process and the lower (presumably) slope for [19] P. Heremans, G. Van den Bosch, R. Bellens, G. Groeseneken, and
the AHI-initiated trap generation process. H. E. Maes, “Temperature dependence of the channel hot-carrier degra-
Trap generation and trapping at various parts of the device dation of n-channel MOSFET’s,” IEEE Trans. Electron Devices, vol. 37,
no. 4, pp. 980–993, Apr. 1990, doi: 10.1109/16.52433.
differently affects the time kinetics of different parameters of [20] V.-H. Chan and J. E. Chung, “Two-stage hot-carrier degradation and its
the device. It is understood that the present TCAD maturity is impact on submicrometer LDD NMOSFET lifetime prediction,” IEEE
sufficient to handle the connection between generated charges Trans. Electron Devices, vol. 42, no. 5, pp. 957–962, May 1995, doi:
and changes in the device parametric drift. 10.1109/16.381994.
[21] A. Raychaudhuri, M. J. Deen, W. S. Kwan, and M. I. H. King,
R EFERENCES “Features and mechanisms of the saturating hot-carrier degradation
in LDD NMOSFETs,” IEEE Trans. Electron Devices, vol. 43, no. 7,
[1] R. B. Fair and R. C. Sun, “Threshold voltage instability in MOSFET’s pp. 1114–1122, Jul. 1996, doi: 10.1109/16.502423.
due to channel hot hole emission,” in IEDM Tech. Dig., Washington,
DC, USA, Dec. 1980, pp. 746–749, doi: 10.1109/IEDM.1980.189944. [22] S. Mahapatra, C. D. Parikh, V. R. Rao, C. R. Viswanathan, and J. Vasi,
[2] E. Takeda and N. Suzuki, “An empirical model for device degradation “A comprehensive study of hot-carrier induced interface and oxide trap
due to hot-carrier injection,” IEEE Electron Device Lett., vol. EDL-4, distributions in MOSFETs using a novel charge pumping technique,”
no. 4, pp. 111–113, Apr. 1983, doi: 10.1109/EDL.1983.25667. IEEE Trans. Electron Devices, vol. 47, no. 1, pp. 171–177, Jan. 2000,
[3] D. S. Ang and C. H. Ling, “On the time-dependent degradation of LDD doi: 10.1109/16.817583.
n-MOSFETs under hot-carrier stress,” Microelectron. Rel., vol. 39, no. 9, [23] S. Mahapatra, C. D. Parikh, V. R. Rao, C. R. Viswanathan, and
pp. 1311–1322, 1999, doi: 10.1016/S0026-2714(99)00053-0. J. Vasi, “Device scaling effects on hot-carrier induced interface and
[4] D. Varghese, M. A. Alam, and B. Weir, “A generalized, I B -independent, oxide-trapped charge distributions in MOSFETs,” IEEE Trans. Electron
physical HCI lifetime projection methodology based on universality of Devices, vol. 47, no. 4, pp. 789–796, Apr. 2000, doi: 10.1109/16.830995.
hot-carrier degradation,” in Proc. IEEE Int. Rel. Phys. Symp., May 2010, [24] T. Tsuchiya, T. Kobayashi, and S. Nakajima, “Hot-carrier-injected oxide
pp. 1091–1094, doi: 10.1109/IRPS.2010.5488666. region and hot-electron trapping as the main cause in Si nMOS-
[5] L. Brunet et al., “Impact of backside interface on hot carriers degra- FET degradation,” IEEE Trans. Electron Devices, vol. ED-34, no. 2,
dation of thin film FDSOI Nmosfets,” in Proc. IEEE Int. Rel. Phys. pp. 386–391, Feb. 1987, doi: 10.1109/T-ED.1987.22934.
Symp. (IRPS), Anaheim, CA, USA, Apr. 2012, pp. 3B.2.1–3B.2.5, doi: [25] T. Tsuchiya, “Trapped-electron and generated interface-trap effects
10.1109/IRPS.2012.6241806. in hot-electron-induced MOSFET degradation,” IEEE Trans. Electron
[6] Y. Qu et al., “Ultra fast (<1 ns) electrical characterization of self- Devices, vol. ED-34, no. 11, pp. 2291–2296, Nov. 1987, doi: 10.1109/T-
heating effect and its impact on hot carrier injection in 14 nm ED.1987.23234.
FinFETs,” in IEDM Tech. Dig., Dec. 2017, pp. 39.2.1–39.2.4, doi: [26] B. S. Doyle et al., “The generation and characterization of electron and
10.1109/IEDM.2017.8268520. hole traps created by hole injection during low gate voltage hot-carrier
[7] A. Bravaix, C. Guerin, V. Huard, D. Roy, J. M. Roux, and E. Vincent, stressing of n-MOS transistors,” IEEE Trans. Electron Devices, vol. 37,
“Hot-carrier acceleration factors for low power management in DC-AC no. 8, pp. 1869–1876, Aug. 1990, doi: 10.1109/16.57138.
stressed 40 nm NMOS node at high temperature,” in Proc. Int. Rel. Phys. [27] B. Doyle, M. Bourcerie, J.-C. Marchetaux, and A. Boudou, “Interface
Symp., Apr. 2009, pp. 531–548, doi: 10.1109/IRPS.2009.5173308. state creation and charge trapping in the medium-to-high gate voltage
[8] J. J. Sanchez, K. K. Hsueh, and T. A. DeMassa, “Drain-engineered range (Vd /2 ≥ Vg ≥ Vd ),” IEEE Trans. Electron Devices, vol. 37, no. 3,
hot-electron-resistant device structures: A review,” IEEE Trans. Elec- pp. 744–754, Mar. 1990, doi: 10.1109/16.47781.
tron Devices, vol. 36, no. 6, pp. 1125–1132, Jun. 1989, doi: [28] C. Hu, S. C. Tam, F.-C. Hsu, P.-K. Ko, T.-Y. Chan, and K. W. Terrill,
10.1109/16.24357. “Hot-electron-induced MOSFET degradation–model, monitor, and
[9] F.-C. Hsu and K.-Y. Chiu, “Temperature dependence of hot-electron- improvement,” IEEE J. Solid-State Circuits, vol. SSC-20, no. 1,
induced degradation in MOSFET’s,” IEEE Electron Device Lett., pp. 295–305, Feb. 1985, doi: 10.1109/JSSC.1985.1052306.
vol. EDL-5, no. 5, pp. 148–150, May 1984, doi: 10.1109/EDL. [29] S. Mahapatra, S. Shukuri, and J. Bude, “CHISEL flash EEPROM.
1984.25865. I. Performance and scaling,” IEEE Trans. Electron Devices, vol. 49,
[10] K. Mistry and B. Doyle, “The role of electron trap creation in enhanced no. 7, pp. 1296–1301, Jul. 2002, doi: 10.1109/TED.2002.1013289.
hot-carrier degradation during AC stress (n-channel MOSFET),” IEEE
[30] K. F. Schuegraf and C. Hu, “Hole injection SiO2 breakdown model for
Electron Device Lett., vol. 11, no. 6, pp. 267–269, Jun. 1990, doi: very low voltage lifetime extrapolation,” IEEE Trans. Electron Devices,
10.1109/55.55276. vol. 41, no. 5, pp. 761–767, May 1994, doi: 10.1109/16.285029.
[11] S. Mahapatra, D. Saha, D. Varghese, and P. B. Kumar, “On the
generation and recovery of interface traps in MOSFETs subjected to [31] M. A. Alam, J. Bude, and A. Ghetti, “Field acceleration for oxide
NBTI, FN, and HCI stress,” IEEE Trans. Electron Devices, vol. 53, breakdown-can an accurate anode hole injection model resolve the
E vs. 1/E controversy?” in Proc. IEEE 38th Annu. Int. Rel. Phys.
no. 7, pp. 1583–1592, Jul. 2006, doi: 10.1109/TED.2006.876041.
[12] W. Arfaoui et al., “Energy-driven hot-carrier model in advanced nodes,” Symp., San Jose, CA, USA, Apr. 2000, pp. 21–26, doi: 10.1109/REL
in Proc. IEEE Int. Rel. Phys. Symp., Jun. 2014, pp. XT.12.1–XT.12.5, PHY.2000.843886.
doi: 10.1109/IRPS.2014.6861189. [32] P. Heremans, H. E. Maes, and N. Saks, “Evaluation of hot carrier degra-
[13] D. Varghese, P. Moens, and M. A. Alam, “On-state hot carrier dation of N-channel MOSFET’s with the charge pumping technique,”
degradation in drain-extended NMOS transistors,” IEEE Trans. Elec- IEEE Electron Device Lett., vol. EDL-7, no. 7, pp. 428–430, Jul. 1986.
tron Devices, vol. 57, no. 10, pp. 2704–2710, Oct. 2010, doi: [33] A. Chasm et al., “Hot-carrier analysis on nMOS Si FinFETs with
10.1109/TED.2010.2059632. solid source doped junction,” in Proc. IEEE Int. Rel. Phys. Symp.
[14] T. Kamata, K. Tanabashi, and K. Kobayashi, “Substrate current due to (IRPS), Pasadena, CA, USA, Apr. 2016, pp. 4B-4–4B-4B-4-6, doi:
impact ionization in MOS-FET,” Jpn. J. Appl. Phys., vol. 15, no. 6, 10.1109/IRPS.2016.7574535.
pp. 1127–1133, Jun. 1976. [34] E. Li, E. Rosenbaum, J. Tao, G. C.-F. Yeap, M. Lin, and P. Fang, “Hot
[15] K. K. Ng and G. W. Taylor, “Effects of hot-carrier trapping in n- and p- carrier effects in nMOSFETs in 0.1 μm CMOS technology,” in Proc.
channel MOSFET’s,” IEEE Trans. Electron Devices, vol. ED-30, no. 8, IEEE 37th Annu. Int. Rel. Phys. Symp., San Diego, CA, USA, Mar. 1999,
pp. 871–876, Aug. 1983, doi: 10.1109/T-ED.1983.21229. pp. 253–258, doi: 10.1109/RELPHY.1999.761622.
[16] K. R. Hofmann, C. Werner, W. Weber, and G. Dorda, “Hot-electron [35] S. J. Wang, J. M. Sung, and S. A. Lyon, “Relationship between hole
and hole-emission effects in short n-channel MOSFET’s,” IEEE Trans. trapping and interface state generation in metal-oxide-silicon structures,”
Electron Devices, vol. ED-32, no. 3, pp. 691–699, Mar. 1985, doi: Appl. Phys. Lett., vol. 52, no. 17, pp. 1431–1433, Apr. 1988, doi:
10.1109/T-ED.1985.22000. 10.1063/1.99690.

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
12 IEEE TRANSACTIONS ON ELECTRON DEVICES

[36] Z. Chen et al., “On the mechanism for interface trap generation in MOS [55] X. Federspiel, F. Cacho, and D. Roy, “Experimental characterization of
transistors due to channel hot carrier stressing,” IEEE Electron Device the interactions between HCI, off-state and BTI degradation modes,” in
Lett., vol. 21, no. 1, pp. 24–26, Jan. 2000, doi: 10.1109/55.817441. Proc. IEEE Int. Integr. Rel. Workshop Final Rep., South Lake Tahoe,
[37] C. Min-Liang, L. Chung-Wai, W. T. Cochran, W. Jungling, C. Dziuba, CA, USA, Oct. 2011, pp. 133–136.
and Y. Tungsheng, “Suppression of hot-carrier effects in submicrometer [56] W. Arfaoui, X. Federspiel, P. Mora, M. Rafik, D. Roy, and A. Bravaix,
CMOS technology,” IEEE Trans. Electron Devices, vol. ED-35, no. 12, “Experimental analysis of defect nature and localization under hot-
pp. 2210–2220, Dec. 1988, doi: 10.1109/16.8795. carrier and bias temperature damage in advanced CMOS nodes,” in Proc.
[38] K.-L. Chen, S. A. Saller, I. A. Groves, and D. B. Scott, “Reliability IEEE Int. Integr. Rel. Workshop Final Rep., Oct. 2013, pp. 78–83, doi:
effects on MOS transistors due to hot-carrier injection,” IEEE Trans. 10.1109/IIRW.2013.6804163.
Electron Devices, vol. ED-32, no. 2, pp. 386–393, Feb. 1985, doi: [57] S. Mahapatra and R. Saikia, “On the universality of hot carrier
10.1109/T-ED.1985.21953. degradation: Multiple probes, various operating regimes, and different
[39] A. Joshi and D.-L. Kwong, “Study of the growth temperature depen- MOSFET architectures,” IEEE Trans. Electron Devices, vol. 65, no. 8,
dence of performance and reliability of thin MOS gate oxides,” IEEE pp. 3088–3094, Aug. 2018, doi: 10.1109/TED.2018.2842129.
Trans. Electron Devices, vol. 39, no. 9, pp. 2099–2107, Sep. 1992, doi: [58] S. E. Rauch and G. La Rosa, “The energy-driven paradigm of NMOS-
10.1109/16.155883. FET hot-carrier effects,” IEEE Trans. Device Mater. Rel., vol. 5, no. 4,
[40] J.-S. Goo, Y.-G. Kim, H. L’Yee, H.-Y. Kwon, and H. Shin, “An analytical pp. 701–705, Dec. 2005, doi: 10.1109/TDMR.2005.860560.
model for hot-carrier-induced degradation of deep-submicron n-channel [59] J. H. Stathis and E. Cartier, “Atomic hydrogen reactions with Pb centers
LDD MOSFETs,” Solid-State Electron., vol. 38, no. 6, pp. 1191–1196, at the (100) Si/SiO2 interface,” Phys. Rev. Lett., vol. 25, pp. 2745–2748,
1995, doi: 10.1016/0038-1101(94)00221-Z. Apr. 1994, doi: 10.1103/PhysRevLett.72.2745.
[41] C.-L. Lou, W.-K. Chim, D. S.-H. Chan, and T. Pan, “A novel [60] A. Stesmans, “Passivation of Pb0 and Pb1 interface defects in thermal
single-device DC method for extraction of the effective mobility (100) Si/SiO2 with molecular hydrogen,” Appl. Phys. Lett., vol. 68,
and source-drain resistances of fresh and hot-carrier degraded drain- no. 15, pp. 2076–2078, Apr. 1996, doi: 10.1063/1.116308.
engineered MOSFET’s,” IEEE Trans. Electron Devices, vol. 45, no. 6, [61] G. Pobegen, M. Nelhiebel, and T. Grasser, “Detrimental impact
pp. 1317–1323, Jun. 1998, doi: 10.1109/16.678559. of hydrogen passivation on NBTI and HC degradation,” in Proc.
[42] S. Mukhopadhyay et al., “Trap generation in IL and HK layers IEEE Int. Rel. Phys. Symp. (IRPS), Anaheim, CA, USA, Apr. 2013,
during BTI/TDDB stress in scaled HKMG N and P MOSFETs,” in pp. XT.10.1–XT.10.6.
Proc. IEEE Int. Rel. Phys. Symp., Waikoloa, HI, USA, Jun. 2014, [62] K. Joshi et al., “HKMG process impact on N, P BTI: Role of ther-
pp. GD.3.1–GD.3.11, doi: 10.1109/IRPS.2014.6861146. mal IL scaling, IL/HK integration and post HK nitridation,” in Proc.
[43] W. Arfaoui, X. Federspiel, A. Bravaix, P. Mora, A. Cros, and D. Roy, IEEE Int. Rel. Phys. Symp. (IRPS), Anaheim, CA, USA, Apr. 2013,
“Application of compact HCI model to prediction of process effect in pp. 4C.2.1–4C.2.10.
28FDSOI technology,” in Proc. IEEE Int. Integr. Rel. Workshop Final [63] C. Prasad et al., “Self-heat reliability considerations on Intel’s 22 nm tri-
Rep. (IIRW), South Lake Tahoe, CA, USA, Oct. 2014, pp. 69–72, doi: gate technology,” in Proc. IEEE Int. Rel. Phys. Symp. (IRPS), Anaheim,
10.1109/IIRW.2014.7049513. CA, USA, Apr. 2013, pp. 5D.1.1–5D.1.5.
[44] A. Bravaix et al., “Impact of the gate-stack change from 40 nm node [64] D. Varghese et al., “Off-state degradation in drain-extended NMOS tran-
SiON to 28 nm high-K metal gate on the hot-carrier and bias tempera- sistors: Interface damage and correlation to dielectric breakdown,” IEEE
ture damage,” in Proc. IEEE Int. Rel. Phys. Symp. (IRPS), Anaheim, Trans. Electron Devices, vol. 54, no. 10, pp. 2669–2678, Oct. 2007.
CA, USA, Apr. 2013, pp. 2D.6.1–2D.6.9, doi: 10.1109/IRPS.2013. [65] S. Chakravarthi, A. Krishnan, V. Reddy, C. F. Machala, and S. Krishnan,
6531961. “A comprehensive framework for predictive modeling of negative bias
[45] C. Guerin, V. Huard, and A. Bravaix, “The energy-driven hot-carrier temperature instability,” in Proc. IEEE Int. Rel. Phys. Symp., Phoenix,
degradation modes of nMOSFETs,” IEEE Trans. Device Mater. Rel., AZ, USA, Apr. 2004, pp. 273–282.
vol. 7, no. 2, pp. 225–235, Jun. 2007, doi: 10.1109/TDMR.2007. [66] T.-C. Ong, M. Levi, P.-K. Ko, and C. Hu, “Recovery of threshold voltage
901180. after hot-carrier stressing,” IEEE Trans. Electron Devices, vol. ED-35,
[46] T. Di Gilio and A. Bravaix, “Lifetime prediction of ultra-thin gate no. 7, pp. 978–984, Jul. 1988, doi: 10.1109/16.3354.
oxide PMOSFETs submitted to hot hole injections,” in Proc. IEEE Int. [67] D. Varghese, D. Saha, S. Mahapatra, K. Ahmed, F. Nouri, and M. Alam,
Integr. Rel. Workshop, S. Lake Tahoe, CA, USA, Oct. 2005, p. 5, doi: “On the dispersive versus arrhenius temperature activation of NBTI time
10.1109/IRWS.2005.1609562. evolution in plasma nitrided gate oxides: Measurements, theory, and
[47] P. Magnone et al., “Impact of hot carriers on nMOSFET variability in 45- implications,” in IEDM Tech. Dig., Washington, DC, USA, Dec. 2005,
and 65-nm CMOS technologies,” IEEE Trans. Electron Devices, vol. 58, pp. 684–687, doi: 10.1109/IEDM.2005.1609444.
no. 8, pp. 2347–2353, Aug. 2011, doi: 10.1109/TED.2011.2156414. [68] E. de Schrijver, P. Heremans, R. Bellens, G. Groeseneken, and
[48] M. Bina et al., “Predictive hot-carrier modeling of n-channel MOS- H. E. Maes, “Analysis of post-stress effects in passivated MOSFET’s
FETs,” IEEE Trans. Electron Devices, vol. 61, no. 9, pp. 3103–3110, after hot-carrier stress,” in Proc. 21st Eur. Solid State Device Res.
Sep. 2014, doi: 10.1109/TED.2014.2340575. Conf. (ESSDERC), Montreux, Switzerland, 1991, pp. 437–440, doi:
[49] A. Kerber, S. Cimino, F. Guarin, and T. Nigam, “Assessing device 10.1016/0167-9317(91)90259-G.
reliability margin in scaled CMOS technologies using ring oscillator cir- [69] N. S. Saks and D. B. Brown, “Interface trap formation via the two-stage
cuits,” in Proc. IEEE Electron Devices Technol. Manuf. Conf. (EDTM), H+ process,” IEEE Trans. Nucl. Sci., vol. 36, no. 6, pp. 1848–1857,
Feb. 2017, pp. 28–30, doi: 10.1109/EDTM.2017.7947495. Dec. 1989, doi: 10.1109/23.45378.
[50] C. Schlunder et al., “Circuit relevant HCS lifetime assessments [70] S. Ling, A.-M. El-Sayed, F. Lopez-Gejo, M. B. Watkins, V. V. Afanas’ev,
at single transistors with emulated variable loads,” in Proc. IEEE and A. L. Shluger, “A computational study of Si–H bonds as pre-
Int. Rel. Phys. Symp. (IRPS), Apr. 2017, pp. 2D-2.1–2D-2.7, doi: cursors for neutral E’ centres in amorphous silica and at the Si/SiO2
10.1109/IRPS.2017.7936262. interface,” Microelectron. Eng., vol. 109, pp. 310–313, Sep. 2013, doi:
[51] M. Duan et al., “Interaction between hot carrier aging and PBTI 10.1016/j.mee.2013.03.028.
degradation in nMOSFETs: Characterization, modelling and lifetime [71] M. Masuduzzaman, A. E. Islam, and M. A. Alam, “Physics and mech-
prediction,” in Proc. IEEE Int. Rel. Phys. Symp. (IRPS), Monterey, anisms of dielectric trap profiling by multi-frequency charge pumping
CA, USA, Apr. 2017, pp. XT-5.1–XT-5.7, doi: 10.1109/IRPS.2017. (MFCP) method,” in Proc. IEEE Int. Rel. Phys. Symp., Montreal, QC,
7936419. Canada, Apr. 2009, pp. 13–20, doi: 10.1109/IRPS.2009.5173218.
[52] A. Kerber and E. A. Cartier, “Reliability challenges for CMOS tech- [72] D. Ielmini, A. S. Spinelli, M. A. Rigamonti, and A. L. Lacaita,
nology qualifications with hafnium oxide/titanium nitride gate stacks,” “Modeling of SILC based on electron and hole tunneling. II. steady-
IEEE Trans. Device Mater. Rel., vol. 9, no. 2, pp. 147–162, Jun. 2009, state,” IEEE Trans. Electron Devices, vol. 47, no. 6, pp. 1266–1272,
doi: 10.1109/TDMR.2009.2016954. Jun. 2000, doi: 10.1109/16.842972.
[53] S. Mukhopadhyay et al., “Trap generation in IL and HK layers [73] M. A. Alam, “SILC as a measure of trap generation and predictor of
during BTI/TDDB stress in scaled HKMG N and P MOSFETs,” in T B D in ultrathin oxides,” IEEE Trans. Electron Devices, vol. 49, no. 2,
Proc. IEEE Int. Rel. Phys. Symp., Waikoloa, HI, USA, Jun. 2014, pp. 226–231, Feb. 2002, doi: 10.1109/16.981211.
pp. GD.3.1–GD.3.11, doi: 10.1109/IRPS.2014.6861146. [74] E. Y. Wu, J. Sune, and W. Lai, “On the Weibull shape factor of intrinsic
[54] E. Amat et al., “Channel hot-carrier degradation in short chan- breakdown of dielectric films and its accurate experimental determina-
nel devices with high-k/metal gate stacks,” in Proc. Spanish tion. Part II: Experimental results and the effects of stress conditions,”
Conf. Electron Devices, Santiago de Compostela, Spain, Feb. 2009, in IEEE Trans. Electron Devices, vol. 49, no. 12, pp. 2141–2150,
pp. 238–241. Dec. 2002.

Authorized licensed use limited to: University of Exeter. Downloaded on June 09,2020 at 15:01:54 UTC from IEEE Xplore. Restrictions apply.

You might also like