• 沒有找到結果。

Chapter 3 Electrical Characteristics of Locally Strained PMOSFETs with Poly-SiGe

3.4 Effects of the Strain

The stress from PE-SiN layer was first examined by probing blanket monitor

sample deposited on Si wafers. We confirmed that the stress is compressive and

increases monotonically with increasing thickness. The stress is around –95 MPa for

100 nm SiN. Figures 3.7(a) and 3.7(b) show the output characteristics of PMOSFETs

with the identical geometry at different temperatures. Saturation current increases

monotonically with increasing capping layer thickness. Figures 3.8 (a) and (b) illustrate

18

the Id-Vg characteristics at different temperatures. The transconductance of the devices

exhibits similar trends, as shown in Figs. 3.9 (a) and (b). It should be noted that, in Fig.

3.10, there is no significant difference of the capacitance between unstrained and

strained channel devices. From these results it is inferred that the holes mobility is

improved by SiN capping layer.

Figs. 3.11 (a) and (b) show the percentage increase of the saturation current of the SiN-capping relative to the control devices as a function of channel lengths at 25°C and

125°C, respectively. The enhancement is less pronounced at high temperature owing to

the enhanced phone scattering rate. The mobility enhancement is more significant as

channel length is scaled down. Subthreshold swing of the devices seems not to be

affected by the induced strain, as shown in Figs. 3.12 (a) and (b), even as operated at a raised temperature (125°C).

Figs. 3.13 (a) and (b) show the threshold voltage as a function of channel length.

Fig. 3.14 illustrates the changes in threshold voltage for devices with SiN layer capping relative to that of control devices. For long channel devices (e.g., L > 2 µm), a nearly

constant shift is observed in Fig. 3.14. This phenomenon could be explained by the

reduction in the interface trap density extracted from charge pumping measurements, as

shown in Fig. 3.15. Since the he precursors for SiN deposition are SiH4 and NH3, a large

19

amount of hydrogen species can be introduced during processing and passivate the

interface states. As also can be seen in the figure, the amount of change in threshold

voltage increases with decreasing channel length, especially as channel length is scaled below 1 µm. Such phenomenon is presumably caused by the splitting of valence band

edges by the compressive strain that increases with decreasing channel length.

20

Chapter 4

Effects of SiN Capping on NBTI of PMOSFETs

4.1 Static and Dynamic NBTI

4.1.1 Brief Review of NBTI Mechanisms

For the aggressive scaling of CMOS technologies, an ultra thin gate oxide is

essential to achieve high drive current under lower power operation. The integrity and

reliability of such a thin gate oxide are therefore crucial for ULSI manufacturing.

Recently, negative-bias-temperature instability (NBTI) has been identified as one of the

major reliability concerns for deep sub-micron PMOSFETs [45]. It was observed that a

large number of interface states and positive fixed charges were generated during

negative-bias-temperature stressing (NBTS), resulting in a negative shift in threshold

voltage showing a power-law dependence on stress time:

b

th At

V = (4-1)

This phenomenon becomes more significant as gate oxide is scaled down, and may

even become the limiting factor for deep sub-micron p-channel devices. The shift in

threshold voltage and degradation in transconductance have been suggested to be due to

the interfacial electrochemical reactions related to the holes from the channel inversion

21

layer. The exponential value of the power law equation is around 0.25, which could be

explained by the diffusion-controlled electrochemical reactions. Based on the t0.25-like

time evolution, a generalized reaction-diffusion model for interfacial charge formation

based on the trivalent silicon and its hydrogen compounds was proposed [46].

In the Svensson model [47], an interface trap is a trivalent silicon atom with an

unsaturated valence electron (dangling bond) at the Si-SiO2 interface. It is denoted by

Si≡ iSi and acts as an active interface trap. A post-metal-anneal in a forming gas

(typically 10% H2 in N2 ambient) is widely used to passivate the interface dangling

bonds, and introduces a lot of hydrogen-terminated trivalent Si bonds which are

electrically inactive at the actual interface. If the terminated hydrogen is released from

the Si≡ −Si H bond by some dissociation mechanism, the remaining interface

trivalent silicon (dangling bond) is restored as an active interface trap. Various

mechanisms have been proposed for the dissociation process.

High electric fields can dissociate the silicon-hydrogen bond, according to the

model [48]

0

3 3

SiSiHSiSii+H , (4-2)

where H0 is a neutral interstitial hydrogen atom or atomic hydrogen. Recent

first-principle calculations show that the positively charged hydrogen or proton H+ is the

22

only stable charge state of hydrogen at the interface, and that H+ reacts directly with the

SiH to form an interface trap, according to the reaction [47]:

3 3 2

SiSiH+H+SiSii+H . (4-3)

The SiH is polarized in this model such that the mobile positive H+ migrates

towards the negatively charged dipole region in the SiH molecule. The H+ atom then

reacts with the H- to form H2, leaving behind a positively charges Si dangling bond.

A different model considers the interaction of SiH with “hot holes” or holes near or

at the Si/SiO2 interface [46]. Dissociation involving holes is given by

3 3

SiSiH+h+SiSii+H+. (4-4)

Fig. 4.1 is a schematic view of reaction-diffusion model for Nit generation.

The fixed charge (Qf) also contributes to threshold voltage shift near the SiO2/Si

interface. Qf is a by-product of trivalent Si defect in the oxide, generated with the

reaction

0

3 3

OSiH+h+OSi++H (4-5)

As discussed in Ref. 46, the interface trap density (Nit) and fixed oxide charge

density (Nf) are shown to increase as:

23

here C and C’ are appropriate constant values, Eox is electric field in the oxide, Tox

is oxide thickness and aging time t. Ogawa et al. [46] found that the generation of fixed

oxide charges is independent of oxide thickness, but is inversely proportional to oxide

thickness for interface trap generation. This suggests that NBTI is worse for thinner

oxide, but this phenomenon is not always observed and highly dependent on the process

conditions.

The stable interface traps are only formed if by-product species, X, diffuses away

from the interface into the oxide bulk.

diffusion

interface bulk

X ⎯⎯⎯⎯→X (4-8)

while “X” could be H-related species.

As proved by Jeppson and Svensson [49], the observed t0.25 behavior of the

interface trap generation suggests the generation process is diffusion controlled. Nit

buildup equals the total number of released H species. Hole-assisted reaction breaks

interfacial SiH bonds, resulting in Nit generation:

n

it N X

N = (D t)S , (4-9)

where Dx is the diffusion coefficient of X in the oxide, time exponent n depends on

the type of H species trapped and released in the oxide bulk [50].

The model that has often been invoked to explain the t0.25 dependence of the trap

24

generation rates is only partially correct [46]. In fact, there may exist six regimes of the

reaction-diffusion model for Nit generation [51] as shown in Fig. 4.2. During the early

stress stage, the generation of interface states and hydrogen species is limited by the

dissociation rate (regime1, Nit ~ t1). Then, the process enters the quasi-equilibrium

regime (regime2, Nit ~ t0) immediately. After some stress time, the transport of

hydrogen species limits the dissociation process (regime3, Nit ~ t0.25). However, the

reaction-controlled regime could directly merge into the diffusion-controlled regime.

The rate of ΔNit changes after H diffusion front reaches the SiO2/poly interface. Either

H absorption into poly ensures faster H removal and higher rate of ΔNit (regime4) or H

reflection from poly would result in ΔNit saturation (regime5). Finally, ΔNit should

eventually saturate when all SiH bonds are broken (regime6).

4.1.2 Dynamic NBTI

The conventional NBTI based on static experimental data. During normal

operations of digital circuits, the applied bias to the gate of PMOSFETs in a CMOS

inverter is switched between “high” and “low” voltages. During “low” phase of

PMOSFETs applied bias, the “electric passivation” effect may effectively reduce the

interface traps generated during the “high” phase. The dynamic NBTI (DNBTI) effect

greatly prolongs the lifetime of PMOSFETs operating in a digital circuit, while the

25

conventional static NBTI measurement underestimates the PMOSFET lifetime.

Furthermore, the DNBTI effect is dependent on temperature and gate oxide thickness

[52]. A physical model is proposed for DNBTI that involves the interaction between

hydrogen and silicon dangling bonds [53]. According to this reaction-diffusion theory,

the Δ Vth is attributed to the creation of interface traps as a consequence of

dissociation of Si-H bonds, and subsequent diffusion of the released hydrogen species

towards the gate electrode. In the recovery process, released hydrogen re-passivates Si

dangling bonds [53]. The ΔVth recovery progresses in accordance with the power law

dependency as follows:

n

V =A-B tth

∆ ⋅ , (4-10)

where B/A ratio indicates the ratio of a recovery reaction coefficient to ΔVth just

after NBT stress. This finding has significant impact on the determination of maximum

operation voltage as well as lifetime for future scaling of CMOS devices. Therefore, it is

critically important to investigate NBTI under such dynamic stress conditions.

26 4.2 Experimental Results and Discussion

4.2.1 Static NBTI Characterization

Figure 4.3 shows the results of NBTI stress performed at 25°C with gate overdrive

bias of –3.9 V (VGO = VG - Vth). As can be seen in Fig. 4.3(a), larger change in threshold

voltage shift, ΔVth, is observed in the device with SiN capping layer. The shift curves

show a fractional power-law dependence on time, and the exponential values with

SiN-capping sample are larger than the control sample. According to the classical

reaction-diffusion model, such dependence implies that NBTI is controlled by the

diffusion. From Fig. 4.3(b), the maximum value of transconductance, Gmmax, degrades

gravely in the device with SiN layer. The increase of interface trap density, ΔNit, for

devices with SiN capping layer are larger than that of control case, as shown in Fig.

4.3(c). The exponential value ofΔNit in power-law relationship is higher for devices

with SiN capping. Similar phenomenon is also observed for the generation of oxide

trapped charges, ΔNot, as shown in Fig. 4.3(d).

The above results clearly indicate that the use of PECVD SiN capping may result

in degraded NBTI. Two plausible origins are postulated to explain the worsen NBTI: (1)

Higher density of Si-H bonds at the oxide/Si interface, since the SiN layer contains a

27

large amount of hydrogen species because of the use of SiH4 and NH3 precursor gases;

(2) Higher strain energy stored in the channel. The energy may help trigger the

electrochemical reactions at the interface. This is evidenced by the higher exponent

value of the power-time dependence for devices with SiN capping.

Figure 4.4 shows the results after NBTI stress at 125°C. The trends are similar to

those occurring at 25°C. The devices with SiN-capping have aggravated NBTI issue. It

should be noted that in Fig. 4.4(a), Fig. 4.4(c) and Fig. 4.4(d), nearly saturation inΔVth,

ΔNit andΔNot are observed for the device with SiN capping when the stress time is

longer than 1000 sec. The NBTI is related to the amount of hydrogen species bonded at

the oxide/Si interface, and ∆Nit should eventually saturate when nearly all Si-H bonds

are broken. The disparity between the two splits is postulated to be caused by the local

strain induced in the SiN capping split. The existence of strain weakens the bond

strength of hydrogen species that in turn accelerates the bond-breaking process. Such

phenomenon is not observed in the control devices due to the lack of high strain in the

channel.

Figures 4.5 ~ 4.7 show ΔVth and ΔNit characteristics as a function of time under

different VGO at 25°C for the three splits of samples. ΔVth andΔNit for all cases

increase with increasing stress voltage. The exponential value of ΔVth and ΔNit in

28

power-law relationship are similar under different stress conditions in each split.

Because of ΔNit ∝ Eox1.5, higher VGO can induce larger amount of Nit and degradation.

Figures 4.8 ~ 4.10 illustrate ΔVth and ΔNit evolution under different VGO aging at

125°C for the three splits of samples. The trend for ΔVth at 125°C is similar to the

results at 25°C. However, the exponential value of ΔNit in power-law relationship

decreases with increasing stress bias for the devices with SiN capping. It can be

concluded that larger amounts of SiH bonds are broken by higher stress bias during

early stress aging. Figures 4.11 and 4.12 illustrate the temperature dependence

(Arrehenius plots) of ΔVth, ΔNit andΔNot under 1000 sec stressing. The slope

(activation energy) is larger for the devices with SiN capping.

4.2.2 DNBTI and AC stress

To simulate the switching operation of the device in the CMOS circuits, the gate

voltage during NBTI stress is switched between negative and positive bias for all splits

as shown in Figures 4.13 ~ 4.15. The condition during stress periods was for 125°C,

VGO = –4 V, while that for passivation periods VG was set at several biases. Other

terminals were grounded during the measurement. As can be seen in Figs. 4.13(a),

4.14(a) and 4.15(a), theΔVth, reductions of VG = +1 V during passivation are more

significant than VG = 0 V. We found that the degradation in the stress periods worsens

29

when higher strain is contained in the channel. The trends in interface trap density shift

are similar among the three splits, as shown in Fig. 4.13(b), Fig. 4.14(b) and Fig.

4.15(b), which have weaker dependence on the recover voltage. BecauseΔNit is

proportionate to ΔVth even at different positive biases during DNBT stressing, as

shown in Fig. 4.16(a), Fig. 4.17(a) and Fig. 4.18(a), it can be concluded that ΔVth is

directly affected by ΔNit during entire dynamic stress period. However, it should be

noted that in Fig. 4.16(b), Fig. 4.17(b) and Fig. 4.18(b), the decreases in ΔVth do not

follow the consistent decrease in ΔS, especially for the case with passivation bias VG =

+1 V. It may be due to the electron trapping related to the positive hydrogen charges in

the gate dielectric according to reactions (4-8). The B/A ratio of DNBTI is shown in

Table. 4.1. The recovery reaction coefficients are similar between strained and

unstrained cases.

Furthermore, the frequency dependences of NBT degradation were measured and

shown in Figs. 4.19 to 4.21 with 50% duty cycle. ΔVth and ΔNit are plotted as a

function of frequency, as shown in Fig.4.17. Both ΔVth and ΔNit are strongly

dependent on frequency for devices with SiN capping layer. The results indicate that the

stress under higher AC frequency shows lessΔVth andΔNit degradation. It is due to the

shortened stress time with increased frequency. It results in parts of broken SiH bonds

30

being recovered before entering the next stress state. Interface trap generation would be

contained by increase AC stress frequency and Δ Vth is suppressed indirectly.

Reduction in NBT degradation for strained-channel devices under AC stress is more

significant than the control case, mainly because the amount of broken SiH bonds with

SiN capping layer is larger than the control sample.

31

Chapter 5 Conclusions

5.1 Conclusions

Using poly-SiGe as gate material to fabricate PMOSFETs has numerous

advantages, including reduced gate depletion effect and improved gate sheet resistance.

Compressive PECVD SiN layer could significantly enhance the drive current of PMOS devices at either room or raised operating temperatures (125 °C) due to increase of hole

mobility, as shown in our experiment. The degrees of mobility enhancement are

enlarged as devices geometry is scaled down. Despite this merit, our results also

indicate that the SiN capping may aggravate the NBTI characteristics. A high amount of

hydrogen species contained in the PE-SiN layer as well as the strain energy stored in the

channel may be the culprits for the worsened reliability. The interface trap density

change and threshold voltage shift recover significantly during passivation periods.

We’ve also observed that the strained channel device is influenced strongly by AC stress

frequency.

32

References

[1] K. Mistry, M. Armstrong, C. Auth, S. Cea, T. Coan, T. Ghani, T. Hoffmann, A.

Murthy, J. Sandford, R. Shaheed, K. Zawadzki, K. Zhang, S. Thompson and M.

Bohr, “Delaying Forever: Uniaxial Strained Silicon Transistors in a 90nm CMOS Technology”, VLSI Symp. Tech. Dig., pp.50, 2004.

[2] R. Khamankar, H. Bu, C. Bowen, S. Chakravarthi, P.R. Chidambaram, M. Bevan, A. Krishnan, H. Niimi, B. Smith, J. Blatchford, B. Hornung, J.P. Lu, P. Nicollian, B. Kirkpatrick, D. Miles, M. Hewson, D. Farber, L. Hall, H. AlShareef, A.

Varghese, A. Gurba, V. Ukraintsev, B. Rathsack, J. DeLoach, J. Tran, C.

Kaneshige, M. Somervell, S. Aur, C. Machala and T. Grider, “An Enhanced 90nm High Performance Technology with Strong Performance Improvements from Stress and Mobility Increase through Simple Process Changes”, VLSI Symp. Tech.

Dig., pp.162, 2004.

[3] J. L. Hoyt, H. M. Nayfeh, S. Eguchi, I. Aberg, G. Xia, T. Drake, E. A. Fitzgerald, and D. A. Antoniadis, “Strained silicon MOSFET technology”, IEDM Tech. Dig., pp.23-26, December 2002.

[4] B. H. Lee, A. Mocuta, S. Bedell, H. Chen, D. Sadana, K. Rim, P. O'Neil, R. Mo, K.

Chan, C. Cabral, C. Lavoie, D. Mocuta, A. Chakravarti, R. M. Mitchell, J.

Mezzapelle, F. Jamin, M. Sendelbach, H. Kermel, M. Gribelyuk, A. Domenicucci, K. A. Jenkins, S. Narasimha, S. H. Ku, M. Ieong, I. Y. Yang, E. Leobandung, P.

Agnello, W. Haensch, and J. Welser, “Performance enhancement on sub-70nm strained silicon SOI MOSFETs on ultra-thin thermally mixed strained silicon/SiGe on insulator (TM-SGOI) substrate with raised S/D”, IEDM Tech. Dig., pp.946-948, December 2002.

33

[5] C. Ge, C. Lin, C. Ko, C. Huang, Y. Huang, B. Chan, B. Perng, C. Sheu, P. Tsai, L.

Yao, C. Wu, T. Lee, C. Chen, C. Wang, S. Lin, Y. Yeo, and C. Hu,

“Process-strained Si (PSS) CMOS technology featuring 3D strain engineering”, IEDM Tech. Dig., pp.73-76, December 2003.

[6] T. Ghani, M. Armstrong, C. Auth, M. Bost, P. Charvat, G. Glass, T. Hoffmann, K.

Johnson, C. Kenyon, J. Klaus, B. McIntyre, K. Mistry, A. Murthy, J. Sandford, M.

Silberstein, S. Sivakumar, P. Smith, K. Zawadzki, S. Thompson, and M. Bohr, “A 90nm high volume manufacturing logic technology featuring novel 45nm gate length strained silicon CMOS transistors”, IEDM Tech. Dig., pp.978-980, December 2003.

[7] C. S. Smith, Phys. Rev., Vol. 94, pp.42-49, 1954.

[8] F. Ootsuka, S. Wakahara, K. Ichinose, A. Honzawa, S. Wada, H. Sato, T. Ando, H.

Ohta, K. Watanabe, and T. Onai, “A highly dense, high-performance 130nm node CMOS technology for large scale system-on-a-chip applications”, IEDM Tech.

Dig., pp.575-578, December 2000.

[9] 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 transistor design”, IEDM Tech.

Dig., pp.247-250, December 2000.

[10] G. Timp, A. Agarwal, F. H. Baumann, T. Boone, M. Buonanno, R. Cirelli, V.

Donnelly, M. Foad, D. Grant, M. Green, H. Gossmann, S. Hillenius, J. Jackson, D.

Jacobson, R. Kleiman, A. Kornblit, F. Klemens, J. T. Lee, W. Mansfield, S.

Moccio, A. Murrell, M. O'Malley, J. Rosamilia, J. Sapjeta, and P. Silverman, “Low leakage, ultra-thin gate oxides for extremely high performance sub-100nm nMOSFETs”, IEDM Tech. Dig., pp.930-932, December 1997.

[11] B. Maiti, P. J. Tobin, V. Misra, R. I. Hegde, K. G. Reid, and C. Gelatos, “High

34

performance 20Å NO oxynitride for gate dielectric in deep sub-quarter micron CMOS technology”, IEDM Tech. Dig., pp.651-654, December 1997.

[12] T. Aoyama, K. Suzuki, H. Tashiro, Y. Tada, and H. Arimoto, “Flat-band voltage shifts in P-MOS devices caused by carrier activation in P+-polycrystalline silicon and boron penetration”, IEDM Tech. Dig., pp.627-630, December 1997.

[13] H. P. Tuinhout, A. H. Montree, J. Schmitz, and P. A. Stolk, “Effects of gate depletion and boron penetration on matching of deep submicron CMOS transistors”, IEDM Tech. Dig., pp.631-634, December 1997.

[14] T. King, J. R. Pfiester, and K. C. Saraswat, “A variable-work-function polycrystalline-Si1-xGex gate material for submicrometer CMOS technologies”, IEEE Electron Device Lett., Vol. 12, pp. 533 - 535, October 1991.

[15] D. A. Antoniadis and J. E. Chung, “Physics and technology of ultra short channel MOSFET devices”, IEDM Tech. Dig., pp.21-24, December 1991.

[16] T. King, J. R. Pfiester, J. D. Shott, J. P. McVittie, and K. C. Saraswat, “A polycrystalline-Si1-xGex-gate CMOS technology”, IEDM Tech. Dig., pp.253-256, December 1990.

[17] V. Z. Li, M. R. Mirabedini, R. T. Kuehn, J. J. Wortman, and M. C. Öztürk, “Single gate 0.15µm CMOS devices fabricated using RTCVD in-situ boron doped Si1-xGex gates”, IEDM Tech. Dig., pp.833-836, December 1997.

[18] C. Salm, D. T. van Veen, D. J. Gravesteijn, J. Holleman,and P. H. Woerlee,

“Diffusion and Electrical Properties of Boron and Arsenic Doped Poly-Si and Poly-GexSi1–x (x ~ 0.3) as Gate Material for Sub-0.25 µm Complementary Metal Oxide Semiconductor Applications”, J. Electrochem. Soc., vol.144, pp.3665–3673, 1997.

[19] T. King, J. P. McVittie, K. C. Saraswat, and J. R. Pfiester, “Electrical properties of heavily doped polycrystalline silicon-germanium films”, IEEE Trans. Electron

35 Devices, Vol. 41, pp.228-232, February 1994.

[20] P. Hellberg, S. Zhang, and C. S. Petersson, “Work function of boron-doped polycrystalline SixGe1-x films”, IEEE Electron Device Lett., Vol. 18, pp.456-458, September 1997.

[21] P. Heremans, R. Bellens, G. Groeseneken, and H. E. Maes, “Consistent model for the hot-carrier degradation in n-channel and p-channel MOSFET's”, IEEE Trans.

[21] P. Heremans, R. Bellens, G. Groeseneken, and H. E. Maes, “Consistent model for the hot-carrier degradation in n-channel and p-channel MOSFET's”, IEEE Trans.

相關文件