• 沒有找到結果。

Chapter 4 Strain-induced Change of Gate Direct Tunneling Current

4.2 Physical Model

4.2.3 Electron Direct Tunneling Current for nMOSFETs

4.3.1.4 Uniaxial Transverse Stress

It can be observed that the experimental data measured in Ref. [2], [40] have the same trend and magnitude of the strain-induced change of HDT current under uniaxial longitudinal and transverse stress. This phenomenon can also be expected in our model since the strain-induced valence band edge shifts and the 2D DOS effective masses are equivalent under longitudinal and transverse stress on (001) wafer.

4.3.2 Electron Direct Tunneling Current for nMOSFETs

Fig. 4.10 shows the simulated results for nMOSFETs under uniaxial longitudinal compressive stress. Note that the calculation includes the first two subbands of 2 valleys and the first subband of

Δ Δ 4 valleys since the subband energy of second subband of Δ2 valleys are close to that of the first subband of Δ 4 valleys as shown in Fig4.10(b). The circles are the experimental data measured by four-point bending jig at VG=1V for the device samples consisting of arsenic doped polysilicon gate,

1.3nm nitrided SiO2 gate dielectric, and 1017 cm-3 boron doped p well in Ref. [38].

The diamonds and squares are the data from Ref. [2] and [40] as described in the pMOSFETs case. It can be seen that the simulation results well reproduce the experimental data. The mechanism for the change of EDT current can also be explained by the carrier repopulation. Under uniaxial compressive stress, the lowest subband (Δ2 valleys) shifts up while the second subband (Δ 4 valleys) shifts down as shown in Fig. 4.8(a) and Fig. 4.10(b). Consequently, a number of carriers repopulate from the lowest subband to the second subband, which has lower tunneling barrier. In addition, the barrier of the lowest subband also decreases. Thus, the EDT current increases. Note that this mechanism is reverse to that for pMOSFETs under uniaxial compressive stress. Moreover, the analysis for these two cases can also be applied to explain the change of direct tunneling current under other stress conditions.

Fig. 4.11 shows the simulation results for different substrate doping concentration, 5×1017 cm-3 and 1017 cm-3. It can be seen that the change of EDT current for 1017 cm-3 doping concentration are smaller than that for 5×1017 cm-3. The results are also consistent with the experimental data.

4.4 Conclusion

Using the modified subband energy expression of triangular potential approximation and the WKB approximation, we have demonstrated an efficient and reasonably accurate physical model to calculate HDT and EDT current for longitudinal uniaxial compressive stressed silicon device. The improved one band EMA and the data calculated by six-band k‧p method were used to extract the quantization effective mass. The simulated results correspond to the experimental data,

versus

G

G J

ΔJ / σ , published by former work.

In our model, the subband energy is a function of vertical electric field and stress.

Therefore, this model can be used directly to characterize the - curve and be applied easily to different stress conditions on various wafer orientations with extracting the corresponding physical parameters.

JG VG

Chapter 5

Conclusions

In the chapter 2, the strain tensors have been expressed as a function of normal, longitudinal, and transverse stress on (001), (110), and (111) wafers, respectively.

Then, the strain-altered band structures, band edge shifts, constant energy surface, 2D energy contour, and effective masses for various stress conditions and wafer orientations have been calculated by deformation potential theory and k‧p framework for conduction and valence band, respectively. Utilizing these simulated results as tools to estimate the device performance, we concluded that the best advantageous strains among these stress types and wafer orientations for nMOSFETs have shown to be uniaxial and biaxial tension on (001) wafer while for pMOSFETs they are uniaxial longitudinal compression on both (001) and (110) wafer. Finally, we have examined the influences of additional transverse or normal strain and have found that the additional transverse tensile stress on (001) wafer can further enhance the hole mobility.

In chapter 3, the quantization effective masses, the 2D DOS effective masses, and the 3D DOS effective masses of silicon under uniaxial and biaxial stress on (001) wafer have been extracted. In addition, the strain-altered conduction and valence effective DOS have been derived as a function of the strain-induced band edge shift and strain-altered 3D DOS effective mass. From the calculated results, the effective DOS drops quickly due to the band edge splitting when stress increases from zero to 1GPa, and then tends to a constant determined by 3D DOS effective mass of lowest valley for further increased stress. Furthermore, the Fermi energy of bulk silicon with

non-degenerate doping has also been derived as a function of the strain-induced conduction or valence band edge shifts and the strain-altered effective DOS. The calculated results have shown that the Fermi energy is dominated by band shifts under large stress because the effective DOS tends to constant. Finally, the intrinsic carrier concentration has been derived and expressed as a function of strain-altered effective DOS and strain-induced band gap narrowing. The calculated results have shown that the intrinsic carrier density increases rapidly due to the strain-induced band gap narrowing when the stress is larger than 1GPa.

In Chapter 4, a triangular potential approximation based physical model for HDT current in pMOSFETs and EDT current in nMOSFETs under longitudinal uniaxial compressive stress has been presented. A modified subband energy expression, which comprises the vertical electric field component and stress component, is used to evaluate the subband energy in the inversion layer of MOSFETs under gate bias and stress. Then, an improved one band effective mass approximation and data calculated by six-band k‧p method are used to extract quantization effective mass. Moreover, WKB approximation is utilized to evaluate transmission probability and tunneling current. The simulated results agree with the experimental data published by former works. The primarily reason accounting for the decrease of HDT current for pMOSFETs while increase uniaxial compressive stress is that a number of carriers redistribute from higher subband into the lowest subband due to the stress induced subband energy shift. Since the barrier of the lowest subband for hole tunneling is higher and the corresponding transmission probability is smaller than other subbands, the total tunneling current decreases while stress increases. The reverse mechanism of above can also be used to explain the opposite trend of EDT current for nMOSFETs under uniaxial compressive stress. Thus, the proposed model provides a simple method to assess the influence of external stress for direct tunneling

currents in MOSFETs qualitatively and quantitatively. Moreover, with extracting the corresponding physical parameters, our model can be applied directly to various wafer orientations and different stress conditions. For example, the device with different magnitude and type of stress in poly gate and channel. Alternatively, the longitudinal and transverse stresses exist in the device in the meantime.

References

[1] S. E. Thompson, M. Armstrong, C. Auth, M. Alavi, M. Buehler, R. Chau, S. Cea, T. Ghani, G. Glass, T. Hoffman, C.-H. Jan, C. Kenyon, J. Klaus, K. Kuhn, Z. Ma, B. Mcintyre, K. Mistry, A. Murthy, B. Obradovic, R. Nagisetty, P. Nguyen, S.

Sivakumar, R. Shaheed, L. Shifren, B. Tufts, S. Tyagi, M. Bohr, and Y.

El-Mansy, “A 90-nm logic technology featuring strained-silicon,” IEEE Trans.

Electron Devices, vol. 51, no. 11, pp. 1790-1797, 2004.

[2] S. E. Thompson, G. Sun, Y. S. Choi, and T. Nishida, “Uniaxial-process-induced strained-Si: extending the CMOS roadmap,” IEEE Trans. Electron Devices, vol.

53, no. 5, pp. 1010-1020, 2006.

[3] K. N. Yang, H. T. Huang, M. C. Chang, C. M. Chu, Y. S. Chen, M. J. Chen, Y.

M. Lin, M. C. Yu, M. Jang, C. H. Yu, and M. S. Liang, “A physical model for hole direct tunneling current in p+ poly-gate pMOSFETs with ultrathin gate oxides,” IEEE Trans. Electron Devices, vol. 47, no. 11, pp. 2161-2166, 2000.

[4] Y. T. Hou, M. F. Li, Y. Jin, and W. H. Lai, “Direct tunneling hole currents through ultrathin gate oxides in metal-oxide-semiconductor devices,” J. Appl.

Phys., vol. 91, no. 1, pp.258-264, 2002.

[5] S. H. Lo, D. A. Buchanan, Y. Taur, and W. Wang, “Quantum-mechanical modeling of electron tunneling current from the inversion layer of ultra-thin-oxide nMOSFET’s,” IEEE Electron Device Lett., vol. 18, pp. 209–211, 1997.

[6] X. Yang, J. Lim, G. Sun, K. Wu, T. Nishida, and S. E. Thompson,

“Strain-induced changes in the gate tunneling currents in p-channel metal–oxide–semiconductor field-effect transistors,” Appl. Phys. Lett., vol. 88, pp. 052 108, 2006.

[7] Roy R. Craig, Jr., Mechanics of materials second edition, John Wiley & Sons Inc., 1999.

[8] H. A. Rueda, “Modeling of mechanical stress in silicon isolation technology and its influence on device characteristics,” dissertation of degree of doctor of philosophy, university of Florida, 1999.

[9] Wikipedia (http://en.wikipedia.org/wiki/Main_Page)

[10] Fusahito Yoshida, “Fundamentals of elastic plastic mechanics,” KYORITSU SHUPPAN Co., Ltd., 1997.

[11] Y. Kanda, “A graphical representation of the piezoresistance coefficients in Silicon,” IEEE Trans. Electron Devices, Vol. ED-29, no. 1, pp. 64-70, 1982.

[12] Y. Kanda, “Effect of stress on Germanium and Silicon p-n junctions,” Jpn. J.

Appl. Phys., Vol. 6, No. 4, pp. 475-486, 1967.

[13] MEMS and nanotechnology clearinghouse (http://www.memsnet.org/material/) [14] Almaz Optics, Inc. (http://www.almazoptics.com/material.htm)

[15] KORTH KRISTALLE GMBH (http://www.korth.de/eng/index.htm)

[16] J.-S. Lim, S. E. Thompson, and J. G. Fossum, “Comparison of threshold-voltage shifts for uniaxial and biaxial tensile-stressed n-MOSFETs,” IEEE Electron Device Lett., Vol. 25, no. 11, pp. 731-733, 2004.

[17] C. Herring and E. Vogt, “Transport and deformation-potential theory for many-valley semiconductors with anisotropic scattering,” Phys. Rev., Vol. 101, no. 3, pp. 944-961, 1956.

[18] Y. Sun,S. E. Thompson, and T. Nishida, “Physics of strain effects in semiconductors and metal-oxide-semiconductor field-effect transistors,” J. Appl.

Phys. 101, 104503, 2007.

[19] M. V. Fischetti, Z. Ren, P. M. Solomon, M. Yang, and K. Rim, “Six-band k‧p calculation of the hole mobility in silicon inversion layers: dependence on surface orientation, strain, and silicon thickness,” J. Appl. Phys., vol. 94, no. 2, pp. 1079-1095, 2003.

[20] M. Cardona and F. H. Pollak, “Energy-band structure of germanium and silicon - k.p method,” Phys. Rev., Vol. 142, no. 2, pp. 530-543, 1966.

[21] S. Rodrı´guez, J. A. Lo´pez-Villanueva, I. Melchor, and J. E. Carceller, “Hole confinement and energy subbands in a silicon inversion layer using the effective mass theory,” J. Appl. Phys., Vol. 86, no. 1, pp. 438-444, 1999.

[22] I. Balslev, “Influence of uniaxial stress on the indirect absorption edge in Silicon and Germanium,” Phys. Rev., Vol. 143, no. 2, pp. 636-647, 1966.

[23] M. Lundstrom, Fundamentals of carrier transport second edition, Cambridge, 2000.

[24] N. Mohta and S. E. Thompson, “Mobility enhancement-the next vector to extend Moore’s law,” IEEE CIRCUITS & DEVICES MAGAZINE, pp. 18-23, 2005.

[25] Y. Sun, G. Sun, S. Parthasarathy, S. E. Thompson, “Physics of process induced uniaxially strained Si,” Materials Science and Engineering B, 135, pp. 179–183, 2006.

[26] S.E. Thompson, S. Suthram, Y. Sun, G. Sun, S. Parthasarathy, M. Chu, and T.

Nishida, “Future of strained Si/semiconductors in nanoscale MOSFETs,”

Invited Paper.

[27] T. Ando, A. B. Fowler, and F. Stern, “Electronic properties of two-dimensional systems,” Rev. Mod. Phys., vol. 54, no. 2, pp. 437-670, 1982.

[28] F. Stern, “Self-Consistent Results for n-Type Si Inversion Layers,” Phys. Rev. B,

vol. 5, no. 12, pp. 4891-4899, 1972.

[29] F. Stern, W. E. Howard, “Properties of Semiconductor Surface Inversion Layers in the Electric Quantum Limit,” Phys. Rev., vol. 163, no. 3, pp. 816-835, 1967.

[30] B. L. Anderson and R. L. Anderson, Fundamentals of semiconductor devices first edition, Mc Graw Hill, 2005.

[31] D. A. Neamen, Fundamentals of semiconductor physics and devices first edition, Mc Graw Hill, 2003.

[32] S. Takagi, M. Takayanagi, and A. Toriumi, “Characterization of inversion-layer capacitance of holes in Si MOSFET’s,” IEEE Trans. Electron Devices, vol. 46, no. 7, pp. 1446-1450, 1999.

[33] L. F. Register, E. Rosenbaum, and K. Yang, “Analytic model for direct tunneling current in polycrystalline silicon-gate metal-oxide-semiconductor devices,” Appl. Phys. Lett., vol. 74, pp. 457–459, 1999.

[34] N. Yang, W. K. Henson, J. R. Hauser, and J. J. Wortman, “Modeling study of ultrathin gate oxides using direct tunneling current and capacitance-voltage measurements in MOS devices,” IEEE Trans. Electron Devices, vol. 46, pp.

1464–1471, July 1999.

[35] Y. T. Hou and M. F. Li, “A novel simulation algorithm for Si valence hole quantization of inversion layer in metal-oxide-semiconductor devices,” Jpn.

Appl. Phys., vol. 40, part 2, no. 2B, pp. L144-147, 2001.

[36] Y. T. Hou and M. F. Li, “Hole quantization effects and threshold voltage shift in pMOSFET—assessed by improved one-band effective mass approximation,”

IEEE Trans. Electron Devices, vol. 48, no. 6, pp. 1188-1193, 2001.

[37] Y. T. Hou and M. F. Li, “A simple and efficient model for quantization effects of hole inversion layers in MOS devices,” IEEE Trans. Electron Devices, vol. 48, no. 12, pp. 2893-2898, 2001.

[38] J. S. Lim, X. Yang, T. Nishida, and S. E. Thompson, “Measurement of conduction band deformation potential constants using gate direct tunneling current in n-type metal oxide semiconductor field effect transistors under mechanical stress,” Appl. Phys. Lett., vol. 89, pp. 073 509, 2006.

[39] C. Y. Hsieh and M. J. Chen, “Measurement of channel stress using gate direct tunneling current in uniaxially stressed nMOSFETs,” IEEE Electron Device Lett., vol. 28, no. 9, pp. 818-820, 2007.

[40] X. Yang, Y. Choi, T. Nishida, and S. E Thompson, “Gate direct tunneling currents in uniaxial stressed MOSFETs,” IEDST, pp. 149-152, 2007.

[41] H. H. Mueller and M. J. Schulz, “Simplified method to calculate the band bending and the subband energies in MOS capacitors,” IEEE Trans. Electron Devices, vol. 44, no. 9, pp. 1539-1543, 1997.

[42] Y. T. Hou, M. F. Li, W. H. Lai, and Y. Jin, “Modeling and characterization of direct tunneling hole current through ultrathin gate oxide in p-metal-oxide- semiconductor field-effect transistors” Appl. Phys. Lett., vol. 78, no. 25, pp.

4034-4036, 2001.

[43] P. Liu, G. Lua, X. Liu, Y. Xia, Y. Chen, Y. Zhou, Y. Li, J. Pan, S. Ge,

“Investigation of weak damage in Al0.25Ga0.75As/GaAs by using RBS/C and Raman spectroscopy,” Phys. Lett. A, 286, pp. 332-337, 2001.

Table 2.1 The deformation potentials, Luttinger parameters, elastic stiffness constants, and split-off energy for Si, Ge, and GaAs.

Si Ge GaAs

a a(eV) 2.46 1.24 1.16

b a (eV) -2.1 -2.9 -2.0

d a (eV) -4.8 -5.3 -4.8

γ1a 4.22 13.4 6.98

γ2a 0.39 4.24 2.06

γ3a 1.44 5.69 2.93

S11 (10-12 m2/N) 7.68 b 9.64 b 11.75 c S12 (10-12 m2/N) -2.14 b -2.6 b -3.65 c S44 (10-12 m2/N) 12.6 b 14.9 b 16.8 c

Δ (eV) 0 0.044 d 0.29 d 0.34 d

aSee Ref. 18.

bSee Ref. 12.

cSee Ref. 43.

dSee Ref 30.

Table 2.2 The normal, longitudinal, and transverse direction for (001), (110), and (111) wafer.

wafer orientation normal direction (out of plane)

longitudinal direction (in-plane)

transverse

direction (in-plane) (001) [001] [110] [110]

(110) [110] [110] [001]

(111) [111] [110] [11 2 ]

Table 2.3 The stress tensor and strain tensor for biaxial stress on (001) wafer, uniaxial stress along [110], [110], [001], [111], and [11 2 ] direction.

Biaxial d [110] [110] [001] [111] [11 2 ]

aThe form of stress tensor is defined by Equation (2.2).

bThe form of strain tensor is defined by Equation (2.6).

c

σk indicates the stress applied along k-direction.

dFor biaxial stress on (001) wafer, k is along [100] or [010] and σ[ ]100[ ]010 .

Table 2.4 The resultant strain tensors in response to the combination of normal, longitudinal, and transverse stress for the three wafer orientations

Wafer orientation Strain tensor

Biaxial stress on (001) wafer

Uniaxial stress on (001) wafer

Uniaxial stress on (110) wafer

Uniaxial stress on (111) wafer

Table 2.5 Numerical values of effective mass for silicon conduction band in inversion layer given by [27].

surface (100) (110) (111)

valleys lower higher lower higher all

degeneracy 2 4 4 2 6

Normal mass (m0) 0.916 0.190 0.315 0.190 0.258

Conductivity mass (m0) 0.190 0.315 0.283 0.315 0.296 DOS effective mass (m0) 0.190 0.417 0.324 0.417 0.358

Table 2.6 The conductivity, transverse, and quantization effective masses of the top band for bulk silicon with various stress conditions and wafer orientations. The quantization effective masses of the second band are also listed.

wafer Stress type

mc,1st (m0) mtran,1st

(m0)

mnorm,1st

(m0)

mnorm,2nd

(m0)

σ <0 0.12 1.37 0.28 0.22

Uniaxial

longitudinal σ >0 0.46 0.18 0.21 0.24

σ <0 1.37 0.12 0.28 0.22

Uniaxial

transverse σ >0 0.18 0.46 0.21 0.24

σ <0 0.22 0.22 0.29 0.27

(001)

Biaxial

σ >0 0.28 0.28 0.18 0.29

σ <0 0.12 0.28 1.37 0.15

Uniaxial

longitudinal σ >0 0.46 0.21 0.18 0.17

σ <0 0.28 0.18 0.28 0.22

(110)

Uniaxial

transverse σ >0 0.22 0.29 0.22 0.22

σ <0 0.12 0.17 0.47 0.18

Uniaxial

longitudinal σ >0 0.46 0.29 0.19 0.20

σ <0 0.37 0.23 0.68 0.17

(111)

Uniaxial

transverse σ >0 0.19 0.25 0.18 0.19

Table 2.7 Comparison the effective masses between the 1GPa uniaxial longitudinal compressive stress with and without additional 1GPa uniaxial transverse tensile stress for (001) and (110) wafer.

wafer Additional transverse stress

mc,1st

(m0)

mc,2nd

(m0)

mtran,1st

(m0)

mnorm,1st

(m0)

mnorm,2nd

(m0)

Without 0.12 0.59 1.37 0.28 0.22

(001)

With 0.12 0.3 1.88 0.28 0.18

Without 0.12 0.59 0.28 1.37 0.15

(110)

With 0.12 1.21 0.29 1.32 0.12

Table 4.1. Equations for subband calculation.

Description Equation Oxide electric field

ox ox

ox t

F =V

Potential drop due to poly depletion

Substrate band bending

ox

Silicon surface field

Si

Subband energies Equation (4.1) for pMOSFETs Equation (4.10) for nMOSFETs Inversion carrier density per

subband ⎟⎟

Total inversion layer carrier per

area Ninv =

Nij

Total inversion QM channel

thickness ox ox

ij Average QM channel thickness =

s

Silicon potential drop

q Ionized impurity density per area

q N NdeplSiVdepl sub

=

( ) P Δ R

( ) P

Δ V

( ) P

V

t

Δ

( ) P

V

s

Δ

( ) P

Δ F

Δ A

σ

z

σ

y

σ

x

τ

yx

τ

yz

τ

zy

τ

zx

τ

xy

τ

xz

Fig. 2.1. (a) An arbitrary force ΔR

( )

P acting on an infinitesimal area ΔA at point P.

The normal component of the force is ΔF

( )

P and the tangential components of the force are ΔVs

( )

P and ΔVt

( )

P along two orthogonal directions in the plane. (b) Schematic of the nine components defining the stress state at an arbitrary point in three dimensions.

σ

y

σ

y

2 Δ L 2

Δ L

τ

yz

τ

zy

2 π

θ

θ

y uz

z

uy

Fig. 2.2. (a) Schematic of the deformation of a body applied to normal stress along y-axis; and (b) schematic the deformation of a body applied to pure shear stress. Dash line indicates the size and shape of the original body before deformation and solid line indicates those of the body after deformation

[100]

[010]

] 10 1 [

] 110 [

] 001 [

Fig. 2.3. Schematic of the surface orientation and the corresponding stress directions for (001) wafer. The shadow region indicates the wafer surface. The surface normal is [001], the longitudinal (channel) direction is [110], and the transverse direction, which is perpendicular to the channel in the plane, is [110].

] 110 [

] 10 1 [

] 1 00 [

Fig. 2.4. Schematic of the surface orientation and the corresponding stress directions for (110) wafer. The shadow region indicates the wafer surface. The surface normal is [110], the longitudinal (channel) direction is [110], and the transverse direction, which is perpendicular to the channel in the plane, is [001].

[100]

[010]

[001]

] 10 1 [

] 111 [

] 2 11 [

Fig. 2.5. Schematic of the surface orientation and the corresponding stress directions for (111) wafer. The shadow region indicates the wafer surface. The surface normal is [111], the longitudinal (channel) direction is [110], and the transverse direction, which is perpendicular to the channel in the plane, is [112].

Fig. 2.6. Silicon valence band structures for (a) unstressed, (b) 1GPa uniaxial longitudinal compressive, (c) 1GPa uniaxial longitudinal tensile, (d) 1GPa uniaxial transverse compressive, and (e) 1GPa uniaxial transverse tensile stress on (001) wafer.

Fig. 2.7. Silicon valence band structures for (a) 1GPa biaxial compressive and (b) 1GPa biaxial tensile stress on (001) wafer.

Fig. 2.8. Silicon valence band structures for (a) unstressed, (b) 1GPa uniaxial longitudinal compressive, (c) 1GPa uniaxial longitudinal tensile, (d) 1GPa uniaxial transverse compressive, and (e) 1GPa uniaxial transverse tensile stress on (110) wafer.

Fig. 2.9. Silicon valence band structures for (a) unstressed, (b) 1GPa uniaxial longitudinal compressive, (c) 1GPa uniaxial longitudinal tensile, (d) 1GPa uniaxial transverse compressive, and (e) 1GPa uniaxial transverse tensile stress on (111) wafer.

Fig. 2.10. Strain-induced hole subband energy shift versus (a) uniaxial longitudinal and (b) uniaxial transverse stress on (001) wafer.

Fig. 2.11. Strain-induced hole subband energy shift versus biaxial stress on (001) wafer.

Fig. 2.12. Comparison between the strain-induced hole subband energy shift calculated by 4×4 and 6×6 Hamiltonian for (a) uniaxial longitudinal and (b) biaxial stress on (001) wafer. The solid line indicates the subband energy calculated by the 6×6 Hamiltonian. The dotted line indicates the subband energy calculated by 4 by 4 Hamiltonian.

Fig. 2.13. Strain-induced hole subband energy shift versus (a) uniaxial longitudinal and (b) uniaxial transverse stress on (110) wafer.

Fig. 2.14. Strain-induced subband energy shift versus (a) uniaxial longitudinal and (b) uniaxial transverse stress on (111) wafer.

Top band

Second band

Third band

Fig. 2.15. Hole constant energy surface of unstressed bulk silicon for three lowest bands.

Top band

Second band

Third band

(a) (b) Fig. 2.16. Hole constant energy surface of silicon under 1GPa uniaxial longitudinal (a)

compressive and (b) tensile stress on (001) wafer for three lowest bands.

Top band

Second band

Third band

(a) (b) Fig. 2.17. Hole constant energy surface of silicon under 1GPa uniaxial transverse (a)

compressive and (b) tensile stress on (001) wafer for three lowest bands.

Top band

Second band

Third band

(a) (b) Fig. 2.18. Hole constant energy surface of Silicon under 1GPa biaxial (a) compressive

and (b) tensile stress on (001) wafer for three lowest bands.

Top band

Second band

Third band

(a) (b)

Fig. 2.19. Hole constant energy surface of silicon under 1GPa uniaxial longitudinal (a) compressive and (b) tensile stress on (110) wafer for three lowest bands.

Top band

Second band

Third band

(a) (b)

Fig. 2.20. Hole constant energy surface of silicon under 1GPa uniaxial transverse (a) compressive and (b) tensile stress on (110) wafer for three lowest bands.

Top band

Second band

Third band

(a) (b)

Fig. 2.21. Hole constant energy surface of silicon under 1GPa uniaxial longitudinal (a) compressive and (b) tensile stress on (111) wafer for three lowest bands.

Top band

Second band

Third band

(a) (b)

Fig. 2.22. Hole constant energy surface of silicon under 1GPa uniaxial longitudinal (a) compressive and (b) tensile stress on (111) wafer for three lowest bands.

Top band

Second band

Third band

Fig. 2.23. Contour map in kx, ky plane (kz=0) of unstressed bulk silicon on (001) wafer for three lowest valence bands.

Top band

Second band

Third band

(a) (b) Fig. 2.24. Contour map in kx, ky plane (kz=0) of silicon under 1GPa uniaxial

longitudinal (a) compressive and (b) tensile stress on (001) wafer for three lowest valence bands.

Top band

Second band

Third band

(a) (b) Fig. 2.25. Contour map in kx, ky plane (kz=0) of silicon under 1GPa uniaxial transverse

(a) compressive and (b) tensile stress on (001) wafer for three lowest valence bands.

Top band

Second band

Third band

(a) (b) Fig. 2.26. Contour map in kx, ky plane (kz=0) of silicon under 1GPa biaxial (a)

compressive (b) tensile stress on (001) wafer for three lowest valence bands.

Top band

Second band

Third band

(a) (b) Fig. 2.27. Contour map in kx, ky plane (kz=0) of silicon under 1GPa uniaxial

longitudinal (a) compressive and (b) tensile stress on (110) wafer for three lowest valence bands.

Top band

Second band

Third band

(a) (b) Fig. 2.28. Contour map in kx, ky plane (kz=0) of silicon under 1GPa uniaxial transverse

(a) compressive and (b) tensile stress on (110) wafer for three lowest valence bands.

Top band

Second band

Third band

(a) (b) Fig. 2.29. Contour map in kx, ky plane (kz=0) of silicon under 1GPa uniaxial

longitudinal (a) compressive and (b) tensile stress on (111) wafer for three lowest valence bands.

Top band

Second band

Third band

(a) (b) Fig. 2.30. Contour map in kx, ky plane (kz=0) of silicon under 1GPa uniaxial transverse

(a) compressive and (b) tensile stress on (111) wafer for three lowest valence bands.

Band structure

Constant energy surface

Energy contour in the surface

(a) (b) Fig. 2.31. Band structures, constant energy surface, and energy contour of bulk

silicon under 1GPa longitudinal compressive stress with additional 1GPa transverse tensile stress on (a) (001) and (b) (110) wafer, respectively.

相關文件