Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-swr86 Total loading time: 0 Render date: 2024-07-23T19:26:31.800Z Has data issue: false hasContentIssue false

14 - MOS systems on high-mobility channel materials

from Part III - Real MOS systems

Published online by Cambridge University Press:  05 October 2014

Olof Engström
Affiliation:
Chalmers University of Technology, Gothenberg
Get access

Summary

Motivation for high-mobility channel materials in MOSFETs

The progress of information processing relies on advancements in decreasing the switching time of transistors in logic circuits built on CMOS technology. Defining this quantity as the time it takes to move the transistor between its OFF and ON states, there are two device parameters of specific importance: the sub-threshold voltage swing and the transition time for charge carriers along the channel. The first property, discussed in Chapter 11, depends on the capacitive coupling between gate and channel and is connected with the prospect of realizing high-k gate oxides. As we saw in Section 11.4, the second quality depends on the carrier mobility of the channel material and on extrinsic scattering mechanisms originating from the influence of phonons or charge in the gate oxide or from surface roughness (Laux, 2007). For very short channels, below about 20 nm, additional charge interactions are expected for hot carriers and also between channel charge and carriers in the highly doped source and drain areas (Fischetti et al., 2007; Kuhn, 2011). These limitations have motivated the search for new channel materials with higher intrinsic mobility.

The first steps for such improvements were done by creating strain and compressive stress in silicon to change lattice properties, which modifies the semiconductor band structure and gives rise to lower effective mass and thus higher mobility for electrons in n-channel and holes in p-channel devices, the latter built on Ge/Si compounds (Sugii et al., 1999; Fischetti et al., 2002; Leadley et al., 2010). A more radical encroachment into silicon technology has been the introduction of III–V materials for n-channels to speed up carrier transfer. Not surprisingly, the advantage in one parameter earned by such a modification must be offset by a disadvantage in another. Increased mobility is connected with smaller effective masses, which gives lower density of states in the semiconductor energy bands. InGaAs is an interesting candidate in this connection and, as will be further developed below, this compound requires a larger semiconductor surface potential drop in order to achieve a high enough channel charge for acceptable drive current. Its lower density of states demands a larger gate voltage swing for switching the transistor and gives rise to an inferior sub-threshold slope, thus counteracting the improvement in switching time offered by the larger mobility.

Type
Chapter
Information
The MOS System , pp. 333 - 351
Publisher: Cambridge University Press
Print publication year: 2014

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Afanas’ev, V. V. and Stesmans, A. (2004). Energy band alignment at the (100)Ge/HfO2 interface. Appl. Phys. Lett. 84, 2319.CrossRefGoogle Scholar
Afanas’ev, V. V., Stesmans, A., Brammertz, G. et al. (2009). Energy barriers at interfaces between (100) InxGa1-xAs (0 < x < 0.53) and atomic-layer deposited Al2O3 and HfO2. Appl. Phys. Lett. 94, 202110.CrossRefGoogle Scholar
Allen, F. G. and Gobeli, G. W. (1966). Energy structure in photoemission from Cs covered silicon and geranium. Phys. Rev. 144, 558.CrossRefGoogle Scholar
Aspnes, D. E. and Studna, A. A. (1983). Dielectric functions and optical parameters of Si, Ge, GaP, GaAs, GaSb, InP, InAs, and InSb. Phys. Rev B 27, 985.CrossRefGoogle Scholar
Chelikowsky, J. R. and Cohen, M. L. (1976). Nonlocal pseudopotential calculation for the electronic structure of eleven diamond and zinc-blende semiconductors. Phys. Rev. B 14, 556.CrossRefGoogle Scholar
Chelikowsky, J. R., Wagener, T. J., Weaver, J. H. and Jin, A. (1989). Valence- and conduction-band densities of states for tetrahedral semiconductors: Theory and experiment. Phys. Rev. B 40, 9644.CrossRefGoogle ScholarPubMed
Cohen, M. L. and Phillips, J. C. (1965). Spectral analysis of photoemissive yields in Si, Ge, GaAs, GaSb, InAs and InSb. Phys Rev. 139, A912.CrossRefGoogle Scholar
Engström, O. (2012). A model for internal photoemission at high-k oxide/silicon energy barriers. J. Appl. Phys. 112, 064115.CrossRefGoogle Scholar
Engström, O. (2013). Response to “Comment on ‘A model for internal photoemission at high-k oxide/silicon energy barriers’.”. J. Appl. Phys. 113, 166101.CrossRefGoogle Scholar
Engström, O., Przewlocki, H. M., Mitrovic, I. Z. and Hall, S. (2014). Internal photoemission technique for high-k oxide/semiconductor band offset determination: the influence of semiconductor bulk properties. Accepted for publication in Proc. ESSDERC 2014.CrossRef
Fischetti, M. V., Gámiz, F. and Hänsch, W. (2002). On the enhanced electron mobility in strained-silicon inversion layers. J. Appl. Phys. 92, 7320.CrossRefGoogle Scholar
Fischetti, M. V. and Laux, S. E. (1991). Monte Carlo simulation of transport in technologically significant semiconductors of the diamond and zinc-blende structures – Part II: Submicrometer MOSFETs. IEEE Trans. Electron Dev. 38, 650.CrossRefGoogle Scholar
Fischetti, M. V., O’Reagan, T. P., Narayanan, S., Sachs, C., Kim, J. and Zhang, Y. (2007). Theoretical study of some physical aspects of electronic transport in nMOSFETs at the 10 nm gate-length. IEEE Trans. Electron Dev. 54, 2116.CrossRefGoogle Scholar
Heine, V. (1965). Theory of surface states. Phys. Rev. 138, A1689.CrossRefGoogle Scholar
Huang, M. L., Chang, Y. C., Chang, Y. H., Lin, T. D., Kwo, J. and Hong, M. (2009). Energy-band parameters of atomic layer deposited Al2O3 and HfO2 on InxGa1-xAs. Appl. Phys. Lett. 94, 052106.CrossRefGoogle Scholar
Kane, E. O. (1962). Theory for photoemission from semiconductors. Phys. Rev. 127, 131.CrossRefGoogle Scholar
Kane, E. O. (1967). Electron scattering by pair production in silicon. Phys. Rev. 159, 624.CrossRefGoogle Scholar
Kepa, J., Stesmans, A. and Afanas’ev, V. V. (2013). Thermally induced degradation of condensation grown (100)Ge0.75Si0.25/SiO2 interfaces revealed by electron spin resonance. Appl. Phys. Lett. 102, 122104.CrossRefGoogle Scholar
Kuhn, K. J. (2011). Moore’s crystal ball: Device physics and technology past the 15 nm generation. Microel. Eng. 88, 1044.CrossRefGoogle Scholar
Laux, S. E. (2007). A simulation study of the switching times of 22- and 17 nm gate-length SOI nFETs on high mobility substrates and Si. IEEE Tans. Electron Dev. 54, 2304.CrossRefGoogle Scholar
Le Royer, C. (2011). Interfaces and performance: What future for nanoscale Ge and GeSi based CMOS?. Microel. Eng. 88, 1541.CrossRefGoogle Scholar
Leadley, D., Dobbie, A., Myronov, M. et al. (2010). In Balestra, F. (ed.) Nanoscale CMOS. Hoboken, NJ: Wiley, p. 69.Google Scholar
Lee, C. H., Nishimura, T., Saido, N., Nagashio, K., Kita, K. and Toriumi, A. (2009). Record-high mobility in Ge n-MOSFETs exceeding Si universality. Proc. IEDM, p.457.CrossRef
Lee, M. L., Fitzgerald, E. A., Bulsara, M. T., Currie, M. T. and Lochtefeld, A. (2005). Strained Si, SiGe and Ge channels for high-mobility metal-oxide semiconductor field effect transistors. J. Appl. Phys. 97, 011101.CrossRefGoogle Scholar
Mantl, S. and Buca, D. (2010). From thin Si/SiGe buffers to SSOI. In Balestra, F. (ed.) Nanoscale CMOS. Hoboken, NJ: Wiley, p. 127.Google Scholar
Mitrovic, I. Z., Althobaiti, M., Weerakkody, A. D. et al. (2013). Interface engineering of Ge using thulium oxide: Band line-up study. Microel. Eng. 109, 204.CrossRefGoogle Scholar
Mönch, W. (2011). Branch-point energies and the band-structure lineup at Schottky contacts and heterostructures. J. Appl. Phys. 109, 113724.CrossRefGoogle Scholar
Negara, M. A., Goel, N., Bauza, D., Ghibaudo, G. and Hurley, P. K. (2011). Interface state densities, low frequency noise and electron mobility in surface channel In0.53Ga0.47As n-MOSFETs with ZrO2 gate dielectric. Microel. Eng. 88, 1095.CrossRefGoogle Scholar
Nguyen, N. V., Xu, M., Kirillov, O. A. et al. (2010). Band offset of Al2O3/Ga1-xAs (x = 0.53 and 0.75) and the effect of postdeposition annealing. Appl. Phys. Lett. 96, 052107.CrossRefGoogle Scholar
O’Connor, E., Monaghan, S., Cherkaoui, K., Povey, I. M. and Hurley, P. K. (2011). Analysis of the minority carrier response of n-type and p-type Au/NiAl2O3/In0.53Ga0.47As/InP capacitors following an optimized (NH4)2S treatment. Appl. Phys. Lett. 99, 212901.CrossRefGoogle Scholar
O’Reagan, T. P. and Hurley, P. K. (2011). Calculation of the capacitance-voltage characteristic of GaAs, In0.53Ga0.47As and InAs metal-oxide-semiconductor structures. Appl. Phys. Lett. 99, 163502.CrossRefGoogle Scholar
O’Reagan, T. P., Hurley, P. K., Sorée, B. and Fischetti, M. V. (2010). Modelling capacitance-voltage response of In0.47Ga0.53As metal-oxide-semiconductor structures: Charge quantization and nonparabolic corrections. Appl. Phys. Lett. 96, 213514.CrossRefGoogle Scholar
Packan, P., Akbar, S., Armstrong, M. et al. (2009). High performance 32 nm logic technology featuring 2nd generation high-k + metal gate transistors. Proc. IEDM, p. 659.
Porod, W. and Ferry, D. K. (1983). Modification for the virtual-crystal approximation for ternary III-V compounds. Phys. Rev. B 27, 2587.CrossRefGoogle Scholar
Prégaldiny, F., Lallement, C. and Mathiot, D. (2004). Accounting for quantum mechanical effects from accumulation to inversion in a fully analytical surface potential based MOSFET model. Solid-State Electron. 48, 781.CrossRefGoogle Scholar
Przewlocki, H. M., Brzezinska, D. and Engström, O. (2012). Photoemission yield and the electron escape depth determination in metal-oxide-semiconductor structures on N+-type and P+-type silicon substrates. J. Appl. Phys. 111, 114510.CrossRefGoogle Scholar
Robertson, J. (2000) Band offsets of wide-band-gap oxides and implications for future electronic devices. J. Vac. Sci. Technol. B 18, 1786.CrossRefGoogle Scholar
Seguini, G., Perego, M., Spiga, S., Fancuilli, M. and Dimoulas, A. (2007). Conduction band offset of HfO2 on GaAs. Appl. Phys. Lett., 91, 192902.CrossRefGoogle Scholar
Shin, B., Weber, J. R., Long, R. D., Hurley, P. K., Van de Walle, C. G. and McIntyre, P. C. (2010). Origin and passivation of fixed charge in atomic layer deposited aluminum oxide gate insulators on chemically treated InGaAs substrates. Appl. Phys. Lett. 96, 152908.CrossRefGoogle Scholar
Somers, P., Stesmans, A., Souriau, L. and Afanas’ev, V. V. (2012). Electron spin resonance features of the GePb1 dangling bond defect in condensation-grown (100)Si/SiO2/Si1-xGex/SiO2 heterostructures. J. Appl. Phys. 112, 074501.CrossRefGoogle Scholar
Souriau, L., Terzieva, V., Vandervorst, W. et al. (2008). High Ge content SGOI substrates obtained by the Ge condensation technique: A template for growth of strained epitaxial Ge. Thin Solid Films 517, 23.CrossRefGoogle Scholar
Stern, F. and Howard, W. E. (1967). Properties of semiconductor surface inversion layers in the electric quantum limit. Phys. Rev. 163, 816.CrossRefGoogle Scholar
Stesmans, A. and Afanas’ev, V. V. (1998). Electron spin resonance features of interface defects in thermal (100)Si/SiO2. J. Appl. Phys. 83, 2449.CrossRefGoogle Scholar
Stesmans, A., Somers, P. and Afanas’ev, V. V. (2012). Nontrigonal Ge dangling bond interface defect in condensation-grown (100)Si1-xGex/SiO2. Phys. Rev. B 79, 195301.CrossRefGoogle Scholar
Sugii, N., Nakagawa, K., Yamaguchi, S. and Miyao, M. (1999). Role of Si1-xGe buffer on mobility enhancement in a strained-Si n-channel metal-oxide-semiconductor field-effect transistor. Appl. Phys. Lett. 75, 2948.CrossRefGoogle Scholar
Takagi, S., Tezuka, T., Irisawa, T. et al. (2009). Device structures and carrier transport properties of advanced CMOS using high mobility channels. In Deleonibus, S. (ed.) Electronic Device Architectures for the Nano-CMOS Era. Pan Stanford Publishing, p. 81.Google Scholar
Tersoff, J. Theory of semiconductor heterojunctions: The role of quantum dipoles. Phys. Rev. B 30, 4874.
Trinkaus, H., Buca, D., Minamisawa, R. A., Holländer, B., Luyberg, M. and Mantl, S. (2012). Anisotropy of strain relaxation in (100) and (110) Si/SiGe heterostructures. J. Appl. Phys. 111, 014904.CrossRefGoogle Scholar
Walkup, R. E. and Raider, S. I. (1088). In situ measurement of Si(g) production during dry oxidation of crystalline silicon. Appl. Phys. Lett. 53, 888.CrossRefGoogle Scholar
Wang, L., Asbeck, P. M. and Taur, Y. (2010). Self-consistent 1-D Schrödinger-Poisson solver for III-V heterostructures accounting for conduction band non-parabolicity. Solid-State Electron. 54, 1257.CrossRefGoogle Scholar
Weber, O., Damlencourt, J.-F., Andrieu, F. et al. (2006). Fabrication and mobility characteristics of SiGe surface channel pMOSFETs with a HfO2/TiN gate stack. IEEE Trans. Electron Dev. 53, 449.CrossRefGoogle Scholar
Yu, T., Jun, C. G., Zhuge, L. J., Wu, X. M. and Wu, Z. F. (2013). Effect of NH3 plasma treatment in the interfacial property between ultrathin HfO2 and strained Si0.65Ge0.35 substrate. J. Appl. Phys. 113, 044105.CrossRefGoogle Scholar
Zhang, Q., Zhou, G., Xing, H. G. et al. (2012a). Tunnel field effect transistor heterojunction band alignment by internal photoemission spectroscopy. Appl. Phys. Lett. 100, 102104.CrossRefGoogle Scholar
Zhang, R., Iwasaki, T., Taoka, N., Takenaka, M. and Takagi, S. (2012b). Al2O3/GeOx/Ge gate stacks with low interface trap density fabricated by electron cyclotron resonance plasma postoxidation. Appl. Phys. Lett. 98, 112902.CrossRefGoogle Scholar
Zhang, W. F., Nishimula, T., Nagashio, K., Kita, K. and Toriumi, A. (2013). Conduction band offset at GeO2/Ge interface determined by internal photoemission and charge-corrected x-ray photoelectron spectroscopies. Appl. Phys. Lett. 102, 102106.CrossRefGoogle Scholar
Zhao, H., Chen, Y.-T., Yum, J. H., Wang, Y., Zhou, F., Xue, F. and Lee, J. C. (2010). Effects of barrier layers on device performance of high mobility In0.7Ga0.3As metal-oxide-semiconductor field-effect-transistors. Appl. Phys. Lett. 96, 102101.CrossRefGoogle Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×