Hostname: page-component-77c89778f8-vsgnj Total loading time: 0 Render date: 2024-07-19T14:44:20.787Z Has data issue: false hasContentIssue false

A New SiGeC Vertical MOSFET: Single-device CMOS (SD-CMOS)

Published online by Cambridge University Press:  01 February 2011

Carlos J. Augusto
Affiliation:
Carlos.Augusto@QuantumSemi.com, Quantum Semiconductor LLC, San Jose, California, United States
Lynn Forester
Affiliation:
Lynn.Forester@QuantumSemi.com, Quantum Semiconductor LLC, San Jose, California, United States
Get access

Abstract

A new type of silicon-based Vertical MOSFET concept is presented - Single-Device CMOS (SD-CMOS) - in which the same structure can be operated as NFET or as PFET, depending on the biasing conditions [1]. SD-CMOS offers new possibilities for CMOS integration schemes that are simpler - requiring only 5 masks for the “Front- End” - and less costly to manufacture, than any integration scheme requiring the fabrication of two devices with opposite polarities.

In epitaxially grown Vertical MOSFETs, the source, channel and drain can have atomically sharp interfaces, well controlled doping, and channel length controlled by the epitaxial process rather than by lithography and ion-implantation. With epitaxial growth, it is straightforward to do bandgap engineering by incorporating films such as Si1-xGex, and/or Si1-yCy and/or Si1-x-yGexCy, into any of the aforementioned regions. Suitable band offsets at the source/channel interface [2] can suppress DIBL, which is a key limitation to CMOS scaling. These advantages of Vertical MOSFETs are crucial for future CMOS technology nodes, such as 22nm and below.

SD-CMOS has a metallic drain region with work-function close to the mid-gap energy of the channel material, which can be a homogenous material, a random alloy, or a superlattice. The source region has a very narrow bandgap, achievable with (Si1-yCy)m-(Si1-xGex)n superlattices, whose mid-gap level is aligned with that of the channel region, and with band offsets with the channel region that are nearly symmetric for the conduction and valence bands. The source contact is a metal with a work-function close to the mid-gap level of the source and channel regions, which in turn are aligned with the work-function of the drain. The potential barrier, for electrons and holes from the source contact to the source region, is required to be just a few KT. For operation as NMOS and PMOS with nearly symmetric threshold voltages, the work-function of the gate electrode is also aligned with the mid-gap energy level of the channel.

SD-CMOS is unique due to its band alignments: symmetric band edges, from source to drain, with respect to the mid-gap energy (the “mirror” line). Such configuration can only be obtained in the absence of doping, which if present would immediately break that symmetry. The conduction and valence band edges are required to be asymmetric with respect to a cross-section line crossing the channel through the middle of the gate.

The conduction (valence) band offset between source and channel sets the barrier height for electrons (holes) in the OFF condition for NMOS (PMOS), while applying a voltage at the gate leads to the accumulation of electrons (holes) at the source/channel interface, thereby pushing the Fermi-Level in the source above (below) the conduction (valence) band edge of the channel for the ON condition. Band diagrams and a CMOS fabrication flow for SD-CMOS will be presented. [1] US Patent 6,674,099 [2] US Patent 5,914,504

Type
Research Article
Copyright
Copyright © Materials Research Society 2010

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

1. Augusto, Carlos J. R. P., US Patent No. 6 674 099, (6 January 2004).Google Scholar
2. Augusto, Carlos J. R. P., US Patent No. 7 023 030 (4 April 2006).Google Scholar
3. DESSIS 2D device simulator, ISE Integrated Systems Engineering AG, TCAD release 10.0.Google Scholar
4. Fischetti, M.V., Jin, S., Tang, T.-W., Asbeck, P., Taur, Y., Laux, S. E., Sano, N., IEEE 13th International Workshop on Computational Electronics, 2009, IWCE '09; DOI: 10.1109/IWCE.2009.5091145.Google Scholar
5. Sturm, J. C., Manoharan, H., Lenchyshyn, L. C., Thewalt, M. L. W., Rowell, N. L., Noël, J.-P., Houghton, D. C., Phys. Rev. Lett. 66, pp. 13621365, (1991); DOI: 10.1103/PhysRevLett.66.1362.Google Scholar
6. Eberl, K., Brunner, K. and Winter, W., Thin Solid Films, 294 (1–2), pp. 98104 (1997); DOI: 10.1016/S0040–6090(96)09269–3.Google Scholar
7. Solomon, P. M. and Laux, S. E., IEDM Tech. Dig., 2001, pp. 9598; DOI: 10.1109/IEDM.2001.979425.Google Scholar
8. O'Connor, I., Liu, J., Gaffiot, F., Prégaldiny, F., Lallement, C., Maneux, C., Goguet, J., Frégonèse, S., Zimmer, T., Anghel, L., Dang, T.-T., Leveugle, R., “CNTFET Modeling and Reconfigurable Logic-Circuit Design”, IEEE Trans. Circ. Syst.-I, Vol. 54, No. 11, Nov. 2007, pp. 23652379; DOI: 10.1109/TCSI.2007.907835.Google Scholar
9. Yu, W. J., Kim, U. J., Kang, B. R., Lee, I. H., Lee, E.-H., Lee, Y. H., “Adaptive Logic Circuits with Doping-Free Ambipolar Carbon Nanotube Transistors”, NANO LETTERS, 2009 Vol. 9, No. 4, 14011405; DOI: 10.1021/nl803066v.Google Scholar
10. Jabeur, K., Navarro, D., O'Connor, I., Gaillardon, P. E., Jamaa, M. H. B., Clermidy, F., “Reducing transistor count in clocked standard cells with ambipolar double-gate FETs”, Nanoscale Architectures (NANOARCH), 2010 IEEE/ACM International Symposium on, 17–18 June 2010, pp. 4752; DOI: 10.1109/NANOARCH.2010.5510928.Google Scholar
11. See for example Proceedings of the IEEE, Vol. 98, No. 2, February 2010, “Special Issue on Circuit Technology for ULP”, with an introduction by Reuss, R. H., Fritze, M.; DOI: 10.1109/JPROC.2009.2037210.Google Scholar