• 沒有找到結果。

Statistical Characterization and Modeling of the Temporal Evolutions of Delta V-t Distribution in NBTI Recovery in Nanometer MOSFETs

N/A
N/A
Protected

Academic year: 2021

Share "Statistical Characterization and Modeling of the Temporal Evolutions of Delta V-t Distribution in NBTI Recovery in Nanometer MOSFETs"

Copied!
7
0
0

加載中.... (立即查看全文)

全文

(1)

Abstract—NBTI trapped charge characteristics and recovery

mechanisms are examined by a statistical study of individual trapped charge emissions in nanoscale HfSiON/metal gate pMOS-FETs. We measure individual trapped charge emission times in NBTI recovery in a large number of devices. The characteristic time distributions of the first three emitted holes are obtained. The distributions can be well modeled by using a thermally-assisted tunnel (ThAT) detrapping model. NBTI trapped charge energy and spatial distributions and its activation energy distribution in the ThAT model are discussed and extracted. Based on the ThAT model and measured result of single-charge induced Vt shifts,

we develop a statistical NBTI recovery ΔVt evolution model.

Our model can well reproduce the temporal evolutions of a ΔVt

distribution in a number of NBTI stressed nanometer MOSFETs in relaxation.

Index Terms—Activation energy, charge emission time, negative

bias temperature instability, recovery, statistical model.

I. INTRODUCTION

N

EGATIVE Bias Temperature Instability (NBTI) has been recognized as one of the most important reliability issues in ultra-thin gate oxide CMOS devices [1]–[4]. The use of a high-k gate dielectric even expedites NBTI degradation [3], [4]. Unlike most reliability effects, NBTI Vt degradation recovers

partly after the removal of stress [5]. Several circuit techniques exploiting NBTI recovery have been proposed to alleviate NBTI severity in memory and logic circuits [6], [7]. To improve a design window that is tightened by several variability sources, the integration of NBTI degradation and recovery character-istics into a circuit simulation is called for in modern CMOS circuit design [8]. Hence, underlying degradation and recovery

Manuscript received August 22, 2012; revised January 6, 2013 and January 8, 2013; accepted January 10, 2013. Date of publication February 8, 2013; date of current version February 20, 2013. The authors would like to acknowledge financial support from National Science Council, Taiwan, under contract NSC 99-2221-E-009-169-MY3 and from Ministry of Education in Taiwan under ATU Program. The review of this paper was arranged by Editor B. Kaczer.

J.-P. Chiu, Y.-H. Liu, H.-D. Hsieh, C.-W. Li, and T. Wang are with the De-partment of Electronics Engineering, National Chiao Tung University, Hsinchu 300, Taiwan (e-mail: twang@cc.nctu.edu.tw).

M.-C. Chen is with National Nano Device Laboratories, Hsinchu 30078, Taiwan.

Color versions of one or more of the figures in this paper are available online at http://ieeexplore.ieee.org.

Digital Object Identifier 10.1109/TED.2013.2240390

mechanisms of NBTI must be understood in order to make a meaningful projection of a device/circuit lifetime. While the reaction-diffusion (RD) model [2], [9]–[11] and a charge trap-ping model [12]–[17] are commonly adopted for NBTI degra-dation, no consensus regarding NBTI recovery mechanisms has been reached. Hydrogen back-diffusion in the framework of the RD model was first proposed to explain a NBTI recovery phenomenon [11]. A ΔVtrelaxation of the form of ΔVt(t) = Vt(0)(1t/2t0/



1 + t/t0) is anticipated from the back

diffusion theory. Wang et al. used a fast transient technique to characterize NBTI recovery and found a log(t) dependence of ΔVton relaxation time in a time span from 10−3s to 102s [18].

Moreover, they characterized individual trapped hole emissions in small-area devices and proposed a thermally assisted charge tunnel detrapping (ThAT) model for a recoverable component in relaxation [18], [19]. A log(t) dependence was also reported in [20] where the authors proposed a dispersive transport model within the RD framework to overcome the apparent deficiency of the RD model with respect to relaxation time dependence. Alam et al. ascribed a log(t) recovery transient to a fast charge detrapping process on top of slower Nit re-passivation

as encapsulated by the RD model [21], [22]. More recently, Grasser et al. used a time dependent defect spectroscopy method to investigate NBTI recovery and concluded that NBTI recovery is due to thermally-assisted discharging of traps and no diffusion process is involved [17]. Their conclusions are consistent with [18].

In most of earlier works, NBTI behavior was characterized in large-area devices [20], [23] and studies were based on an average and continuous Vt evolution [20]–[23]. In contrast,

NBTI degradation/recovery proceeds in discrete steps in small area devices [18], [24]. Due to the discrete nature of a Vt

evolution, we are able to measure individual trapped charge emission times and each single trapped charge induced Vtshift.

Statistical characterization of NBTI degradation/recovery in small area devices therefore provides important information to help understand a responsible mechanism as well as trapped charge characteristics such as trapped charge energy and spatial distributions and its activation energy distribution. Moreover, a NBTI model accounting for an entire ΔVt distribution is

required in a worst-case circuit simulation. In this paper, a statistical model based on a ThAT model and an extracted activation energy distribution will be developed for a NBTI recovery ΔVtdistribution and its temporal evolutions.

(2)

Fig. 1. Example ΔIdand Vttraces in NBTI relaxation. τ1, τ2and τ3 are

the 1st, the 2nd and the 3rd trapped hole emission times, respectively. Δvt,i

(i = 1, 2, 3) represents a single emitted charge induced threshold voltage shift. II. STATISTICALCHARACTERIZATION

OFNBTI RECOVERY

We characterize NBTI recovery in high-k (HfSiON) gate di-electric and metal gate pMOSFETs. The devices have a nominal gate length of 35 nm, a gate width of 70 nm and an effective oxide thickness of ∼1.0 nm. The devices are first stressed at Vg=−1.8 V for 100 sec. The recovery characterization

scheme is similar to [13], i.e., in a relaxation-measurement-relaxation sequence. Both stress and recovery characterizations are performed at room temperature. In measurement phase, the drain voltage is−0.05 V and the gate voltage is chosen such that the pre-stress drain current is ∼500 nA. Drain current variations are recorded using Agilent B1500. The cumulative relaxation time in each device measurement is 1000 s. A corre-sponding ΔVttrace is obtained from a measured ΔIddivided

by a transconductance [25]. Fig. 1 shows example ΔId and

Vt traces in NBTI relaxation. Each abrupt Vt change (Δvt,i)

in the trace is due to a single trapped hole emission, where i denotes a charge detrapping sequence number. Any identifiable RTN signals are not counted. We collect all Δvt,i in ∼170

devices. The magnitude distributions of Δvt,1, Δvt,2 and all

collected Δvt are plotted in Fig. 2. Emitted trapped charges

have a similar Δvtdistribution, characterized by an exponential

function f (|Δvt|) = exp(−|Δvt|/σ)/σ with a σ of 3.3 mV.

The origins and the distribution of the Δvt have been studied

thoroughly and the exponential distribution is realized due to a random substrate dopant induced current path percolation effect [24], [26].

A. Trapped Charge Emission Times

Individual trapped charge emission times, for example, τ1,

τ2 and τ3 in Fig. 1, are clearly defined. We collect the first

three emitted charge characteristic times (τi, i = 1, 2, 3). The

emission characteristic times scatter over several decades of time. The probability density functions (PDFs) of the log(τi),

i = 1, 2, 3, are shown in Fig. 3. The mean (log(τi)) and the

standard deviation (σi) of the distributions are indicated in the

figure. The relationship between thelog(τi) (i = 1, 2, 3) is

easily identified. The mean increases by the same amount with a sequence number, i.e.,log(τi+1) − log(τi) ≈ 1.04. On the

other side, the standard deviation σiincreases with i from σ1=

Fig. 2. Magnitude distributions of Δvt,1, Δvt,2and all collected Δvt,ifrom

NBTI recovery traces in 170 high-k/metal gate pMOSFETs.

Fig. 3. Probability density distributions of a trapped charge (hole) emission time in a log(τ ) scale. τ1, τ2and τ3are the 1st, the 2nd and the 3rd trapped

hole emission times, respectively. The mean (log(τi)) and the standard

deviation (σi) of the distributions are indicated in the figure. The symbols are

measurement result and the solid lines are from Monte Carlo simulation. 1.070 to σ3= 1.285, but its dependence on a sequence number

i is more subtle and requires consideration of trapped charge spatial and energetic distributions and its activation energy. We collect all the charge emission times in about 170 devices and plot their occurrence number distribution in a log(τ ) scale in Fig. 4. A rather uniform distribution from 10−2 s to 103 s in Fig. 4 is obtained, implying a log(t) dependence of a recovery ΔVtin a large area device.

(3)

Fig. 4. Histogram of all trapped charge emission times in a log(τ ) scale collected from 170 devices. The occurrence number distribution is rather uniform in a period from 10−2s to 103s.

B. Trapped Charge Spatial and Energy Distributions

According to the ThAT model [19], a trapped charge tunnel emission time in a high-k MOSFET can be expressed as

τi= τ0exp

 Ea kT



exp(αILTIL) exp(αkxi) (1)

αIL= 22m∗ILq(Et+ φB)  (1a) αk=2  2m∗HKqEt  (1b)

where the pre-factor τ0is a lumped parameter, Eais activation energy, TIL is an interfacial layer thickness, xi denotes a trapped charge distance to the HK/IL interface and Et is a trapped charge energy. Other variables have their usual def-initions in [19]. The hole tunneling mass used in this work is m∗IL= 0.41m0 [27] and m∗HK= 0.18m0 [28]. According

to (1), a tunneling front moves in a speed of d = 2.3/αk per decade of time. A removable trapped charge density (Nt) in relaxation therefore can be extracted from the emission occur-rence number versus log(τ ) in Fig. 4 as follows:

Nt=no. of emitted charges/device/decade W× L × d

=αk× (no. of emitted charges/device/decade)

2.3× W × L . (2)

Since the number of emitted charges exhibits a uniform distribution approximately in each decade of time in Fig. 4, we obtain a constant removable trapped charge density Nt in space. With respect to a trapped charge energy distribution, we calculated a voltage drop across an interfacial oxide in NBTI stress by a 2-D numerical device simulation [29]. Quantum corrections are not included in the simulation. The calculated voltage drop is about 0.8 V at a stress Vg of −1.8 V. We

therefore assume that removable trapped holes are uniformly distributed in an energy range of 0 to 0.8 eV above the Si valence band-edge, corresponding to an Etvalue of 2.7 eV to 3.5 eV with respect to the valence band-edge of the HfSiON. This assumption is also adopted in [30] and is supported partly

Fig. 5. Schematic representation of a band diagram of a high-k/metal gate pMOSFET in relaxation. xiand Et represent a trapped charge position and

energy.

by a charge pumping measurement result [31]. Fig. 5 illustrates a removable trapped charge distribution and a band diagram in recovery. The calculated value of αk is from 7.2 to 8.1 nm−1. For simplification, we used an average αk(= 7.65 nm−1) in (2) and obtain a Ntof 1.3× 1018cm−3. An average distance (Δx) between two adjacent trapped charges in the gate-to-substrate direction is about Δx = 1/W LNt∼ 0.32 nm. In addition, the ratio of the emission times of two consecutive emitted trapped charges is

log(τi+1) − log(τi) = 1

2.3× [αk· (xi+1 − xi)]

1

2.3× (αk· Δx) . (3)

Equation (3) shows that the mean of the log(τi) increases

with i by the same amount, i.e., αkΔx/2.3 = 1.06, without regard to activation energy. Equation (3) is consistent with the measurement result in Fig. 3.

C. Activation Energy Distribution

Note that our NBTI emission time model (1), unlike a RTN model, does not have explicit electric field dependence. In contrary, an NBTI emission time model in [17] shows an exponential electric field dependence. The difference is in that a removable trapped hole in our model is assumed in an energy range above the valence band-edge in relaxation (Fig. 5). This assumption is reasonable because of a large voltage drop (0.8 V) across the interfacial oxide in NBTI stress. The mea-surement results in literature [17], [19] also do not support an exponential field dependence in NBTI recovery. Since an electric field is a secondary effect in NBTI recovery, the wide spread of the trapped charge emission times in NBTI recovery is believed mainly due to an activation energy distribution caused by local chemistry.

From (1), Eacan be expressed as

(4)

Fig. 6. Relative activation energy (Ea− Ea) distributions extracted from

the τ1, τ2and τ3, respectively. τ0in (4) is chosen such that theEa is about

0.5 eV. The solid line is a Gaussian-distribution fit.

To extract an Ea distribution from the measured τi, average

values of αk and xi are used in (4) for simplification. The average distance between two consecutive emitted charges is 0.32 nm. The extracted Ea distributions from τ1, τ2, and τ3,

respectively, are shown in Fig. 6. A good match between them is obtained. It should be remarked that the distortion of the Ea distribution of i = 1 is understood because our recovery measurement starts with a time delay of 5 ms. Thus, some emitted charges with very short τ (< 5 ms) are not counted. In Fig. 6, an appropriate τ0 is chosen such that the mean of an

Ea distribution is about 0.5 eV [33]. The solid line in Fig. 6 represents a Gaussian distribution fit with a mean of 0.5 eV and a standard deviation of 0.07 eV. To examine the validity of the Eaextraction, we re-calculate the τ ’s distributions based on an extracted Ea distribution by a Monte Carlo method. In the Monte Carlo procedure, the number of removable trapped charges in each device is selected according to a Poisson distribution [24] with an average number of NtW LTHK, where

THKis the thickness of a high-k dielectric. The use of a Poisson

distribution here is an approximation and its correctness has been discussed in [32]. Then, removable trapped charges are randomly placed in the HK layer. For each trapped charge, an Etis selected in a range from 2.7 eV to 3.5 eV and an Ea is selected according to the distribution in Fig. 6. With a trapped charge location, energy and activation energy, an emission time is calculated according to (1). An emission sequence number is then assigned to each trapped charge according to its calculated emission time. A trapped charge with the shortest τ has i = 1, and the second shortest one has i = 2 and so on. The Monte-Carlo simulated τi (i = 1, 2, 3) distributions (solid lines) are

plotted in Fig. 3. A reasonable agreement between simulation and measurement is obtained. The broadening of the log(τi)

with a sequence number i in Fig. 3 can be partly explained as follows. We re-arrange the terms in (4) and obtain the following equation: log(τi) = 1 2.3  Ea kT + 2.3 log(τ0) + αILTIL+ αkxi  . (5)

As i increases, the tunneling distance xi is larger and the variance of the term (αkxi) in the right hand side increases and so does the variance of log(τi).

Fig. 7. Dependence of1 on recovery temperature. The extracted activation

energy is about 0.53 eV. Each data point is an average of ten readings.

The dependence of a hole emission time on recovery temper-ature is investigated. The tempertemper-ature characterization is some-what difficult because some of NBTI traps might be annealed at high temperatures. For this reason, we chose a single-NBTI trap device. We take an average of τ1 from ten measurements

on the same device by repeatedly re-filling a NBTI trap, i.e., a refilling-recovery-refilling sequence. The result in Fig. 7 shows a linear relationship between a hole emission timelog(τ) and 1/kT, suggesting a thermally- assisted process and Arrhenius activation energy.

III. MODELING OF ARECOVERYΔVtDISTRIBUTION

We measure threshold voltage shifts in a large number of devices at different recovery times. The number of emitted holes and a total threshold voltage shift (ΔVt) in each device

are recorded. Fig. 8 shows the measurement results at a re-covery time of 0.1 sec, 10 sec and 1000 sec. The y-axis is a total ΔVtand the x-axis is the number of emitted holes. Each

data point represents a device. A straight line with a slope of 3.3 mV, i.e., an average single-charge induced Vt shift, is

drawn in the figure as a reference. The measurement results scatter along the line. The ΔVtand the emitted charge number

distributions broaden with recovery time in the measurement period. An average of recovery Vt traces in 170 devices is

plotted in Fig. 9, which reflects an observed recovery charac-teristic in a large-area device. A Monte Carlo ΔVtdistribution

model based on the ThAT and extracted trapped charge spa-tial, energetic and activation energy distributions is developed. The simulation flowchart is shown in Fig. 10. For a recovery time tr, the number of emitted charges (N ) is computed by counting all the charges with τi less than tr. For each emitted charge, a Δvt is randomly selected based on the

dis-tribution f (|Δvt|) = exp(−|Δvt|/σ)/σ with σ = 3.3 mV. A

total ΔVt is then calculated as ΔVt=

N

i=1Δvt,i. In total, 5× 105 devices are simulated. The simulated and measured ΔVtdistributions are shown in Fig. 11 at a recovery time of

tr= 0.1 s, 10 s and 1000 s. Our model is in good agree-ment with measureagree-ment. The mean and the variance of the ΔVt distributions versus recovery time are shown in Figs. 9

(5)

Fig. 8. Total Vt shift (ΔVt) versus number of emitted trapped holes in a

device at a recovery time of 0.1 s, 10 s and 1000 s. Each data point represents a device. A straight line with a slope of 3.3 mV is drawn as a reference.

Fig. 9. Evolution of ΔVtwith a recovery time. The solid line represents an

average of measured ΔVttraces in 170 devices. The symbols are the mean of

Monte Carlo simulated ΔVtdistributions. A logarithmic time dependence of

ΔVtis obtained.

dependence in a large-area device as well as an overall ΔVt

distribution and its temporal evolution in small-area devices. In short, the log(t) behavior is a result of a uniform spatial distri-bution of trapped charges while Eaand Etdistributions affect a ΔVt distribution and its temporal evolution in small-area

devices.

Fig. 10. Simulation flowchart of a Monte Carlo based ΔVtdistribution model

for NBTI relaxation.

Fig. 11. Probability density distributions of NBTI recovery ΔVtin 70 nm×

35 nm pMOSFETs from measurement and from a Monte Carlo simulation. The recovery time is 0.1 s, 10 s, and 1000 s.

Fig. 12. Variance of a recovery ΔVtdistribution versus a recovery time in

70 nm× 35 nm pMOSFETs from measurement and from a Monte Carlo simulation.

(6)

REFERENCES

[1] V. Reddy, A. T. Krishnan, A. Marshall, J. Rodriguez, S. Natarajan, T. Rost, and S. Krishnan, “Impact of negative bias temperature instability on digital circuit reliability,” Microelectron. Reliab., vol. 45, no. 1, pp. 31– 38, Jan. 2005.

[2] M. A. Alam and S. Mahapatra, “A comprehensive model of PMOS NBTI degradation,” Microelectron. Reliab., vol. 45, no. 6, pp. 71–81, Jun. 2005. [3] S. Zafar, Y. H. Kim, V. Narayanan, C. Cabral, Jr., V. Paruchuri, B. Doris, J. Stathis, A. Callegari, and M. Chudzik, “A comparative study of NBTI and PBTI (charge trapping) in SiO2/HfO2 Stacks with FUSI, TiN, Re

Gates,” in VLSI Symp. Tech. Dig., 2006, pp. 23–25.

[4] S. Zafar, A. Kumar, E. Gusev, and E. Cartier, “Threshold voltage insta-bilities in high-k gate dielectric stacks,” IEEE Trans. Device Mater. Rel., vol. 5, no. 1, pp. 45–64, Mar. 2005.

[5] S. Mahapatra, K. Ahmed, D. Varghese, A. E. Islam, G. Gupta, L. Madhav, D. Saha, and M. A. Alam, “On the physical mechanism of NBTI in silicon oxynitride p-MOSFETs: Can difference in insulator processing conditions resolve the interface trap generation versus hole trapping controversy?” in

Proc. Int. Rel. Phys. Symp., 2007, pp. 1–9.

[6] T. Siddiqua and S. Gurumurthi, “Enhancing NBTI recovery in SRAM arrays through recovery boosting,” IEEE Trans. Very Large Scale Integr.

(VLSI) Syst., vol. 20, no. 4, pp. 616–629, Apr. 2012.

[7] J. Shin, V. Zyuban, P. Bose, and T. M. Pinkston, “A proactive wear-out recovery approach for exploiting micro-architectural redundancy to extend cache SRAM lifetime,” in Proc. Int. Symp. Comput. Archit., 2008, pp. 353–362.

[8] H. Kufluoglu, V. Reddy, A. Marshall, J. Krick, T. Ragheb, C. Cirba, A. Krishnan, and C. Chancellor, “An extensive and improved circuit sim-ulation methodology for NBTI recovery,” in Proc. Int. Rel. Phys. Symp., 2010, pp. 670–675.

[9] K. O. Jeppson and C. M. Svensson, “Negative bias stress of MOS devices at high electric fields and degradation of MNOS devices,” J. Appl. Phys., vol. 48, no. 5, pp. 2004–2014, May 1977.

[10] S. Ogawa and N. Shiono, “Generalized diffusion-reaction model for the low-field charge build up instability at the Si-SiO2interface,” Phys. Rev.

B, vol. 51, no. 7, pp. 4218–4230, Feb. 1995.

[11] M. A. Alam, “A critical examination of the mechanics of dynamic NBTI for PMOSFETs,” in IEDM Tech. Dig., 2010, pp. 345–348.

[12] C. T. Chan, C. J. Tang, T. Wang, H. C.-H. Wang, and D. D. Tang, “Positive bias and temperature stress induced two-stage drain current degradation in HfSiON nMOSFETs,” in IEDM Tech. Dig., 2005, pp. 563–566. [13] C. T. Chan, C. J. Tang, T. Wang, H. C. -H. Wang, and D. D. Tang,

“Charac-teristics and physical mechanisms of positive bias and temperature stress-induced drain current degradation in HfSiON nMOSFETs,” IEEE Trans.

Electron Devices, vol. 53, no. 6, pp. 1340–1346, Jun. 2006.

[14] T. Grasser, B. Kaczer, W. Goes, H. Reisinger, T. Aichiger, P. Hehenberger, P.-J. Wagner, F. Schanovky, J. Franco, M. T. Luque, and M. Nelhiebel, “The paradigm shift in understanding the bias temperature instability: From reaction-diffusion to switching oxide traps,” IEEE Trans. Electron

Devices, vol. 58, no. 11, pp. 3652–3665, Nov. 2011.

[15] T. Grasser, B. Kaczer, and W. Goes, “An energy level perspective of bias temperature instability,” in Proc. Int. Rel. Phys. Symp., 2008, pp. 28–38. [16] T. Grasser, B. Kaczer, W. Goes, T. Aichinger, P. Hehenberger, and

M. Nelhiebel, “A two-stage model for negative bias temperature insta-bility,” in Proc. Int. Rel. Phys. Symp., 2009, pp. 33–44.

[17] T. Grasser, H. Reisinger, P.-J. Wagner, W. Goes, F. Schanovsky, and B. Kaczer, “The time dependent defect spectroscopy (TDDS) technique for the bias temperature instability,” in Proc. Int. Rel. Phys. Symp., 2010, pp. 16–25.

relaxation,” IEEE Trans. Electron Devices, vol. 54, no. 9, pp. 2143–2154, Sep. 2007.

[22] S. Mahapatra, A. E. Islam, S. Deora, V. D. Maheta, K. Joshi, and M. A. Alam, “A critical re-evaluation of the usefulness of R-D framework in predicting NBTI stress and recovery,” in Proc. Int. Rel. Phys. Symp., 2011, pp. 614–623.

[23] H. Reisinger, O. Blank, W. Heinrigs, A. Muhlhoff, W. Gustin, and C. Schlunder, “Analysis of NBTI degradation- and recovery-behavior based on ultra-fast Vt-measurements,” in Proc. Int. Rel. Phys. Symp., 2006, pp. 448–453.

[24] B. Kaczer, T. Grasser, P. J. Roussel, J. Franco, R. Degraeve, L.-A. Ragnarsson, E. Simoen, G. Groeseneken, and H. Reisinger, “Origin of NBTI variability in deeply scaled pFRTs,” in Proc. Int. Rel. Phys.

Symp., 2010, pp. 26–32.

[25] M. Denais, A. Bravaix, V. Huard, C. Parthasarathy, G. Ribes, F. Perrier, Y. Rey-Tauriac, and N. Revil, “On-the-fly characterization of NBTI in ultra-thin gate oxide PMOSFETs,” in IEDM Tech. Dig., 2004, pp. 109– 112.

[26] J. P. Chiu, Y. T. Chung, T. Wang, M. C. Chen, C. Y. Lu, and K. F. Yu, “A comparative study of NBTI and RTN amplitude distributions in high-k gate dielectric pMOSFETs,” IEEE Electron Device Lett., vol. 33, no. 2, pp. 176–178, Feb. 2012.

[27] H. Y. Yu, Y. T. Hou, M. F. Li, and D.-L. Kwong, “Hole tunneling current through oxynitride/oxide stack and the stack optimization for p-MOSFETs,” IEEE Electron Device Lett., vol. 23, no. 5, pp. 285–287, May 2002.

[28] Y. T. Hou, M. F. Li, H. Y. Yu, Y. Jin, and D.-L. Kwong, “Quantum tunneling and scalability of HfO2and HfAlO gate stacks,” in IEDM Tech.

Dig., 2002, pp. 731–734.

[29] ISE TCAD Manual Release 10.0, Integr. Syst. Eng., Zurich, Switzerland, 2004.

[30] Z. Q. Teo, D. S. Ang, and G. A. Du, “Observation of two gate stress voltage dependence of NBTI induced threshold voltage shift of ultra-thin oxynitride gate P-MOSFET,” in Proc. Int. Rel. Phys. Symp., 2009, pp. 1002–1004.

[31] Y. Y. Liao, S. F. Horng, Y. W. Chang, T. C. Lu, K. C. Chen, T. Wang, and C. Y. Lu, “Profiling of nitride-trap-energy distribution in SONOS flash memory by using a variable-amplitude low-frequency charge-pumping technique,” IEEE Electron Device Lett., vol. 28, no. 9, pp. 828–830, Sep. 2007.

[32] J. P. Chiu, C. W. Li, and T. Wang, “Characterization and modeling of trap number and creation time distributions under negative-bias-temperature stress,” Appl. Phys. Lett., vol. 101, no. 8, pp. 082906-1–082906-3, Aug. 2012.

[33] M. Toledano-Luque, B. Kaczer, E. Simoen, P. J. Roussel, A. Veloso, T. Grasser, and G. Groeseneken, “Temperature and voltage dependences of the capture and emission times of individual traps in high-k dielectrics,”

Microelectron. Eng., vol. 88, no. 7, pp. 1243–1246, Jul. 2011.

Jung-Piao Chiu received the B.S. degree in

elec-tronics engineering from National Chiao Tung Uni-versity, Taiwan, in 2007, and is currently working toward the Ph.D. degree.

His research interests are reliability analysis in advanced CMOS and Flash memory devices.

(7)

Taiwan, in 2012.

His research interests include electric characteri-zation and reliability study of high-k CMOS devices.

Chi-Wei Li is currently working toward the M.S.

de-gree with the Department of Electronics Engineering and the Institute of Electronics from NCTU, Taiwan. His research interests are reliability analysis in high-k devices and RTN.

Taiwan University and the PhD degree from the University of Illinois, Urbana-Champaign.

Currently, he is a Professor at National Chiao-Tung University, and an Editor of IEEE ELECTRON

數據

Fig. 1. Example ΔI d and V t traces in NBTI relaxation. τ 1 , τ 2 and τ 3 are
Fig. 5. Schematic representation of a band diagram of a high-k/metal gate pMOSFET in relaxation
Fig. 7. Dependence of τ 1  on recovery temperature. The extracted activation
Fig. 9. Evolution of ΔV t with a recovery time. The solid line represents an

參考文獻

相關文件

HPM practice in Taiwan: A case study of HPM Tongxun (HPM Newsletter). These articles have documented the process of development and evolution of HPM practice in Taiwan as well

dimensional nanomaterials for photodetectors with ultrahigh gain and wide spectral response. II.  Photon down conversion and light trapping in hybrid ZnS nanopartcles/Si

• In 1976 Hawking argued that the formation and evaporation of black holes leads to a fundamental loss of information from the universe, a breakdown of predictability, as

HPM practice in Taiwan: A case study of HPM Tongxun (HPM Newsletter). These articles have documented the process of development and evolution of HPM practice in Taiwan as well

Attributable to increasing rent of housing and expenses of house maintenance, rising prices in summer clothing and footwear, as well as fresh vegetables, the indices of Clothing

files Controller Controller Parser Parser.

Y., (1997), “Resource recovery of sludge as building and construction materials — a future trend in sludge management,” Journal of Water Science and Technology, Vol. (2000)

The aim of this study is to investigate students in learning in inequalities with one unknown, as well as to collect corresponding strategies and errors in problem solving..