• 沒有找到結果。

Modeling mechanical stress effect on dopant diffusion in scaled MOSFETs

N/A
N/A
Protected

Academic year: 2021

Share "Modeling mechanical stress effect on dopant diffusion in scaled MOSFETs"

Copied!
9
0
0

加載中.... (立即查看全文)

全文

(1)

Modeling Mechanical Stress Effect on Dopant

Diffusion in Scaled MOSFETs

Yi-Ming Sheu, Sheng-Jier Yang, Chih-Chiang Wang, Chih-Sheng Chang, Li-Ping Huang, Tsung-Yi Huang,

Ming-Jer Chen, Senior Member, IEEE, and Carlos H. Diaz, Senior Member, IEEE

Abstract—The effect of shallow trench isolation mechanical stress on MOSFET dopant diffusion has become significant, and affects device behavior for sub-100-nm technologies. This paper presents a stress-dependent dopant diffusion model and demonstrates its capability to reflect experimental results for a state-of-the-art logic CMOS technology. The proposed stress-de-pendent dopant diffusion model is shown to successfully reproduce device characteristics covering a wide range of active area sizes, gate lengths, and device operating conditions.

Index Terms—Dopant diffusion, mechanical stress, modeling, MOSFET, shallow trench isolation (STI), simulation, strain.

I. INTRODUCTION

A

S A RESULT of the aggressive downscaling of CMOS technology, shallow trench isolation (STI)-induced me-chanical stress increases in magnitude with reduced device active areas, causing a nonnegligible impact on device per-formance [1]–[4]. Both experimental work and numerical simulations have been conducted to calculate the STI stress magnitude and distribution encountered in scaled MOSFETs [5]–[9]. The results show that the silicon stress level near the STI region is high. As design rules or layout dimensions scale down, the high-stress region encroaches further into the MOSFET channel. Thus, STI mechanical stress has a signifi-cant influence on state-of-the-art device performance.

Earlier work studying the mechanical stress effect has been focused on the MOSFET drive current shift, either in the form of localized or planar stress conditions [1]–[3], [6], [10]–[14]. Sev-eral studies have been performed to link STI mechanical stress to mobility changes while accounting for the observed current shift [2], [12], [13], although no threshold voltage shift mech-anism has been investigated. Scott et al. [14] have investigated both the drive current and threshold voltage shift, suggesting a difference in stress-induced diffusivity as the plausible origin of the threshold voltage shift. So far, however, there has been no

Manuscript received June 14, 2004. This work was supported by the National Science Council under Contract NSC 93-2215-E-009-024. The review of this paper was arranged by Editor C.-Y. Lu.

Y.-M. Sheu is with the Department of Advanced Device Technology, Taiwan Semiconductor Manufacturing Company 8, Hsinchu 300, Taiwan, R.O.C. He is also with the Department of Electronics Engineering, National Chiao-Tung University, Hsinchu 300, Taiwan, R.O.C. (e-mail: ymsheu@tsmc.com).

S.-J. Yang, C.-C. Wang, C.-S. Chang, L.-P. Huang, T.-Y. Huang, and C. H. Diaz are with the Department of Advanced Device Technology, Taiwan Semi-conductor Manufacturing Company 8, Hsinchu 300, Taiwan, R.O.C.

M.-J. Chen is with the Department of Electronics Engineering, National Chiao-Tung University, Hsinchu 300, Taiwan, R.O.C.

Digital Object Identifier 10.1109/TED.2004.841286

Fig. 1. Schematic cross section of the device along channel length direction with active area sizeX and gate lengthL both as parameters. The stress condition is compressive mainly because of the lower thermal expansion rate of STI oxide compared to silicon, and the thermal gate oxidation-induced volume expansion at the STI edge.

further elaboration on this aspect. On the other hand, there has been a great deal of work devoted to dopant diffusion behavior in silicon under the influence of mechanical stress [15]–[19]. Cowern et al. [15] proposed a strain-induced dopant diffusivity model of boron diffusion in SiGe. Ahn et al. [17] concluded that in the presence of high-stress nitride film, phosphorus diffu-sion in the silicon was retarded, whereas antimony diffudiffu-sion was enhanced. Aziz [18] established a relationship between hydro-static pressure and biaxial strain via thermodynamic formula-tion, while accommodating calculation of the activation energy shift due to strain. Based on Azizs and Cowerns theoretical work [15], [18], Zangenberg et al. [19] critically reviewed the find-ings over the past ten years and further identified the strain ef-fect on boron and phosphorus diffusion in SiGe. However, most studies in the area of mechanical stress-induced dopant diffu-sion changes remain in fundamental research, i.e., at the silicon material level, and have not yet been extended to semiconductor device characterization and modeling.

It is well recognized that the key MOSFET parameters, such as threshold voltage, drain-induced barrier lowering, body factor, and subthreshold swing, are all strongly dependent on dopant distribution details. Thus, it is crucial to examine stress-dependent dopant diffusion for scaled MOSFETs under mechanical stress.

In this paper, we present a stress-dependent diffusion model and incorporate it into a two-dimensional (2-D) process sim-ulation environment to assess the doping distribution effect in scaled MOSFETs. The proposed model is corroborated by ex-tensive experimental data in a sub-100-nm CMOS technology.

(2)

SHEU et al.: MODELING MECHANICAL STRESS EFFECT ON DOPANT DIFFUSION IN SCALED MOSFETs 31

II. STRESS-DEPENDENT DIFFUSION MODEL ANDMODELINGMETHODOLOGY A. Stress-Dependent Diffusion Model

Several approaches dealing with dopant diffusion dependen-cies on strain are briefly described in the preceding section. The general concept of these approaches is to express dopant diffu-sion under mechanical stress in an Arrhenius form [15], [18], [19]. For example, in the case of a compressively strained SiGe layer where Cowern studied boron diffusion [15], the stress con-dition is regarded as biaxial and the dopant diffusion depen-dence follows the Arrhenius form

(1)

where is the dopant diffusivity under strain, is the dopant diffusivity without strain, is the biaxial strain in the plane of the SiGe layer, and is the activation energy per strain.

Analogous to the above strained SiGe layer, we de-velop a stress-dependent dopant diffusion model for dopant diffusion under STI mechanical stress, named the volume-change-ratio-induced diffusion activation energy shift model (VIDAESM). The volume change ratio is a function of position due to nonuniform stress distributions. In this paper, the MOSFET width is large enough to allow the three-dimen-sional stress effect to be reduced to the 2-D one. The activation energy involved is the product of a dopant-dependent coefficient and volume change ratio, meaning that (1) can be rewritten in the case of dopant diffusion under STI mechanical stress

(2)

where is the dopant diffusivity under strain, is the dopant diffusivity without strain, is the volume change ratio due to stress, is the activation energy per volume change ratio de-pending on the dopant species, and is the temperature. When the strain is small, the volume change ratio can be expressed as

(3) where is the strain along the channel length direction, is the strain in the direction perpendicular to the silicon surface,

is the strain along the channel width direction, and is the strain summation of , , and . Note that is zero in the 2-D simulation due to wide structures adopted. Therefore, (2) becomes

(4)

A 2-D numerical process simulator TSUPREM4 is chosen to perform the process simulation. TSUPREM4 is capable of sim-ulating intrinsic dopant diffusion, three-stream dopant-point

Fig. 2. Flow chart of the modeling procedure.

Fig. 3. I–V calibration result of a short-channel nMOSFET with largeX

for (a)V = 0:05 V and (b) V = 1 V. L = 65 nm and X = 10 m. Symbols stand for the silicon data. Solid lines are the calibrated simulation result.

(3)

Fig. 4. SIMS and calibration results of 1-D dopant profile for (a) nMOSFET and (b) pMOSFET.

effect, the dopant clustering effect, and the dislocation loop effect. For assessment of mechanical stress, the simulator also simultaneously solves force balance equations while taking into account thermal expansion, intrinsic stress, geometry rearrangement after etch and deposition processes, and the thermal oxidation process [20]. The stress-dependent diffusion model VIDAESM has been incorporated into the simulator through the user-specified equation interface to adaptively cal-culate stress-dopant diffusivity during the process simulation. B. Modeling Methodology

To model stress-dependent dopant diffusion for various stress levels, we designed and fabricated a series of MOSFETs with various active area sizes. Fig. 1 schematically shows the cross section view of a test device along the channel direction. The mechanical stress effect was explored here with active area size, , and gate length , both used as the main structural parameters.

The flow chart of the modeling procedure is shown in Fig. 2. First, the one-dimensional (1-D) dopant profiles were processed using blanket control wafers, which covered the range of the process conditions of the device wafers. The results were then taken as stress-free dopant profiles and used to calibrate the dopant diffusion parameters without stress-dependent models.

Second, 2-D MOSFET structures were simulated using the mechanical stress model. Calibrated diffusion parameters were employed to simulate a large MOSFET, where the stress level is low. All front-end major process steps from the STI

Fig. 5. Simulated strain distribution in the silicon of entire front-end process for Lg= 65 nm, (a) X = 10 m, (b) X = 0:6 m. A small X causes a much higher strain.

to the source/drain anneal were considered. The corresponding simulation geometries were calibrated using transmission elec-tron microscope cross-sectional images. Some fine tuning of the 2-D dopant profile parameters, such as implant lateral straggles and segregation factors, are needed to fit the silicon device cur-rent–voltage (I–V) characteristics. Fig. 3 shows the calibration result of a short-channel nMOSFET I–V with a large .

Next, with the stress distribution known, the stress-dependent diffusion models were introduced to simulate MOSFETs with varying values. After implementing the stress-depen-dent diffusion model, process simulation results were used

(4)

SHEU et al.: MODELING MECHANICAL STRESS EFFECT ON DOPANT DIFFUSION IN SCALED MOSFETs 33

Fig. 6. Magnitude of strain versusX corresponding to three pointsA, B, and C in Fig. 4.

Fig. 7. Depletion region boundaries with substrate bias V for 65-nm nMOSFET at (a)V = 0:05 V and (b) V = 1 V.

as device simulation inputs. MEDICI was chosen as the nu-merical device simulator. The device modeling parameters, such as carrier mobility, work function, and silicon/oxide interface charges were calibrated to fit the I–V of large MOSFETs. Then, device simulations with various

values were performed and compared with silicon device data. The above procedure was iterated from process to device cycle until the current-voltage data was all satisfactorily reproduced in all cases. The values from the previous work [15] were employed as the initial guess values. It is worth noting that the numerical convergence and simulation speed were not greatly

Fig. 8. I–V comparison among experimental data, simulation without

stress-dependent diffusion model, and simulation with stress-dependent diffusion model for a smallX MOSFET at (a)V = 0:05 V and (b)V = 1 V. L = 65 nm and X = 0:6 m. Symbols stand for the silicon data. Dashed lines are the simulation without stress-dependent diffusion model. Solid lines are the simulation with stress-dependent diffusion model.

influenced after implementing VIDAESM. The simulation time incorporating VIDAESM increases by about 7% compared to that without VIDAESM.

III. EXPERIMENTAL ANDCOMPARISONS

The silicon wafers were fabricated using novel CMOS cesses. The control wafers for 1-D SIMS analysis were pro-cessed using the same thermal steps as device wafers. Fig. 4 displays SIMS results for both n– and pMOSFET. The implant conditions were BF 2 keV cm and As 2 keV

cm for ultrashallow junction calibration and the junc-tion depths are around 260 angstroms for both devices. The cali-brated simulation profile is also plotted in Fig. 4. The calibration procedure included the fine-tuning of implant damage, dopant-point defect pairing diffusion, silicon-oxide dopant segregation, oxidation enhanced diffusion models, dopant clustering models, dopant-defect clustering models, and intrinsic diffusion models. The stress simulation involved the main process steps, which are STI formation, gate oxidations, and poly-gate for-mation in sequence. Viscoelastic oxidation model was used to simulate the stress-dependent oxide growth. The Youngs moduli used were dyne/cm for the silicon and dyne/cm for the oxide layers. The intrinsic stress used is dyne/cm for the STI oxide and dyne/cm for gate spacer oxide. The other parameters follow the default values in the TSUPREM4 manual [20].

(5)

Fig. 9. Comparison of experimental and simulated nMOSFETV at different I level for various L and X . MinimumX for Lg= 65 nm is 0.6 m, for Lg= 0:18 m is 0.74 m, and for Lg = 0:42 m is 1 m. Final set of dopant diffusion parameters can model MOSFETs of different X under various drain voltages and substrate biases. Symbols stand for silicon data. Solid lines represent simulations with stress-dependent diffusion model. (a)I = 10 Am, X = 10 m; (b) I = 10 Am, X = 10 m; (c) I = 10 Am, X = minimum; (d) I = 10 Am, X = minimum.

The stress distribution results for different values are given in Fig. 5. It can be seen that the polarity of the strain in the lateral direction is negative, meaning that the MOSFET core area experiences a compressive stress. On the other hand, strain in the vertical direction is tensile with a magnitude much smaller than . In particular, Fig. 5(b) reveals that

drastically increases in magnitude with decreasing values. Three reference points , , and are chosen to inspect the value of the strain. is at the center of the gate, is 75 nm away from the gate center and is 150 nm from the gate center. The depth of these points is 20 nm from the silicon surface. Fig. 6 highlights the magnitude of the strain versus the value at points , , and in Fig. 5. The negative polarity of the strain means that the general strain conditions in the active area are compressive, and the magnitude increases rapidly as the value of decreases. The compressive stress mainly comes from lower thermal expansion rate of the STI oxide compared to silicon, and the thermal gate oxida-tion-induced volume expansion at the STI edge. As

decreases, the STI approaches the MOSFET core region and increases the magnitude of compressive stress.

The MOSFET channel width in the silicon experiment was fixed at 10 m, large enough to ensure that the stress along the channel width direction is negligible. Simulations were con-ducted to evaluate the mechanical stress along the channel width direction. The results showed that the average strain level for channel width m is around 0.02%, which is at least two orders of magnitude lower than the peak strain level used in this paper. The MOSFET design set consisted of values

from 0.6 to 10 m and from 65 nm to 0.42 m. It has been recognized that boron and phosphorus diffusion are retarded by compressive strain [15], [19], [21]. The stress simulation re-sults show that the MOSFET channel stress and strain magni-tudes for m are around dyne/cm and 0.04%, respectively. As the value shrinks to 0.6 m, the corresponding stress and strain magnitudes become around dyne/cm and 0.4%, respectively. The compres-sive strain level in the channel region is quite close to the strain produced by 10% germanium in silicon, which falls within the range of Cowern’s and Zangenberg’s studies [15], [19].

In the present paper, the impurities introduced to form nMOSFET are boron, indium, arsenic, and phosphorus, while pMOSFET employed boron and arsenic. Boron, arsenic, and phosphorus were all retarded by STI stress as encountered in fit-ting the silicon MOSFET I–V data. Indium was not considered as a fitting variable because it was observed to be almost immo-bile during the thermal process, meaning that a dopant profile change due to mechanical stress would hardly be observed. As will be mentioned later, the nMOSFETs threshold voltage was observed to increase as the value decreases. The subthreshold I–V with a low drain bias is strongly dependent on the accurate doping profile of the MOSFET shallow core region, which is mainly related to arsenic source/drain exten-sion doping and boron halo doping. In high drain bias cases, the subthreshold I–V depends significantly on the deeper part of the MOSFET doping profile, which is related to phosphorus source/drain, indium halo, and boron halo tail doping profiles. As the gate length varies, the extent of the superposition of

(6)

SHEU et al.: MODELING MECHANICAL STRESS EFFECT ON DOPANT DIFFUSION IN SCALED MOSFETs 35

Fig. 10. Experimental and simulated threshold voltage dependence on X of (a) nMOSFET and (b) pMOSFET. NMOSFET threshold voltage V is more dependent on X than the p-type counterpart. Simulation with stress-dependent diffusion model is able to describe stress-inducedV shift.

tilt-implanted halo doping varies accordingly. Moreover, as the substrate bias increases in magnitude, the depletion region further extends into the substrate from the source, channel and drain regions, considerably influencing subthreshold I–V characteristics. Thus, biasing the MOSFET substrate can serve as a means of verification for the stress-dependent diffusion model. Fig. 7 shows the depletion region boundaries of a 65-nm nMOSFET for low and high drain voltages.

After numerical iterations were completed, the effects of gate lengths, gate voltages, drain voltages, and substrate biases si-multaneously matched with the nMOSFET subthreshold I–V data. This is sufficient to claim that the resulting dopant dis-tributions for the whole device core region are correct. To as-sess the creditability of this model, device I–V simulations with and without VIDAESM were performed and compared. Fig. 8 shows the detailed I–V comparison for a small MOSFET with and without VIDAESM. In the absence of VIDAESM, the simulation fails to correctly describe the I–V dependency on . Fig. 9 displays a series of comparisons with measured gate voltage at different drain current levels for different com-binations of gate lengths, active area sizes, drain voltages, and substrate biases. Remarkably, the extracted diffusion parameter set is able to reproduce all the silicon data well. The broad range of gate lengths, active areas, drain voltages and substrate bi-ases employed in this experiment confirm that the VIDAESM model is indeed suited for modeling the mechanical stress effect on scaled MOSFETs. To further ensure the extracted parameter set also valid for pMOSFETs, threshold voltage dependence on is simulated and compared with silicon data for both nMOSFETs and pMOSFETs. Fig. 10 shows the final results. It can be seen that the nMOSFET threshold voltage increases with

TABLE I

IMPURITY1E VALUESEXTRACTED IN THESTUDY

Fig. 11. Net doping contours for (a)X = 10 m, no stress-dependent model, (b)X = 10 m, with stress-dependent model, (c) X = 0:6 m, no stress-dependent model, and (d) X = 0:6 m, with stress-dependent model. ForX = 0:6 m, the source/drain junction is significantly shallower in the MOSFET core region when the stress-dependent diffusion model is turned on. The gate length is 65 nm.

decreasing values while pMOSFET threshold voltage is relatively insensitive to . The trends for both nMOSFETs and pMOSFETs are adequately described by the extracted pa-rameter set.

Finally, Table I lists the extracted values for all impuri-ties involved. The for phosphorus is 30 eV per volume shift ratio and is largest among the impurities. The for arsenic is 14 eV per volume shift ratio, whereas the for boron is 7 eV. These coefficients confirm diffusion retar-dation by the compressive stress in pure silicon, excluding the Ge chemical effect in strained SiGe experiments. Fig. 11 illus-trates the 2-D contour of the nMOSFET net doping concentra-tion for a gate length of 65 nm. As shown in the figure, for m, the dopant contours with and without the stress-dependent model are comparable, while the source/drain junction for m is significantly shallower and ef-fective gate length is longer in the MOSFET core region when the stress-dependent diffusion model is introduced. To more

(7)

Fig. 12. Dopant profiles of (a) vertical direction forX = 0:6 m nMOSFET, (b) vertical direction for X = 10 m nMOSFET, (c) lateral direction forX = 0:6 m nMOSFET, and (d) lateral direction for X = 0:6 m nMOSFET. The vertical profiles are taken at gate edge and the lateral profiles are taken at 15-nm-deep cutlines of the device. Solid lines are simulation with stress-dependent diffusion model and dashed lines are without stress-dependent diffusion model.X = 0:6 m with stress-dependent model device exhibits significant retardation of dopant diffusion.

clearly visualize the effect of the stress-dependent model, one can inspect the dopant profile along specific cutlines. Fig. 12 displays corresponding vertical and lateral doping profiles for a 65 nm gate length nMOSFET with as a parameter. As can be seen, significant dopant diffusion retardation prevails at small values and this explains an increase in threshold voltage as the decreases.

IV. CONCLUSION

This paper has proposed an accurate stress-dependent dopant diffusion model. The model has been implemented into process and device simulators and has been validated by the exten-sive experimental data. A complete set of MOSFET devices with various gate lengths and active areas have been designed and fabricated from state-of-the-art sub-100-nm process for model verification. Retarded dopant diffusion for phosphorus, boron, and arsenic has been observed and explained by cali-brated dopant profiles while accurately accounting for silicon threshold voltage changes and I–V behaviors. The major benefit of this model is that only a single set of physically based dif-fusion parameters is required to reproduce device subthreshold characteristics for different active areas, gate lengths, drain voltages, and substrate biases. The proposed model therefore can serves as a compact and accurate method for practically dealing with STI mechanical stress-dependent dopant diffusion in ultralarge scale devices.

REFERENCES

[1] Y. M. Sheu, K. Y. Y. Doong, C. H. Lee, M. J. Chen, and C. H. Diaz, “Study on STI mechanical stress-induced variations on advanced CMOSFETs,” in Proc. ICMTS, Mar. 2003, pp. 205–208.

[2] R. A. Bianchi, G. Bouche, and O. Roux-dit-Buisson, “Accurate mod-eling of trench isolation-induced mechanical stress effects on MOSFET electrical performance,” in IEDM Tech. Dig., Dec. 2002, pp. 117–120. [3] K. W. Su, Y. M. Sheu, C. K. Lin, S. J. Yang, W. J. Liang, X. Xi, C. S.

Chiang, J. K. Her, Y. T. Chia, C. H. Diaz, and C. Hu, “A scaleable model for STI mechanical stress effect on layout dependence of MOS electrical characteristics,” in Proc. Custom Integrated Circuits Conf., Sep. 2003, pp. 245–248.

[4] A. Toda, N. Ikarashi, H. Ono, S. Ito, T. Toda, and K. Imai, “Local lattice strain distribution around a transistor channel in metal-oxide-semicon-ductor devices,” Appl. Phys. Lett., vol. 79, pp. 4243–4245, Dec. 2001. [5] D. Chidambarrao, J. P. Peng, and G. R. Srinivasan, “Stresses in

sil-icon substrates near isolation trenches,” J. Appl. Phys., vol. 70, pp. 4816–4822, Nov. 1991.

[6] Y. M. Sheu, C. S. Chang, H. C. Lin, S. S. Lin, C. H. Lee, C. C. Wu, M. J. Chen, and C. H. Diaz, “Impact of STI mechanical stress in highly scaled MOSFETs,” in Proc. Int. Symp. VLSI-TSA, Oct. 2003, pp. 269–272. [7] V. Senez, T. Hoffmann, E. Robiliart, G. Bouche, H. Jaouen, M.

Lunen-borg, and G. Carnevale, “Investigations of stress sensitivity of 0.12 CMOS technology using process modeling,” in IEDM Tech. Dig., Dec. 2001, pp. 831–834.

[8] V. Senez, A. Armigliato, I. De Wolf, G. Carnevale, R. Balboni, S. Frab-boni, and A. Benedetti, “Strain determination in silicon microstructures by combined convergent beam electron diffraction, process simulation, and micro-Rama spectroscopy,” J. Appl. Phys., vol. 94, pp. 5574–5583, Nov. 2003.

[9] B. Dietrich, V. Bukalo, A. Fischer, K. F. Dombrowski, E. Bugiel, B. Kuck, and H. H. Richter, “Raman-spectroscopic determination of inho-mogeneous stress in submicron silicon devices,” Appl. Phys. Lett., vol. 82, pp. 1176–1178, Feb. 2003.

[10] K. Rim, J. L. Hoyt, and J. F. Gibbons, “Fabrication and analysis of deep submicron strained-Si N-MOSFETs,” IEEE Trans. Electron

(8)

SHEU et al.: MODELING MECHANICAL STRESS EFFECT ON DOPANT DIFFUSION IN SCALED MOSFETs 37

[11] S. Ito, H. Namba, K. Yamaguchi, T. Hirata, K. Ando, S. Koyama, S. Kuroki, N. Ikezawa, T. Suzuki, T. Saitoh, and T. Horiuchi, “Mechanical stress effect of etch-stop nitride and its impact on deep submicron tran-sistor design,” in IEDM Tech. Dig., Dec. 2000, pp. 247–250.

[12] A. Lochtefeld and D. A. Antoniadis, “Investigating the relationship between electron mobility and velocity in deeply scaled NMOS via mechanical stress,” IEEE Electron Device Lett., vol. 22, no. 12, pp. 591–593, Dec. 2001.

[13] A. Hamada, T. Furusawa, N. Saito, and E. Takeda, “A new aspect of mechanical stress effects in scaled MOS devices,” IEEE Trans. Electron

Devices, vol. 38, no. 4, pp. 895–900, Apr. 1991.

[14] G. Scott, J. Lutze, M. Rubin, F. Nouri, and M. Manley, “NMOS drive current reduction caused by transistor layout and trench isolation-in-duced stress,” in IEDM Tech. Dig., Dec. 1999, pp. 827–830.

[15] N. E. B. Cowern, P. C. Zlam, P. C. van der Sluis, D. J. Gravesteijn, and W. B. Boer, “Diffusion in strained Si(Ge),” Phys. Rev. Lett., vol. 72, pp. 2585–2588, Apr. 1994.

[16] S. Chaudhry and M. E. Law, “The stress assisted evolution of point and extended defects in silicon,” J. Appl. Phys., vol. 82, pp. 1138–1146, Aug. 1997.

[17] S. T. Ahn, H. W. Kennel, J. D. Plummer, and W. A. Tiller, “Film stress-related vacancy supersaturation in silicon under low-pressure chemical vapor deposited silicon nitride films,” J. Appl. Phys., vol. 64, pp. 4914–4918, Nov. 1988.

[18] M. J. Aziz, “Thermodynamics of diffusion under pressure and stress: relation to point defect mechanisms,” Appl. Phys. Lett, vol. 70, pp. 2810–2812, May 1997.

[19] N. R. Zangenberg, J. Fage-Pedersen, J. L. Hansen, and A. N. Larsen, “Boron and phosphorus diffusion in strained and relaxed Si and SiGe,”

J. Appl. Phys., vol. 94, pp. 3883–3890, Sep. 2003.

[20] Synopsys TSUPREM-4 User Guide, Synopsys Inc., Jun. 2003. [21] H. Park, K. S. Jones, J. A. Slinkman, and M. E. Law, “The effects of

strain on dopant diffusion in silicon,” in IEDM Tech. Dig., Dec. 1993, pp. 303–306.

Yi-Ming Sheu was born in Hsinchu, Taiwan, R.O.C.,

in 1968. He received the B.S. degree in material science and engineering from National Cheng-Kung University, Tainan, Taiwan, in 1991, and the M.S. degree in optical science from National Central University, Zhongli, Taiwan, in 1996. He is currently pursuing the Ph.D. degree in the Department of Electronics Engineering at National Chiao-Tung University, Hsinchu.

In 1997, he joined the Taiwan Semiconductor Manufacturing Company (TSMC) Hsinchu, and is currently engaged in advanced CMOS device design.

Sheng-Jier Yang was born in Taipei, Taiwan,

R.O.C., in 1964. He received the B.S. degree from Tatung Institute of Technology, Tatung, Taiwan, in 1987, the M.S. degree from National Central University, Zhongli, Taiwan, in 1989 and the Ph.D. degree from Arizona State University, Tempe, in 2000, all in chemical engineering.

From 2000 to 2002, he was a Device Engineer with PowerOn Technologies, Inc. Phoenix, where he was working on optimizing discrete power MOSFETs for application voltages ranging from 20 to 800 V. Since 2002, he has been with the Taiwan Semiconductor Manufacturing Company (TSMC) Hsinchu, Taiwan. His current focus is on characterizing stress-induced impact on MOSFETs.

Chih-Chiang Wang received the B.S. degree in

physics science from Fu-Jen University, Xinzhuang, Taiwan, R.O.C., and the M.S. degree in electro-op-tical engineering from National Chiao-Tung University, Hsinchu, Taiwan, in 1989 and 1991, respectively.

In 1991, he joined the Electronics Research and Service Organization (ERSO/ITRI) and worked for the 0.7 and 0.5-m DRAM devices design and 0.25 logic devices development departments. He joined the Taiwan Semiconductor Manufacturing Company (TSMC) Hsinchu, in 1998. Since 2001, he has been the Manager of the TCAD team, Device Engineering Division and is responsible for process and device calibration/simulation. His work primarily focuses on aggressive CMOS technology and the related models development. He has been involved in the technologies development and TCAD calibration for 0.18-, 0.15-, and 0.13-m, and 90-nm technologies, and in stress/SiGe-related modeling and simulation.

Chih-Sheng Chang received the B.S. and M.S.

de-grees in electrical engineering from National Taiwan University, Taipei, Taiwan, R.O.C., in 1986 and 1990, respectively, and the Ph.D. degree from the Depart-ment of Electrical and Computer Engineering, Uni-versity of Illinois, Urbana, in 1996.

Since 1999, he has been with Taiwan Semicon-ductor Manufacturing Company (TSMC), Ltd., Hsinchu, where he has worked on TCAD, MOS device design, and RF devices optimization and characterization. He has been the main device designer for TSMC’s 0.15- and 0.13-mm high-speed devices. Currently, he is a Technology Manager in the Communication Technology Department, Logic Technology Division, where he is in charge of the development of RF active and passive devices for 0.18-mm, 0.13-mm, and 90-nm technologies. His current research interests include active and passive devices for RF CMOS technology, metal gate/high- gate stack for advance CMOS technology, and device physics for strained MOS devices.

Li-Ping Huang received the M.S. degree from

National Tsing Hua University of Applied Physics, Hsinchu, Taiwan, R.O.C., in 1991.

In 2001, he joined the TCAD group of the Taiwan Semiconductor Manufacturing Company (TSMC), Hsinchu. Currently, his interests include process and device simulation.

Tsung-Yi Huang was born in Kaohsiung, Taiwan,

R.O.C., in 1969. He received the B.S. degree in physics science from National Tsing-Hua University (NTHU), Hsinchu, Taiwan in 1992, and the M.S. and Ph.D. degrees in electrical engineering from Tsing-Hua University, Hsinchu, in 1994 and 2001, respectively.

From 1997 to 2002, he worked at Avant! as a TCAD Engineer. In July 2002, he joined the Taiwan Semiconductor Manufacturing Company (TSMC), Hsinchu, as the Technological Manager for device profile modeling. His research includes high-voltage device characteristics modeling, junction engineering, and modeling for advanced CMOS technolo-gies. He is currently developing modeling and simulation for dopant diffusion in strained and strain-relaxed SiGe.

(9)

Ming-Jer Chen (S’78–M’79–SM’98) received the B.S. degree in electrical engineering (with highest honors) in 1977 from National Cheng-Kung University, Tainan, Taiwan, and the Ph.D. degree in electronics engineering in 1985 from National Chiao-Tung University (NCTU), Hsinchu, Taiwan.

In 1985, he joined the faculty of the Department of Electronics Engineering, NCTU, where he became Full Professor in 1993. From 1987 to 1992, he served as a consultant for the Taiwan Semiconductor Manufacturing Company (TSMC), Hsinchu, where he led a team from NCTU and Electronics Research and Service Organization (ERSO/ITRI) to build up a series of process windows and design rules. He has been granted six U.S. patents and six Taiwanese patents covering the high-pre-cision analog capacitors, 1-T SRAM cell, dynamic threshold MOS, and ESD protection. From 2000 to 2001, he was a Visiting Professor at the Department of Electrical Engineering and the Center for Integrated Systems, Stanford University, Stanford, CA. His research interests have long been focused on the technology reliability physics and currently on nanoscale electronics. He has graduated 12 Ph.D. students and more than 80 Master’s students.

Dr. Chen is a member of Phi Tau Phi.

Carlos H. Diaz (M’86–SM’98) received the B.S. and

M.S. degrees in electrical engineering from Univer-sidad de Los Andes, Bogota D.C., Colombia, and the Ph.D. degree in electrical engineering from the Uni-versity of Illinois at Urbana.

He joined Taiwan Semiconductor Manufacturing Company (TSMC)’s Research and Development team, Hsinchu, Taiwan, R.O.C., in 1998 where he is currently Deputy Director, Logic Technology Division. Before joining TSMC, he was a Member of the Technical Staff in the ULSI Laboratory, Hewlett-Packard (HP) Laboratories, Palo Alto, CA, working on process integration, device design, and physics reliability. While at HP, he was also a Lecturer on VLSI technology for University of California, Berkeley extension program. He holds 15 U.S. patents, has published more than 60 papers in technical journals and international conferences, and has coauthored one book. Dr. Diaz was the recipient of two SRC inventor awards. He is TSMC’s repre-sentative to the ITRS PIDS TWG. He served at the Solid State Devices and Materials International Conference, the International Conference on Simula-tion of Semiconductor Processes and Devices, the IEEE InternaSimula-tional Reliability Physics Symposium, and the EOS/ESD Symposium program committees. Cur-rently, he is a member of the VLSI Technology Symposium and the IEEE Inter-national Electron Devices Meeting technical program committees.

數據

Fig. 1. Schematic cross section of the device along channel length direction with active area size X and gate length L both as parameters
Fig. 2. Flow chart of the modeling procedure.
Fig. 5. Simulated strain distribution in the silicon of entire front-end process for Lg = 65 nm, (a) X = 10 m, (b) X = 0:6 m
Fig. 6. Magnitude of strain versus X corresponding to three points A, B, and C in Fig
+4

參考文獻

相關文件

(C)John’s love for graffiti began in junior high s chool when his school held a graffiti contest.. (D)John’s love for crows start s in junior high school when he joined

• Thresholded image gradients are sampled over 16x16 array of locations in scale space. • Create array of

The share of India & Taiwan in the World economy and discussed how world export-import market is increasing year by year.. The data shows us that the business between these

This research applied the modeling approach of Grey relational analysis to establish the relations among the factors, such as service seniority, education, experience,

Therefore, the focus of this research is to study the market structure of the tire companies in Taiwan rubber industry, discuss the issues of manufacturing, marketing and

However, the CRM research was seldom used by the Science Park logistic industry; this research used structural equation modeling (SEM) to research the relationship in CRM,

This research of the installation service of telecom carry on integrating sex valuation, and take the operation of Chunghwa Telecom Hsinchu as an example and inquire in

The bottleneck stations with multiple machines are generally in the industry, such as semiconductor wafer manufacturing plants, IC substrate manufacturing plants,