Epitaxial graphene growth on silicon carbide

Last updated

Epitaxial graphene growth on silicon carbide (SiC) by thermal decomposition is a method to produce large-scale few-layer graphene (FLG). Graphene is one of the most promising nanomaterials for the future because of its various characteristics, like strong stiffness and high electric and thermal conductivity. Still, reproducible production of Graphene is difficult, thus many different techniques have been developed. The main advantage of epitaxial graphene growth on silicon carbide over other techniques is to obtain graphene layers directly on a semiconducting or semi-insulating substrate which is commercially available. [1] [2]

Contents

History

The thermal decomposition of bulk SiC was first reported in 1965 by Badami. He annealed the SiC in vacuum to around 2180 °C for an hour to obtain a graphite lattice. [3] In 1975, Bommel et al. then achieved to form monolayer graphite on the C-face as well as the Si-face of hexagonal SiC. The experiment was carried out under UHV with a temperature of 800 °C and hints for a graphene structure could be found in LEED patterns and the change in the carbon Auger peak from a carbide character to a graphite character. [4] [5] New insights in the electronic and physical properties of graphene like the Dirac nature of the charge carriers, half-integer quantum Hall effect or the observation of the 2D electron gas behaviour were first measured on multilayer graphene from de Heer et al. at the Georgia Institute of Technology in 2004. [6] [7] Still, the Nobel Prize in Physics ″for groundbreaking experiments regarding the two-dimensional material graphene″ in 2010 was awarded to Andre Geim and Konstantin Novoselov. An official online document of the Royal Swedish Academy of Sciences about this awarding got under fire. Walter de Heer mentions several objections about the work of Geim and Novoselov who apparently have measured on many-layer graphene, also called graphite, which has different electronic and mechanical properties. [8] Emtsev et al. improved the whole procedure in 2009 by annealing the SiC-samples at high temperatures over 1650 °C in an argon environment to obtain morphologically superior graphene. [9]

Process

The underlying process is the desorption of atoms from an annealed surface, in this case a SiC-sample. Due to the fact that the vapor pressure of carbon is negligible compared to the one of silicon, the Si atoms desorb at high temperatures and leave behind the carbon atoms which form graphitic layers, also called few-layer graphene (FLG). Different heating mechanisms like e-beam heating or resistive heating lead to the same result. The heating process takes place in a vacuum to avoid contamination. Approximately three bilayers of SiC are necessary to set free enough carbon atoms needed for the formation of one graphene layer. This number can be calculated out of the molar densities. [10] Today's challenge is to improve this process for industrial fabrication. The FLG obtained so far has a non-uniform thickness distribution which leads to different electronic properties. Because of this, there's a demand for growing uniform large-area FLG with the desired thickness in a reproducible way. Also, the impact of the SiC substrate on the physical properties of FLG is not totally understood yet. [1]

The thermal decomposition process of SiC in high / ultra high vacuum works out well and appears promising for large-scale production of devices on graphene basis. But still, there are some problems that have to be solved. Using this technique, the resulting graphene consists of small grains with varying thickness (30–200 nm). These grains occur due to morphological changes of the SiC surface under high temperatures. On the other side, at relatively low temperatures, poor quality occurs due to the high sublimation rate. [2]

The growth procedure was improved to a more controllable technique by annealing the SiC-samples at high temperatures over 1650 °C in an argon environment. [11] [9] The desorbed silicon atoms from the surface collide with the argon atoms and a few are reflected back to the surface. This leads to a decrease of the Si evaporation rate. [12] Carrying out the experiment under high temperatures further enhances surface diffusion. This leads to a restructuring of the surface which is completed before the formation of the graphene layer. [2] As an additional advantage, the graphene domains are larger in size than in the initial process (3 x 50 μm2) up to 50 x 50 μm2 . [13] [14]

Of course, the technology always undergoes changes to improve the graphene quality. One of them is the so-called confinement controlled sublimation (CCS) method. Here, the SiC sample is placed in a graphite enclosure equipped with a small leak. By controlling the evaporation rate of the silicon through this leak, a regulation of the graphene growth rate is possible. Therefore, high-quality graphene layers are obtained in a near-equilibrium environment. [7] [15] The quality of the graphene can also be controlled by annealing in the presence of an external silicon flux. By using disilane gas, the silicon vapor pressure can be controlled. [16]

Crystallographic orientation between the SiC and graphene layers

SiC is bipolar and therefore the growth can take place on both the SiC(0001) (silicon-terminated) or SiC(0001) (carbon-terminated) faces of 4H-SiC and 6H-SiC wafers. The different faces result in different growth rates and electronic properties.

Silicon-terminated face

On the SiC(0001) face, large-area single crystalline monolayer graphene with a low growth rate can be grown. [7] These graphene layers do have a good reproducibility. In this case, the graphene layer grows not directly on top of the substrate but on a complex structure. [15] This structure is non-conducting, rich of carbon and partially covalently bonded to the underlying SiC substrate and provides, therefore, a template for subsequent graphene growth and works as an electronic ″buffer layer″. This buffer layer forms a non-interacting interface with the graphene layer on top of it. Therefore, the monolayer graphene grown an SiC(0001) is electronically identical to a freestanding monolayer of graphene. [15] Changing the growth parameters such as annealing temperature and time, the number of graphene layers on the SiC(0001) can be controlled . [2] The graphene always maintains its epitaxial relationship with the SiC substrate and the topmost graphene, which originates from the initial buffer layer, is continuous everywhere across the substrate steps and across the boundary between regions with different numbers of graphene layers. [1]

The buffer layer does not exhibit the intrinsic electronic structure of graphene but induces considerable n-doping in the overlying monolayer graphene film. [17] [18] This is a source of electronic scattering and leads therefore to major problems for future electronic device applications based on SiC-supported graphene structures. [19] This buffer layer can be transformed into monolayer graphene by decoupling it from the SiC substrate using an intercalation process.

It is also possible to grow off axis on 6H-SiC(0001) wafers. Ouerghi obtained a perfect uniform graphene monolayer at the terraces by limiting the silicon sublimation rate with N2 and silicon fluxes in UHV at an annealing temperature of 1300 °C. [20]

A growth on the 3C-SiC(111) face is also possible. Therefore, annealing temperatures over 1200 °C are necessary. First, the SiC loses silicon atoms and the top layer rearranges in a SiC structure. A loss of further silicon atoms leads to a new intermediate distorted stage of SiC which matches almost the graphene (2 x 2) structure. Losing the residual silicon atoms, this evolves into graphene. The first four layers of cubic SiC(111) are arranged in the same order as SiC(0001) so the findings are applicable to both structures. [2]

Carbon-terminated face

The growth on the SiC(0001) face is much faster than on the SiC(0001) face . Also the number of layers is higher, around 5 to 100 layers and a polycrystalline nature appear. [10] In early reports, the regions of graphene growths have been described as ″islands″ since they appear on microscopy images as pockets of graphene on the substrate surface. [14] [21] Hite et al. however found out, that these islands are positioned at a lower level than the surrounding surface and referred them as graphene covered basins (GCBs). The suggestion is, that crystallographic defects in the substrate act as nucleation sites for these GCBs. During the growth of the graphene layers, the GCBs coalesce with each. Because of their different possible orientations, sizes and thickness, the resulting graphene film contains misoriented grains with varying thickness. This leads to large oriental disorder. [2] Growing graphene on the carbon-terminated face, every layer is rotated against the previous one with angles between 0° and 30° relative to the substrate. Due to this, the symmetry between the atoms in the unit cell is not broken in multilayers and every layer has the electronic properties of an isolated monolayer of graphene. [2]

Evaluation of number of graphene layers

LEEM images and reflectivity data of graphene on 4H-SiC(0001), showing a number of dips in reflectivity corresponding to the layer number. Adapted from LEEM graphene layer number counting.png
LEEM images and reflectivity data of graphene on 4H-SiC(0001), showing a number of dips in reflectivity corresponding to the layer number. Adapted from

To optimize the growth conditions, it is important to know the number of graphene layers. This number can be determined by using the quantized oscillations of the electron reflectivity. Electrons have a wave character. If they are shot on the graphene surface, they can be reflected either from the graphene surface or from the graphene-SiC interface. The reflected electrons (waves) can interfere with each other. The electron reflectivity itself changes periodically as a function of the incident electron energy and the FLG thickness. For example, thinner FLG provides longer oscillation periods. The most suitable technique for these measurements is the low-energy electron microscopy (LEEM). [1]

A fast method to evaluate the number of layers is using optical microscope in combination with contrast-enhancing techniques. Single-layer graphene domains and substrate terraces can be resolved on the surface of SiC. [23] The method is particularly suitable for quick evaluation of the surface.

Applications

Furthermore, epitaxial graphene on SiC is considered as a potential material for high-end electronics. It is considered to surpass silicon in terms of key parameters like feature size, speed and power consumption and is therefore one of the most promising materials for future applications.

Saturable absorber

Using a two-inch 6H-SiC wafer as substrate, the graphene grown by thermal decomposition can be used to modulate a large energy pulse laser. Because of its saturable properties, the graphene can be used as a passive Q-switcher. [24]

Metrology

The quantum Hall effect in epitaxial graphene can serve as a practical standard for electrical resistance. The potential of epitaxial graphene on SiC for quantum metrology has been shown since 2010, displaying quantum Hall resistance quantization accuracy of three parts per billion in monolayer epitaxial graphene. [25] Over the years precisions of parts-per-trillion in the Hall resistance quantization and giant quantum Hall plateaus have been demonstrated. Developments in encapsulation and doping of epitaxial graphene have led to the commercialisation of epitaxial graphene quantum resistance standards

Hall sensors

Other

The graphene on SiC can be also an ideal platform for structured graphene (transducers, membranes). [2]

Open problems

Limitations in terms of wafer sizes, wafer costs and availability of micromachining processes have to be taken into account when using SiC wafers. [2]

Another problem is directly coupled with the advantage. of growing the graphene directly on a semiconducting or semi-insulating substrate which is commercially available. But there's no perfect method yet to transfer the graphene to other substrates. For this application, epitaxial growth on copper is a promising method. The carbon's solubility into copper is extremely low and therefore mainly surface diffusion and nucleation of carbon atoms are involved. Because of this and the growth kinetics, the graphene thickness is limited to predominantly a monolayer. The big advantage is that the graphene can be grown on Cu foil and subsequently transferred to for example SiO2. [26]

See also

Related Research Articles

<span class="mw-page-title-main">Boron nitride</span> Refractory compound of boron and nitrogen with formula BN

Boron nitride is a thermally and chemically resistant refractory compound of boron and nitrogen with the chemical formula BN. It exists in various crystalline forms that are isoelectronic to a similarly structured carbon lattice. The hexagonal form corresponding to graphite is the most stable and soft among BN polymorphs, and is therefore used as a lubricant and an additive to cosmetic products. The cubic variety analogous to diamond is called c-BN; it is softer than diamond, but its thermal and chemical stability is superior. The rare wurtzite BN modification is similar to lonsdaleite but slightly softer than the cubic form.

<span class="mw-page-title-main">Chemical vapor deposition</span> Method used to apply surface coatings

Chemical vapor deposition (CVD) is a vacuum deposition method used to produce high-quality, and high-performance, solid materials. The process is often used in the semiconductor industry to produce thin films.

<span class="mw-page-title-main">Silicon carbide</span> Extremely hard semiconductor

Silicon carbide (SiC), also known as carborundum, is a hard chemical compound containing silicon and carbon. A semiconductor, it occurs in nature as the extremely rare mineral moissanite, but has been mass-produced as a powder and crystal since 1893 for use as an abrasive. Grains of silicon carbide can be bonded together by sintering to form very hard ceramics that are widely used in applications requiring high endurance, such as car brakes, car clutches and ceramic plates in bulletproof vests. Large single crystals of silicon carbide can be grown by the Lely method and they can be cut into gems known as synthetic moissanite.

<span class="mw-page-title-main">Epitaxy</span> Crystal growth process relative to the substrate

Epitaxy refers to a type of crystal growth or material deposition in which new crystalline layers are formed with one or more well-defined orientations with respect to the crystalline seed layer. The deposited crystalline film is called an epitaxial film or epitaxial layer. The relative orientation(s) of the epitaxial layer to the seed layer is defined in terms of the orientation of the crystal lattice of each material. For most epitaxial growths, the new layer is usually crystalline and each crystallographic domain of the overlayer must have a well-defined orientation relative to the substrate crystal structure. Epitaxy can involve single-crystal structures, although grain-to-grain epitaxy has been observed in granular films. For most technological applications, single-domain epitaxy, which is the growth of an overlayer crystal with one well-defined orientation with respect to the substrate crystal, is preferred. Epitaxy can also play an important role while growing superlattice structures.

<span class="mw-page-title-main">Molecular-beam epitaxy</span> Crystal growth process

Molecular-beam epitaxy (MBE) is an epitaxy method for thin-film deposition of single crystals. MBE is widely used in the manufacture of semiconductor devices, including transistors, and it is considered one of the fundamental tools for the development of nanotechnologies. MBE is used to fabricate diodes and MOSFETs at microwave frequencies, and to manufacture the lasers used to read optical discs.

<span class="mw-page-title-main">Graphene</span> Hexagonal lattice made of carbon atoms

Graphene is an allotrope of carbon consisting of a single layer of atoms arranged in a hexagonal lattice nanostructure. The name is derived from "graphite" and the suffix -ene, reflecting the fact that the graphite allotrope of carbon contains numerous double bonds.

<span class="mw-page-title-main">Dangling bond</span>

In chemistry, a dangling bond is an unsatisfied valence on an immobilized atom. An atom with a dangling bond is also referred to as an immobilized free radical or an immobilized radical, a reference to its structural and chemical similarity to a free radical.

A wetting layer is an monolayer of atoms that is epitaxially grown on a flat surface. The atoms forming the wetting layer can be semimetallic elements/compounds or metallic alloys. Wetting layers form when depositing a lattice-mismatched material on a crystalline substrate. This article refers to the wetting layer connected to the growth of self-assembled quantum dots. These quantum dots form on top of the wetting layer. The wetting layer can influence the states of the quantum dot for applications in quantum information processing and quantum computation.

Stranski–Krastanov growth is one of the three primary modes by which thin films grow epitaxially at a crystal surface or interface. Also known as 'layer-plus-island growth', the SK mode follows a two step process: initially, complete films of adsorbates, up to several monolayers thick, grow in a layer-by-layer fashion on a crystal substrate. Beyond a critical layer thickness, which depends on strain and the chemical potential of the deposited film, growth continues through the nucleation and coalescence of adsorbate 'islands'. This growth mechanism was first noted by Ivan Stranski and Lyubomir Krastanov in 1938. It wasn't until 1958 however, in a seminal work by Ernst Bauer published in Zeitschrift für Kristallographie, that the SK, Volmer–Weber, and Frank–van der Merwe mechanisms were systematically classified as the primary thin-film growth processes. Since then, SK growth has been the subject of intense investigation, not only to better understand the complex thermodynamics and kinetics at the core of thin-film formation, but also as a route to fabricating novel nanostructures for application in the microelectronics industry.

Carbide-derived carbon (CDC), also known as tunable nanoporous carbon, is the common term for carbon materials derived from carbide precursors, such as binary (e.g. SiC, TiC), or ternary carbides, also known as MAX phases (e.g., Ti2AlC, Ti3SiC2). CDCs have also been derived from polymer-derived ceramics such as Si-O-C or Ti-C, and carbonitrides, such as Si-N-C. CDCs can occur in various structures, ranging from amorphous to crystalline carbon, from sp2- to sp3-bonded, and from highly porous to fully dense. Among others, the following carbon structures have been derived from carbide precursors: micro- and mesoporous carbon, amorphous carbon, carbon nanotubes, onion-like carbon, nanocrystalline diamond, graphene, and graphite. Among carbon materials, microporous CDCs exhibit some of the highest reported specific surface areas (up to more than 3000 m2/g). By varying the type of the precursor and the CDC synthesis conditions, microporous and mesoporous structures with controllable average pore size and pore size distributions can be produced. Depending on the precursor and the synthesis conditions, the average pore size control can be applied at sub-Angstrom accuracy. This ability to precisely tune the size and shapes of pores makes CDCs attractive for selective sorption and storage of liquids and gases (e.g., hydrogen, methane, CO2) and the high electric conductivity and electrochemical stability allows these structures to be effectively implemented in electrical energy storage and capacitive water desalinization.

<span class="mw-page-title-main">Walter de Heer</span> Dutch physicist

Walter Alexander "Walt" de Heer is a Dutch physicist and nanoscience researcher known for discoveries in the electronic shell structure of metal clusters, magnetism in transition metal clusters, field emission and ballistic conduction in carbon nanotubes, and graphene-based electronics.

<span class="mw-page-title-main">Silicene</span> Two-dimensional allotrope of silicon

Silicene is a two-dimensional allotrope of silicon, with a hexagonal honeycomb structure similar to that of graphene. Contrary to graphene, silicene is not flat, but has a periodically buckled topology; the coupling between layers in silicene is much stronger than in multilayered graphene; and the oxidized form of silicene, 2D silica, has a very different chemical structure from graphene oxide.

Bilayer graphene is a material consisting of two layers of graphene. One of the first reports of bilayer graphene was in the seminal 2004 Science paper by Geim and colleagues, in which they described devices "which contained just one, two, or three atomic layers"

In materials science, the term single-layer materials or 2D materials refers to crystalline solids consisting of a single layer of atoms. These materials are promising for some applications but remain the focus of research. Single-layer materials derived from single elements generally carry the -ene suffix in their names, e.g. graphene. Single-layer materials that are compounds of two or more elements have -ane or -ide suffixes. 2D materials can generally be categorized as either 2D allotropes of various elements or as compounds.

A rapidly increasing list of graphene production techniques have been developed to enable graphene's use in commercial applications.

<span class="mw-page-title-main">Electronic properties of graphene</span>

Graphene is a semimetal whose conduction and valence bands meet at the Dirac points, which are six locations in momentum space, the vertices of its hexagonal Brillouin zone, divided into two non-equivalent sets of three points. The two sets are labeled K and K'. The sets give graphene a valley degeneracy of gv = 2. By contrast, for traditional semiconductors the primary point of interest is generally Γ, where momentum is zero. Four electronic properties separate it from other condensed matter systems.

<span class="mw-page-title-main">Discovery of graphene</span>

Single-layer graphene was first unambiguously produced and identified in 2004, by the group of Andre Geim and Konstantin Novoselov, though they credit Hanns-Peter Boehm and his co-workers for the experimental discovery of graphene in 1962; while it had been explored theoretically by P. R. Wallace in 1947. Boehm et al. introduced the term graphene in 1986.

Two dimensional hexagonal boron nitride is a material of comparable structure to graphene with potential applications in e.g. photonics., fuel cells and as a substrate for two-dimensional heterostructures. 2D h-BN is isostructural to graphene, but where graphene is conductive, 2D h-BN is a wide-gap insulator.

<span class="mw-page-title-main">Low-energy plasma-enhanced chemical vapor deposition</span>

Low-energy plasma-enhanced chemical vapor deposition (LEPECVD) is a plasma-enhanced chemical vapor deposition technique used for the epitaxial deposition of thin semiconductor films. A remote low energy, high density DC argon plasma is employed to efficiently decompose the gas phase precursors while leaving the epitaxial layer undamaged, resulting in high quality epilayers and high deposition rates.

Francesca Iacopi is an engineer, researcher and an academic. She specializes in materials and nanoelectronics engineering and is a professor at the University of Technology Sydney. She is a chief investigator of the ARC Centre of Excellence in Transformative Meta-Optical Systems, a Fellow of the Institution of Engineers Australia, and a senior member of Institute of Electrical and Electronics Engineers.

References

  1. 1 2 3 4 Hibino, Hiroki; Kageshima, Hiroyuki; Nagase, Masao (2010). "Graphene Growth on Silicon Carbide". NTT Technical Review. 8 (8).
  2. 1 2 3 4 5 6 7 8 9 Mishra, Neeraj; Boeckl, John; Motta, Nunzio; Iacopi, Francesca (2016). "Graphene growth on silicon carbide: A review". Physica Status Solidi A. 213 (9).
  3. Badami, D.V. (1965). "X-Ray studies of graphite formed by decomposing silicon carbide". Carbon. 3 (1): 53–57. doi:10.1016/0008-6223(65)90027-8.
  4. Van Bommel, A.J.; Crombeen, J.E.; Van Tooren, A. (1975). "LEED and Auger electron observations of the SiC(0001) surface". Surface Science. 48 (2): 463–472. doi:10.1016/0039-6028(75)90419-7.
  5. Hass, J.; de Heer, W. A.; Conrad, E. H. (2008). "The growth and morphology of epitaxial multilayer graphene". Journal of Physics: Condensed Matter. 20 (32): 323202. doi:10.1088/0953-8984/20/32/323202.
  6. Berger, Claire; Song, Zhimin; Li, Xuebin; Ugbazghi, Asmerom Y.; Feng, Rui; Dai, Zhenting; Marchenkov, Alexei N.; Conrad, Edward H.; First, Phillip N.; de Heer, Walt A. (2004). "Ultrathin Epitaxial Graphite: 2D Electron Gas Properties and a Route toward Graphene-based Nanoelectronics". The Journal of Physical Chemistry B. 108 (52): 19912–19916. arXiv: cond-mat/0410240 . doi:10.1021/jp040650f. S2CID   848033.
  7. 1 2 3 de Heer, Walt A.; Berger, Claire; Ruan, Ming; Sprinkle, Mike; Li, Xuebin; Hu, Yike; Zhang, Baiqian; Hankinson, John; Conrad, Edward (2011). "Large area and structured epitaxial graphene produced by confinement controlled sublimation of silicon carbide". PNAS. 108 (41): 16900–16905. arXiv: 1103.3552 . doi: 10.1073/pnas.1105113108 . PMC   3193246 . PMID   21960446.
  8. Reich, Eugenie Samuel (2010). "Nobel document triggers debate". Nature. 468 (7323): 486. doi: 10.1038/468486a . PMID   21107397.
  9. 1 2 Emtsev, Konstantin V.; Bostwick, Aaron; Horn, Karsten; Johannes, Jobst; Kellogg, Gary L.; Ley, Lothar; McChesney, Jessica L.; Ohta, Taisuke; Reshanov, Sergey A.; Röhrl, Jonas; Rotenberg, Eli; Schmid, Andreas K.; Waldmann, Daniel; Weber, Heiko B.; Seyller, Thomas (2009). "Towards wafer-size graphene layers by atmospheric pressure graphitization of silicon carbide". Nature Materials. 8 (3): 203–207. doi:10.1038/nmat2382. hdl: 11858/00-001M-0000-0010-FA06-C . PMID   19202545.
  10. 1 2 Yu, X. Z.; Hwang, C. G.; Jozwiak, C. M.; Köhl, A.; Schmid, A. K.; Lanzara, A. (2011). "New Synthesis Method for the Growth of Epitaxial Graphene". Journal of Electron Spectroscopy and Related Phenomena. 184 (3–6): 100–106. arXiv: 1104.3907 . doi:10.1016/j.elspec.2010.12.034. S2CID   94674650.
  11. Virojanadara, C.; Syväjarvi, M; Yakimova, R; Johansson, L.I.; Zakharov, A.A.; Balasubramanian, T. (2008). "Homogeneous large-area graphene layer growth on 6H-SiC(0001)". Phys. Rev. B. 78 (24): 245403. doi:10.1103/PhysRevB.78.245403.
  12. Langmuir, Irving (1912). "Convection and Conduction of Heat in Gases". Physical Review. Series I. 34 (6): 401–422. doi:10.1103/PhysRevSeriesI.34.401. S2CID   51668492.
  13. Yazdi, G. Reza; Vasiliauskas, Remigijus; Iakimov, Tihomir; Zakharov, Alexei; Syväjärvi, Mikael; Yakimova, Rositza (2013). "Growth of large area monolayer graphene on 3C-SiC and a comparison with other SiC polytypes". Carbon. 57: 477–484. doi:10.1016/j.carbon.2013.02.022.
  14. 1 2 Tedesco, J. L.; Jernigan, G. G.; Culbertson, J. C.; Hite, J. K.; Yang, Y.; Daniels, K. M.; Myers-Ward, R. L.; Eddy Jr., C. R.; Robinson, J. A.; Trumbull, K.A.; Wetherington, M. T.; Campbell, P. M.; Gaskill, D. K. (2010). "Morphology characterization of argon-mediated epitaxial graphene on C-face SiC". Applied Physics Letters. 96 (22): 222103. arXiv: 1007.5064 . doi:10.1063/1.3442903. S2CID   119290647.
  15. 1 2 3 Ruan, Ming; Hu, Yike; Guo, Zelei; Dong, Rui; Palmer, James; Hankinson, John; Berger, Claire; de Heer, Walt A. (2012). "Epitaxial graphene on silicon carbide: Introduction to structured graphene" (PDF). MRS Bulletin. 37 (12): 1138–1147. doi:10.1557/mrs.2012.231. S2CID   40188237.
  16. Tromp, R. M.; Hannon, J. B. (2009). "Thermodynamics and Kinetics of Graphene Growth on SiC(0001)". Physical Review Letters. 102 (10): 106104. doi:10.1103/PhysRevLett.102.106104. PMID   19392131.
  17. Emtsev, Konstantin V.; Zakharov, Alexei A.; Coletti, Camilla; Forti, Stiven; Starke, Ulrich (2011). "Ambipolar doping in quasifree epitaxial graphene on SiC(0001) controlled by Ge intercalation". Physical Review B. 84 (12): 125423. doi:10.1103/PhysRevB.84.125423.
  18. Starke, U.; Forti, S.; Emtsev, K. V.; Coletti, C. (2012). "Engineering the electronic structure of epitaxial graphene by transfer doping and atomic intercalation". MRS Bulletin. 37 (12): 1177–1186. doi:10.1557/mrs.2012.272.
  19. Varchon, F.; Feng, R.; Hass, J.; Li, X.; Ngoc Nguyen, B.; Naud, C.; Mallet, P.; Veuillen, J.-Y.; Berger, C.; Conrad, E. H.; Magaud, L. (2007). "Electronic Structure of Epitaxial Graphene Layers on SiC: Effect of the Substrate". Physical Review Letters. 99 (12): 126805. arXiv: cond-mat/0702311 . doi:10.1103/PhysRevLett.99.126805. PMID   17930540. S2CID   26406232.
  20. Ouerghi, Abdelkarim; Silly, Mathieu G.; Marangolo, Massimiliano; Mathieu, Clair; Eddrief, Mahmoud; Pichter, Matthieu; Sirotti, Fausto; El Moussaoui, Souliman; Belkhou, Rachid (2012). "Large-Area and High-Quality Epitaxial Graphene on Off-Axis SiC Wafers". ACS Nano. 6 (7): 6075–6082. doi:10.1021/nn301152p. PMID   22702396.
  21. Camara, Nicolas; Tiberj, Antoine; Jouault, Benoit; Caboni, Alessandra; Jabakhanji, Bilal; Mestres, Narcis; Godignon, Philippe; Camassel, Jean (2010). "Current status of self-organized epitaxial graphene ribbons on the C face of 6H–SiC substrates" (PDF). Journal of Physics D: Applied Physics. 43 (37): 374011. doi:10.1088/0022-3727/43/37/374011.
  22. Hibino, H.; Kageshima, H.; Maeda, F.; Nagase, M.; Kobayashi, Y.; Kobayashi, Y.; Yamaguchi, H. (2008). "Thickness Determination of Graphene Layers Formed on SiC Using Low-Energy Electron Microscopy". e-Journal of Surface Science and Nanotechnology. 6. Surface Science Society Japan: 107–110. doi: 10.1380/ejssnt.2008.107 . ISSN   1348-0391.
  23. Yager, Tom; Lartsev, Arseniy; Mahashabde, Sumedh; Charpentier, Sophie; Davidovikj, Dejan; Danilov, Andrey; Yakimova, Rositsa; Panchal, Vishal; Kazakova, Olga; Tzalenchuk, Alexander; Lara-Avila, Samuel; Kubatkin, Sergey (2013). "Express Optical Analysis of Epitaxial Graphene on SiC: Impact of Morphology on Quantum Transport". Nano Letters. 13 (9): 4217–4223. doi:10.1021/nl402347g. PMID   23941358.
  24. Yu, Haohai; Chen, Xiufang; Zhang, Huaijin; Xu, Xiangang; Hu, Xiaobo; Wang, Zhengping; Wang, Jiyang; Zhuang, Shidong; Jiang, Minhua (2010). "Large Energy Pulse Generation Modulated by Graphene Epitaxially Grown on Silicon Carbide". ACS Nano. 4 (12): 7582–7586. doi:10.1021/nn102280m. PMID   21058692.
  25. Tzalenchuk, Alexander; Lara-Avila, Samuel; Kalaboukhov, Alexei; Paolillo, Sara; Syväjärvi, Mikael; Yakimova, Rositsa; Kazakova, Olga; Janssen, TJBM; Fal'Ko, Vladimir; Kubatkin, Sergey (March 2010). "Towards a quantum resistance standard based on epitaxial graphene". Nature Nanotechnology. 5 (3): 186–189. arXiv: 0909.1193 . doi:10.1038/nnano.2009.474. PMID   20081845. S2CID   35430873.
  26. Gao, Li; Guest, Jeffrey R.; Guisinger, Nathan P. (2010). "Epitaxial Graphene on Cu(111)". Nano Letters. 10 (9): 3512–3516. doi:10.1021/nl1016706. PMID   20677798.