Sie sind auf Seite 1von 47

Growing Graphene via Chemical Vapor Deposition

Benjamin Pollard Department of Physics, Pomona College May 2, 2011

Graphene, a two-dimensional nanoscale allotrope of carbon, is a promising material with many useful properties, including those of light transparency and electrical conductivity. Over the past few years research on graphene increased dramatically because of new methods to produce and study it. Since then researchers have proposed uses for graphene ranging from exible touch screens to vacuum membranes. Many of these proposals rely on graphene grown via chemical vapor deposition (CVD), a relatively new technique for producing large-area lms of contiguous, multi-domain graphene. Once created, CVD graphene is transferable to diverse substrates, making the technique versatile for many applications. One such application is as an electrode in an organic solar cell. This investigation explores the growth of graphene via chemical vapor deposition on copper and the subsequent transfer of that graphene to a silicon substrate, keeping in mind a potential application as a transparent electrode in an organic solar cell.

Contents
1 Background 1.1 1.2 1.3 1.4 Carbon Allotropes . . . . . . . . . . . . . . . . . . . . . . . . . . Properties of Graphene . . . . . . . . . . . . . . . . . . . . . . . 5 5 8 10 12

Solar Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Previous work at Pomona College . . . . . . . . . . . . . . . . . .

2 Theory 2.1 2.2 2.3 2.4 Structure of Graphene . . . . . . . . . . . . . . . . . . . . . . . . Graphene Growth Domains . . . . . . . . . . . . . . . . . . . . . Electrical Properties of Graphene . . . . . . . . . . . . . . . . . . Optical Properties of Graphene . . . . . . . . . . . . . . . . . . .

15 15 16 17 19

3 Fabrication and Measurement Techniques 3.1 3.2 3.3 3.4 3.5 Exfoliation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chemical Growth Methods . . . . . . . . . . . . . . . . . . . . . Chemical Vapor Deposition . . . . . . . . . . . . . . . . . . . . . Transfer Process . . . . . . . . . . . . . . . . . . . . . . . . . . . Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

20 20 21 23 25 26

4 Experimental Methods

28

4.1 4.2 4.3

CVD Growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Transfer Process . . . . . . . . . . . . . . . . . . . . . . . . . . . Characterization and Results . . . . . . . . . . . . . . . . . . . .

28 30 31

5 Discussion and Future Work

43

6 Acknowledgements

44

List of Figures
1 Macroscale photos and atomic structure diagrams of diamond and graphite. Image in the public domain, under the Creative Commons License, commons.wikimedia.org. . . . . . . . . . . . 2 Depictions of graphene, graphite, carbon nanotubes and buckyballs (adapted from [15]). . . . . . . . . . . . . . . . . . . . . . . 3 4 5 6 Diagram of a solar cell. . . . . . . . . . . . . . . . . . . . . . . . . Diagram of the band structure of ITO. . . . . . . . . . . . . . . . Diagram showing the graphene lattice unit cell. [14] . . . . . . . Diagram showing armchair and zig-zag cuts along a graphene lattice. [13] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 Fermi surface showing Dirac Cones and the zero-gap nature of graphene [14]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 Diagram of CVD growth on copper. [26] . . . . . . . . . . . . . . 18 23 16 7 11 13 15 6

9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Schematic diagram of the transfer process to an arbitrary substrate. 26 Photo of copper foil after graphene growth. . . . . . . . . . . . . Photographs of growth furnace setup. . . . . . . . . . . . . . . . Optical microscope image of graphene on copper foil. . . . . . . . Optical microscope image of graphene on copper foil. . . . . . . . SEM image of copper domains, 200m. . . . . . . . . . . . . . . 28 29 31 32 32 33 34 34 35 36 36 37 38 39

SEM image of copper domains, 20m. . . . . . . . . . . . . . . . SEM image of graphene domain boundary, 2m. . . . . . . . . . SEM image of copper steps, 2m. . . . . . . . . . . . . . . . . . . SEM image of black spots, 5m. . . . . . . . . . . . . . . . . . . Raman Spectrum of graphene on copper, UC Riverside. . . . . . Raman Spectrum of graphene on copper, Pomona. . . . . . . . . Raman Spectrum of graphene on SiO2 , UC Riverside. . . . . . . Optical thickness measurement of drop-cast PMMA. . . . . . . . Optical thickness measurement of spin-cast PMMA. . . . . . . . Microscope images of SiO2 wafer with evaporated PMMA over time. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

41 42 42 43

25 26 27

Photo of SiO2 wafter after graphene transfer. . . . . . . . . . . . Microscope image of SiO2 wafer after graphene transfer. . . . . . Microscope image of SiO2 wafer after graphene transfer. . . . . .

1
1.1

Background
Carbon Allotropes

Carbon, one of the most common atoms on Earth, occurs naturally in many forms and as a component in countless substances. However, there are only a handful of materials made solely of carbon and nothing else. These are called allotropes of carbon. Two of these carbon allotropes have been collected from nature and used by humans for centuries; they are graphite and diamond (see Figure 1). Carbon can also occur as an unordered mess of atoms; this is called amorphous carbon and will not be covered here. A related form, also outside the scope of this investigation, is glassy carbon which has a semi-ordered structure with bonds resembling other forms. Finally, there are three nanoscale forms of carbon that have attracted widespread attention over the last half-decade because of their novel properties. These carbon nanostructures are called buckyballs, carbon nanotubes, and graphene. Diamond is the most stable form of pure carbon. Formed under high temperatures and pressures under the earths crust, diamond is a tetrahedral lattice with a carbon atom at each vertex. Each carbon atom thus forms four covalent bonds with four neighboring atoms, completely lling its outer electron shell and resulting in one of the hardest and most valued substances in human history. Pure diamond has a wide bandgap and thus acts as a transparent insulator, which as a single crystal gives diamond its dazzling optical properties. Impurities and dopants in diamond lead to other colors and bandgaps, giving rise to rarer gemstones and specialized research respectively. All other allotropes of carbon can be conceptualized as variations on the lattice structure of graphene (see Figure 2). Graphene is actually the most recent

Figure 1: Macroscale photos and atomic structure diagrams of diamond and graphite. Image in the public domain, under the Creative Commons License, commons.wikimedia.org.

Figure 2: Depictions of graphene, graphite, carbon nanotubes and buckyballs (adapted from [15]). carbon nanomaterial to be widely studied, but its basic structure is simple. Consider a 6-carbon ring of atoms, and then tessellate that hexagon to form a 2D hexagonal lattice similar to the surface of a honeycomb. Such a 2D sheet is known as graphene. Graphenes properties are striking in a number of respects, but perhaps most notable is that a single graphene sheet is quite stable and mechanically resilient, as well as very electrically conductive. By far the most common form of pure carbon is graphite. Graphite is simply many layers of graphene stacked on top of each other. While each sheet is tightly bound, only weak bonds exist between layers. This enables the layers to slide laterally, making graphite slippery. Thus, graphite is commonly used as a lubricant. Graphites other common usage is as the core of a pencil, where akes of graphite slide o the bulk material and remain as a mark on paper. Carbon nanotubes (CNTs) are another nanoscale allotrope of carbon. They

can be thought of as ribbons of graphene that have been rolled into a tube. While only nanometers in diameter, CNTs can grow to millimeters in length. Due to the strength of the bonds in a hexagonal carbon lattice, nanotubes are one of the strongest bers ever discovered. Additionally, due to the extra quantum connement imposed on electrons along the circumferential axis, carbon nanotubes can display both metallic and semiconducting electric properties. The electrical nature of a nanotube stems from its physical shape, making CNTs intriguing materials for pure research and numerous electromechanical applications. Lastly, a buckyball is created by collapsing yet another dimension. Conceptually, a buckyball is a small segment of a carbon nanotube that has been pinched together at both ends to form a hollow sphere of carbon atoms. Named after Buckminster Fuller, a architectural engineer and science-ction writer who designed domes with a similar shape, the 60-carbon buckyball was the rst carbon nanomaterial to gain widespread attention. Buckyballs have many proposed uses, such as encapsulation of reactive compounds in chemistry, isolation of quantum systems to make a functional qubit, and fundamental quantum experiments in which an entire buckyball acts as both a particle and a wave.

1.2

Properties of Graphene

Graphene is a nanoscale allotrope of carbon. Unlike graphite, the most common allotrope, graphene is quasi-two-dimensional, since electrons can only move between carbon atoms in the 2D lattice. The extra quantum connement of the electrons due to the lack of a third dimension gives graphene various novel properties. For example, electrons interact with carbon atoms in the lattice to create a system that acts like a single mobile charge carrier. The carrier moves ballistically over the graphene surface, enabling graphene sheets to conduct electricity

very well. [15] Other complex interactions between electrons and the hexagonal lattice make graphene transparent, exible and strong. [6] These properties and others have compelled many researchers over the last half-decade to study graphene for a diverse array of uses. While graphite has been used for ages in a range of purposes from lubricant to pencils, researchers only began widely studying graphene around the year 1990. For the rst decade, research was hindered due to the diculty of producing it in an electrically isolated environment and without defects. However, in 2004, two researchers named Andre Geim and Konstantin Novoselov at Manchester University discovered a new method for producing graphene through mechanical exfoliation. [19] Called the scotch tape method, the procedure is detailed below under Exfoliation. Geim and Novoselovs discovery gave researchers access to pure graphene in a number of desirable environments for experimentation. Furthermore, they found that if silicon dioxide was used as a substrate for the graphene akes, the akes appeared as a discoloration under any optical microscope. These properties made it possible to perform electrical and nanomechanical experiments on graphene that began to showcase the materials novel properties. The scotch tape method sparked widespread research on graphene in many areas of physics and materials science, and won them the Nobel Prize in Physics in 2010. Aside from its use as a transparent electrode in an organic solar cell, graphene has been considered for a huge range of purposes over the last half-decade. Its mechanical strength makes it attractive as a optically transparent membrane. Experiments on gases or substances in vacuum are envisioned in which graphene is used as a window, preserving a closed environment while still allowing precise optical measurements. [6] Graphene has also been proposed as a protective coating on metals such as copper to prevent corrosion. While not noticeably altering 9

the appearance of the underlying metal, graphene eectively prevents oxidation from the surrounding environment. [27] Graphene is also of interest as a simple 2D quantum system. Experiments have already been completed that show interesting quantum phenomena in a clear and denitive way. The Quantum Hall eect, now a common test for graphene purity, is one such example. [13] Further research using graphene promises to test predictions from applied quantum theories and provide insight into fundamental quantum physics.

1.3

Solar Cells

Among all the many exciting applications of graphene, use in organic solar cells stands out as both important and accessible. Therefore, solar cells have guided recent work on graphene at Pomona College as a context and end-goal of graphene growth and transfer processes. Fundamentally, solar cells convert photons of light into electric current. Otherwise known as photovoltaic cells, they work by transferring the energy of an incoming photon to a valence electron. This electron then has enough energy to escape the connes of the atom it was bound to, leaving behind a hole of positive charge. The electron and hole must be forced to separate spatially and enter electrodes on opposite ends of the cell. They will then create a voltage dierence between the two electrodes, which generates a current when connected to any electrical device drawing power. Solar cells are commonly fabricated by depositing dierent materials one after another, creating a stack in which each layer performs a specic function (see Figure 3). Special materials are required to perform each required function. Semiconductors with precisely-tuned band gaps create an environment in which electrons are excited by incoming photons and yet do not recombine with their 10

holes before they have a chance to separate. This is known as the active layer. Electrodes with specic work functions sandwich the active layer and pick up electrons or holes. Charge separation is best achieved if the contact area between the active layer and the electrodes is maximized, so it is best (at least for organic cells) if the electrodes completely cover the area of the cell. However, light must still be able to pass through the top electrode to interact with electrons in the semiconductor. Thus, the top electrode must be both transparent and conducting.

Figure 3: Diagram of a solar cell. Materials that are transparent and conducting are uncommon, since typical conductors are metals, which in turn are typically opaque. In the language of semiconductors, this is because metals lack a band gap between valence and conduction levels. More simply, the Fermi level in a metal is surrounded by available electron states, so that any excited electrons can easily transition into accessible higher-energy states and move between atoms in the crystal lattice [11]. This is why metals are good conductors of heat and electric current. That same freedom of electron mobility also makes metals opaque. When a photon of light enters a metal, valence electrons can easily absorb that photon and

11

temporarily enter a higher-energy state, ultimately settling back to the valence band and releasing a phonon. Therefore light is quickly absorbed by the metal and turned into heat energy instead of passing through uninterrupted. Conventional organic solar cells use thin lms of a material called indium tin oxide (or ITO) as a transparent conducting electrode. ITO is a wide-bandgap n-type semiconductor. Its bandgap is larger than the energy of visible light, so electrons in the material cannot absorb those photons and thus ITO is visibly transparent. Conduction in ITO is achieved through loosely-bound electrons from oxygen atoms creating vacancies which are lled by nearby electrons in the main indium lattice, creating energy levels called impurity states close to the conduction band (see Figure 4). [8] Because of these special properties, ITO is widely used in liquid crystal displays and touch screens in addition to solar cells. Indium, however, is a rare metal, and given its increasing use due to the rise of these technologies, many fear a worldwide shortage of indium in the near future. Even now, the price of indium is increasing drastically. Furthermore, ITOs transparency and conductivity are imperfect, and an increase in one comes at the cost of the other. [22]

1.4

Previous work at Pomona College

Students under the direction of Prof. David Tanenbaum and his colleagues have been investigating carbon nanostructure synthesis and properties for about a decade. The lab started by investigating carbon nanotubes grown via chemical vapor deposition with the work of Matthew Ferguson, James McFarland, Elias Penilla and Ajoy Vase. Ian Frank (class of 2008), now a graduate student at Harvard, transitioned to graphene by looking at exfoliated graphene with Paul McEuen at Cornell University. Using the cheese grater approach (detailed be-

12

Figure 4: Diagram of the band structure of ITO.

13

low), Ian investigated suspended graphene akes using both electromechanical resonance and the tip of an atomic force microscope (AFM). By applying resonant electric elds or pushing on the akes with an AFM tip, Ian was able to measure the mechanical properties of pure few-layer graphene. Subsequently, Ian did a senior thesis at Pomona College on the scotch tape method of graphene exfoliation (also detailed below). After establishing a setup for creating exfoliated graphene at Pomona, Ian investigated the potential for optical and electron beam lithography on it with the aim of producing graphene devices of any shape and reasonably small size. I had the privilege of assisting Ian with the optical lithography aspect of this work during my rst year at Pomona. [13] Scott Berkley (class of 2009) also spent two summers at Cornell University working with Paul McEuen and David Tanenbaum. During the rst summer, Scott learned the cheese grater approach (below) of graphene exfoliation and used the resulting suspended graphene akes to further investigate the mechanical properties of graphene with an AFM tip. [4] During his second summer at Cornell, Scott transitioned to the scotch tape method (below). He characterized the number of layers that could result from that method and began investigating the dierent interactions with laser light which arose from various numbers of layers. [3] Scott went on to complete his thesis on organic solar cells, during which time I worked separately on investigating an e-beam lithography system for the electron microscope at Pomona College.

14

2
2.1

Theory
Structure of Graphene

Graphene is a 2D sheet of carbon atoms arrayed in a hexagonal honeycomb lattice. The sheet is held together with sp2 bonds between the carbon atoms separated by a distance of about 1.4 angstroms, making the sheet quite strong. A few such layers stacked on top of each other is still considered graphene; it takes at least 10 layers (and in some respects more like 100) before a sample becomes bulk graphite. There are about 3.4 angstroms between stacked sheets. [28]

Figure 5: Diagram showing the graphene lattice unit cell. [14] The honeycomb lattice can be analyzed with a two-atom unit cell as a Bravais lattice (see Figure 5). By mentally duplicating and translating this cell by a set amount along set translation vectors, the entire lattice can be constructed. There are two possible cuts along a honeycomb lattice; these are entitled armchair and zig-zag due to the appearance of the resulting jagged edge along such a cut (see Figure 6). The orientation of a lattice, specically whether a cut or a current is along the armchair or zig-zag direction, has interesting 15

fundamental eects on the electronic behavior of graphene. [13]

Figure 6: Diagram showing armchair and zig-zag cuts along a graphene lattice. [13]

2.2

Graphene Growth Domains

This investigation focuses on the growth of graphene via chemical vapor deposition, as described in detail below. In this process, carbon atoms adhere to the surface of a metal substrate under high temperatures. Once a carbon atom occupies a position on the surface of the substrate it pushes other carbons to the side, creating a one atom thick layer of carbon. As the temperature is lowered the carbon crystallizes into a layer of graphene. [19] Unavoidably, the graphene crystallization will start at various places on the surface of the substrate before the entire area has formed a lattice. Each initial crystallization is referred to as a nucleation site, and establishes an orientation

16

for the lattice that grows from it. As various crystal regions grow out from nucleation sites, their borders will meet and a discrepancy will probably occur between the lattice orientations of each region. This will create a denite boundary between regions. Growth stops when every region is surrounded by such boundaries (or the edge of the substrate). At this point the regions are called domains. In a sense, domain boundaries represent defects in the crystal structure of the graphene, since along those lines the bonding of the carbon atoms does not follow the simple Bravais lattice from a repetition of the unit cell. This acts as a barrier for charge transport phenomena and an exception to graphenes optical properties (both discussed below). Therefore, it is desirable to maximize the size of domains to limit the frequency of domain boundaries.

2.3

Electrical Properties of Graphene

In most conductors, the valence and conduction bands overlap, giving excited electrons many states to occupy as they move throughout the material. Materials with this property are known as metals. Graphene, while an excellent conductor, is not a metal but rather a zero-gap semiconductor. While the valence and conduction bands do not overlap in graphene, they touch at the Fermi level. [29] This can be seen by visualizing the Fermi surface of a 2D graphene lattice, as in Figure 7. The Fermi surface for a lattice material is the energy border between the valence and conduction bands in momentum space. For this border to be dened the Fermi energy must fall inside an energy band and not in a band gap; otherwise the valence and conduction bands do not touch at all. Thus, Fermi surfaces only exist for conductors. Graphenes Fermi surface consists of six double cones with the Fermi energy at the intersection of those

17

cones. Because the cones are linear near this intersection the eective mass of electrons in this region is zero (since eective mass is given by the curvature of the energy bands in momentum space, and thus the curvature of the Fermi surface). [12] This leads to an entirely new transport mechanism in graphene compared to metals. The specics of this regime rely on quantum electrodynamics and Diracs relativistic equation of state. Without going into the details of these theories, the results can be conceptualized by thinking of charge carriers in graphene not as individual electrons, but as interacting groups of electrons that behave as an entirely dierent type of particle. Called a Dirac fermion, these charge carriers travel ballistically over the 2D surface at relativistic speeds. [25] Because of this fundamentally dierent transport regime, pure graphene is able to conduct electricity better than metals, with room-temperature resistivity on the order of 106 cm. [15]

Figure 7: Fermi surface showing Dirac Cones and the zero-gap nature of graphene [14].

18

2.4

Optical Properties of Graphene

Aside from being incredibly conductive and strong, graphene is even more attractive to work with because of its optical transparency. Simply because graphene is thin, photons easily pass through it. In actuality, graphene has a surprisingly high absorption rate for being only one atomic layer thick: 2.3% of incident white light is absorbed by a single graphene sheet. Intriguingly, this value is exactly equal to , where is the ne structure constant (e2 / c). This can be derived using quantum mechanical principles applying to 2D Dirac fermions. [24] All carbon allotropes have a particular anity to light at specic wavelengths. These wavelengths correspond to the vibrational modes of sp2 carboncarbon bonds, such that when any higher-energy light excites the carbon material, photons are re-radiated at those wavelengths. The technique of exciting molecular vibrational modes and measuring re-radiated light is known as Raman spectroscopy, and is the easiest and most reliable method of determining the presence of graphene. Graphene produces two strong optical peaks in Raman spectra: the G peak and the D peak. The G peak is due to individual bonds stretching and compressing, while the D peak is due to breathing modes of the hexagonal rings of carbon atoms. They occur at 1560 and 1360 cm1 respectively. [9] Peaks can also be observed at twice those values due to the next harmonic mode of the oscillation. While graphenes transparency makes it dicult to see on most substrates, a particular interaction occurs between a graphene sheet and a silicon substrate with a 100nm layer of silicon dioxide on top that makes the graphene observable under an optical microscope. Silicon dioxide naturally forms on the surface of a pure silicon wafer, but a silicon dioxide layer of any desired thickness can be de-

19

liberately grown thermally on a silicon wafer for use in the lab. When graphene is deposited or transfered onto such a substrate, the index of refraction at the surface of the silicon dioxide changes due to the graphene lm. This results in a slight discoloration, from pink to purple, in the places where graphene exists on the wafer. [13, 18]

3
3.1

Fabrication and Measurement Techniques


Exfoliation

Graphene has been the subject of intense widespread research for less than a decade. Most of this work used graphene created by a process of mechanical exfoliation called the scotch tape method. In this procedure, pure samples of bulk graphite are placed on the sticky side of common adhesive tape. The tape is pressed on a desired substrate and then peeled away. Flakes of graphene around 50 microns wide are left on the substrate, along with chunks of graphite and adhesive residue. The akes can be discerned under an optical microscope due to thin-lm interference, appearing as a region of slight discoloration. [15] The graphene left by the scotch tape method is pure and clean, which enables researchers to measure its electrical and mechanical properties exactly. However, a fair amount of time and luck are required to manually locate an appropriate ake on the region exposed to the tape. This diculty is compounded if the graphene ake must be positioned in a certain way above or around an existing feature on the substrate, as is commonly desired for many nanoscale experiments. Lastly, for graphene to be used as an electrode on a solar cell it must cover the entire surface area of the cell, which is much larger than the area of a single ake. 20

Aside from the scotch tape method, there is another method of graphene exfoliation used extensively by Paul McEuen at Cornell University to create suspended graphene sheets. I will refer to this method as the cheese grater method. For this technique, pure bulk graphite is attached to the end of a rod for support; a toothpick is sometimes used. [4] A silicon wafer (with a silicon dioxide layer around 0.25 microns) is also prepared with trenches etched into it using radio frequency plasma etching. The graphite is then dragged across the trenches. The trenches, acting like a cheese grater, pull o pieces of the graphite their corners, and akes of few-layer graphene are pulled over the trench. This results in exfoliated graphene suspended over a trench (up to 0.5 microns deep) which is ideal for performing mechanical and electromechanical measurements. [7]

3.2

Chemical Growth Methods

While exfoliation produces very pure single-domain graphene with nearly ideal mechanical and electrical properties, it has one large disadvantage. That is, exfoliation results in graphene akes scattered randomly on a substrate. Each ake is on the order of only microns in size, and much of the substrate remains uncovered. For many applications of graphene discussed earlier (including transparent conducting electrodes for an organic solar cell), a contiguous covering of graphene is needed. To produce contiguous graphene lms, exfoliation cannot be used and chemical methods are needed instead to grow graphene from carbon atoms in another form. Common methods for chemical growth of graphene include reduced graphene oxides, molecular beam epitaxy, plasma-enhanced CVD, and chemical vapor deposition. The rst two will be briey discussed below, while chemical vapor deposition and plasma-enhanced CVD will be covered in

21

the next section. The technique of reduced graphene oxides is really the intersection of exfoliation and chemical growth methods. Exfoliated graphene akes are oxidized, enabling them to be suspended in aqueous solution. This solution is then passed through a lter membrane with pores around 25 nanometers. The graphene oxide akes get caught by the membrane until the entire surface of the lter is covered with graphene sheets. This covering can then be transfered to a more desirable substrate. While the purity of the resulting graphene lm is high, the coverage of the lm is often nonuniform. Parameters must be carefully controlled to get the entire lter area covered. Additionally, the result of this method is a lm of graphene oxide as opposed to just graphene. Graphene oxide lms must be further treated chemically to to make them electrically conducting instead of insulating. [2] Epitaxial graphene is a commonly used technique for creating high quality monolayer graphene. Originally, epitaxial graphene was grown from silicon carbide (SiC). When bulk SiC is heated to around 1500C some of the silicon sublimates, leaving a layer of carbon behind on the surface. [10] Another method of creating epitaxial graphene from SiC is that of molecular beam epitaxy. A graphite lament is loaded into an ultra-high vacuum. As the lament is heated, carbon atoms sublimate o of the graphite. These carbons form a molecular beam in the vacuum, traveling through free space without interacting until they land on a metallic substrate (such as iridium) and form a graphene layer. [23] While molecular beam epitaxy produces high-quality uniform lms over a large surface, it requires an ultra-high vacuum which makes the process tedious and inaccessible to smaller groups.

22

3.3

Chemical Vapor Deposition

A more recent alternative to the scotch tape method is that of chemical vapor deposition, or CVD. In CVD, a metal substrate such as copper is put into a furnace and heated under low vacuum to around 1000C. The heat anneals the copper, increasing its domain size. [1] Methane and hydrogen gases are then owed through the furnace. The hydrogen catalyzes a reaction between methane and the surface of the metal substrate, causing carbon atoms from the methane to be deposited onto the surface of the metal through chemical adsorption (see Figure 8). The furnace is quickly cooled to keep the deposited carbon layer from aggregating into bulk graphite, which crystallizes into a contiguous graphene layer on the surface of the metal. [19]

Figure 8: Diagram of CVD growth on copper. [26] The graphene produced by this method is more likely to carry impurities due to the various materials required for CVD. However, research has shown that such impurities can be suciently minimized to create graphene as pure as exfoliated akes. [20] Additionally, the graphene from CVD tends to wrinkle due to the dierence in thermal expansion between graphene and copper. This is decreased via proper annealing, but is still an ongoing research challenge. [1] Most importantly, graphene from CVD is a contiguous lm as large as the underlying metal substrate, in stark contrast to the random micron-sized akes

23

from the scotch tape method. CVD thus allows graphene to be used as a layer in a solar cell. There are many ways to aect the outcome of a CVD graphene growth run. Since the growth dynamics of carbon deposition and domain growth are not yet fully understood, nding the proper balance of these controls is a largely experimental task. [16] Perhaps the most natural variable to aect a CVD outcome is the amount of the various reaction gases. Increased methane provides more carbon atoms to deposit (and more nucleation sites leading to more domains), while increased hydrogen promotes the reaction and also increases chemical processes on the copper and surrounding environment. The temperature also aects the rate of reaction, as does the speed of changes in temperature. Impurities in the copper substrate detract from the growth process by encouraging nucleation sites and thus hindering the formation of contiguous carbon domains, so proper chemical cleaning of the copper is essential. Annealing time of the copper also aects the level of impurity for the same reason. The geometry of the growth chamber aects the deposition rate of carbon due to its eect on gas ow patterns, specically because of turbulent (instead of laminar) ow regimes. Finally, any leaks in the vacuum system further detract from the growth, as oxygen from the air oxidizes the copper, making the carbon atoms unable to adhere to the copper surface and ruining the deposition. Copper is not the only substrate which can be used in graphene CVD; in fact many transition metals can be used. For example, graphene CVD on nickel is somewhat common, and cobalt has also been used. [5] The main dierences between metal substrates come from dierences in the metals ability to absorb carbon. Nickel and cobalt absorb carbon more than copper, and this leads to an overabundance of carbon on foils which crystallizes into discrete graphite chunks instead of a single graphene sheet. For that reason, nickel and cobalt foils cannot 24

be used and instead thin lms (< 300nm for nickel) must be evaporated onto a silicon substrate before growth. [19] Copper, on the other hand, attracts less carbon and does so only at the surface rather than absorbing it into the bulk of the material, since the weak bonds that hold the carbon atoms to the copper can only be formed with open bonding sites at the surface of the lattice. Therefore copper foils can be used in graphene CVD, simplifying the production process as a whole and making it more robust. While not used in this investigation, it should be noted that a fairly common variant on CVD is that of plasma-enhanced CVD. PECVD works in much the same way as has already been described, but in addition to using a furnace to provide the heat energy for substrate annealing, an RF frequency AC current is passed through the substrate. This spark ionizes the gases in the chamber, enhancing the deposition onto the substrate. [21] While PECVD can be done at much lower furnace temperatures than regular CVD, it requires additional equipment beyond the system available at Pomona College.

3.4

Transfer Process

Another key advantage to CVD graphene growth is the ability to transfer the graphene to an arbitrary substrate (see Figure 9). Once the graphene/copper foil has been removed from the furnace and cooled, a polymer such as polydimethylsiloxane (PDMS) or polymethyl methacrylate (PMMA) can be spincoated onto the graphene as a support, and then the copper removed using an etchant such as ferric chloride (FeCl3 ). This leaves the graphene attached only to the polymer, which can be positioned onto any other substrate (such as a solar cell). A solvent can easily dissolve the polymer, leaving just the graphene on any desired substrate. [1, 19]

25

Figure 9: Schematic diagram of the transfer process to an arbitrary substrate. Various substrates are useful for specic purposes and stages of research. As with the scotch tape method, silicon dioxide allows otherwise-transparent graphene to be seen under an optical microscope. Thus silicon dioxide is a useful substrate to investigate the uniformity of a growth procedure. Silicon dioxide is also a good substrate to perform electrical measurements on the graphene, further measuring its purity and checking its conductive properties. Once the grown graphene has been veried and characterized on silicon, it can just as easily be transfered to substrates involved in producing a solar cell such as glass slides or organic lms.

3.5

Measurement

Since graphene is transparent, verifying with the naked eye that it has indeed grown on a metallic foil is dicult. Raman spectroscopy, however, can be used to quickly verify the presence of graphene. In Raman spectroscopy, a laser is directed at the material in question, and the re-emitted light is measured. The incoming laser light excites characteristic molecular vibrations in the sample, which emit photons at characteristic frequencies. Thus, if the frequencies associated with carbon-carbon bonds are observed, there is graphene on the copper

26

sample. [9] Raman spectroscopy is a quick procedure, but requires a special tool for performing the measurement. These tools are usually purchased as a single unit. A Raman tool is set up in the Chemistry Department at Pomona College, but better tools exist for measuring graphene because of the spectral range which they cover. While not as conclusive, optical and e-beam microscopy can also indicate the presence of graphene on copper by revealing graphene domain boundaries. Once graphene has been transferred to a silicon substrate with a dioxide layer, it can be seen as a slight discoloration under an optical microscope. Atomic force microscopy can also be performed on graphene that has been transfered to a at rigid substrate to directly measure the thickness and thus number of layers of the graphene lm. Both techniques take under an hour and are available in the Pomona Physics Department. It is also useful to measure the thickness of a PMMA lm to investigate the transfer process. The thickness of thin lms on a reective substrate can be measured using a simple spectrometer and light source. If light is shone on such a thin lm, it will reect only at wavelengths equal to integer half multiples of the thickness of the lm, due to interference between the wavefronts reected from the top and bottom of the lm. The intensity of reected light R is governed by the proportionality R cos( 4dn ),

where d is the thickness of the lm, n is the index of refraction of the lm, and is the wavelength of that light. By taking a reectance spectrum of a thin lm illuminated by white light and tting the resulting curve to a function of this form, the parameter d can be extracted as a measure of the lms thickness. [17] Taking about ve minutes, this technique can be performed at Pomona College,

27

though the proper equipment might need to be gathered and assembled each time.

4
4.1

Experimental Methods
CVD Growth

The growth recipe used in this investigation follows that of the McEuen group (see [30]). Graphene was grown on .025mm copper foil from Alfa Aesar. A piece of foil was cut approximately 2cm x 3cm, and a small cut was made in the lower right corner for positional reference later on (see Figure 10). The piece was sandwiched between two glass slides and clipped with plastic alligator clips, and left to atten.

Figure 10: Photo of copper foil after graphene growth.

Before insertion into the furnace, the copper foil was cleaned by dipping it in various solvents. The order of these dips was: 28

1. Acetone (10 seconds) 2. Water (< 10 seconds) 3. Acetic Acid (10 minutes) 4. Water (< 10 seconds) 5. Acetone (10 seconds) 6. Isopropyl alcohol (IPA) (10 seconds)

The remaining IPA was removed using compressed air, and the copper then loaded into the furnace tube (see Figures 11a and 11b). Multiple foils were often loaded for a single run.

(a) Photo of ow regulators, gas tanks and growth furnace.

(b) Photo of growth furnace and pump.

Figure 11: Photographs of growth furnace setup. Once all the foils were loaded, the furnace vacuum system was closed o from the gas tanks and pumped down to around 30 mTorr. Hydrogen was then owed at a pressure of 150 mTorr and the furnace heated to 1000C (taking about 20 minutes). The furnace was held at this temperature for 15 additional minutes to allow the copper to anneal. Then, methane was owed at a rate measured by the coarse ow meter at the 89 mark (approximately 6 Torr) for 29

13 minutes. The furnace heating system was then turned o, and the door opened 2-3 inches. When the temperature reached 450C, the door was opened completely. When the temperature reached 150C, argon ow was started at a rate measured by the ow meter at the 120 mark, and the hydrogen and methane ows were stopped. This left the pressure at around 1.8 Torr. After 2 minutes of argon ow, the pump was turned o and the pump valve closed. The pressure was allowed to gradually rise to room pressure by the ow of argon, after which the argon was stopped and the copper removed. [30]

4.2

Transfer Process

Once graphene was grown on a copper foil, it could be transferred to any other substrate. First, PMMA (4% in Anisole) was drop-cast on the copper. This took about 1.5 hours to evaporate, after which it was cured on a 145C hot plate for 3 minutes. A piece of scotch tape was placed sticking to the bottom edge of the PMMA to act as a handle later. The sample was then O2 plasma cleaned for 2 minutes to remove the graphene on the opposite side of the copper. It was then placed in a ferric chloride bath overnight to remove the copper. After the copper etch the sample was rinsed in dH2 O thoroughly until no ferric chloride color could be seen in the water. The graphene could be stored in this way oating in a glass jar of water. To transfer to a substrate, the substrate was rst cleaned. The PMMA was then placed on the substrate (graphene side down) directly from a bath of dH2 O. Surface tension in the water held it on the surface, however care had to be taken to avoid wrinkles. The sample was placed in a covered acetone bath overnight to remove the PMMA. It was then removed (with a bubble of acetone still covering the surface) and placed in an IPA bath for 1 hour. The sample

30

was then placed in a desiccator box to dry, after which it was ready for use.

4.3

Characterization and Results

Figure 12: Optical microscope image of graphene on copper foil.

After graphene was grown on copper foils using the procedure described above, many techniques were used to characterize it. They were all aimed primarily at verifying that graphene was indeed grown on the copper, and to get a sense of the uniformity of graphene coverage. First, optical microscope images were taken of graphene grown on copper foils at 50x and 100x magnication (see Figures 12 and 13). These images show two salient features. First, the ridges in the copper substrate are clearly visible. They are due to the milling process of copper foils, and are expected. Secondly, faint boundary lines can be seen separating regions of slightly dierent shade. Those are believed to be boundaries between domains of graphene growth. Scanning Electron Microscope (SEM) images were also taken of graphene

31

Figure 13: Optical microscope image of graphene on copper foil.

Figure 14: SEM image of copper domains, 200m.

32

Figure 15: SEM image of copper domains, 20m. grown on copper foils, illustrating many features of these samples. At low magnication (around 500x), copper domain boundaries are clearly visible as patches of varying shade (see Figures 14 and 15). These were only observed on some samples, possibly due to a dierence in copper annealing between the side facing up and the side facing down while in the furnace. At higher magnications (10000x), dark lines criss-cross the surface of the sample (see Figure 16). I believe those are edges between graphene regions; possibly tears in the graphene resulting from a dierence in thermal expansion between graphene and copper. Also at 10000x, steps or ridges in the copper substrate can be seen (see Figure 17). This is evidence that the graphene is covering the copper uniformly, as the copper would oxidize in the furnace as it cooled (and the steps would disappear) if the graphene was not there to protect it. Lastly, black spots can be seen dotting the surface at 5000x (see Figure 18). It is unclear what these spots are, but they seem to follow the graphene tears to some extent. It is possible that these spots are partially oxidized copper that is exposed through breaks in the 33

Figure 16: SEM image of graphene domain boundary, 2m.

Figure 17: SEM image of copper steps, 2m.

34

Figure 18: SEM image of black spots, 5m. graphene. It is also possible, however, that these are clumps of graphite that coalesced as the graphene formed. To ensure that graphene does in fact exist on the copper foil, Raman spectroscopy data was collected at UC Riverside (see Figure 19). Peaks in the graphene spectrum can be seen that are distinct from the spectrum for the foil without graphene. They match the expected values for the G peak (1560 cm1 ) and twice the D peak (the 2D peak, 2720 cm1 ), indicating that graphene is present on the foil after growth. Similar data was taken using Pomonas Raman tool (see Figure 20), but the lower wavelength range of that tool and our relative inexperience at measuring graphene compared to the Riverside group made the data less conclusive. However, a rough D peak can still be seen around 1360 cm1 . Raman spectra were also taken at Riverside of a rst attempt to transfer graphene to a SiO2 substrate (see Figure 21). Firstly, the steady sloping back-

35

Figure 19: Raman Spectrum of graphene on copper, UC Riverside.

Figure 20: Raman Spectrum of graphene on copper, Pomona.

36

Figure 21: Raman Spectrum of graphene on SiO2 , UC Riverside. ground of the copper substrate is visibly absent. Spectra were taken at various points on the surface to characterize dierent optical features; this is shown in the inset. Two types of spectra can be dierentiated. The pink interior point is similar to raw SiO2 , while the patchy and yellow points have additional peaks. I conclude that those additional peaks are due to graphene on the surface, which is supported by the fact that these peaks lie at the G, D and 2D positions. The peaks common to all spectra are therefore from the substrate itself. The smaller peaks on either side of the 2D peak remain mysterious, but might be due to multiple layers of graphene interacting and shifting the vibrational energy levels. One of these peaks is close to the 2G position (3120 cm1 ); the dierence might also be due to these multi-layer interactions. All of this information suggests to me that the graphene has torn and folded up on itself during the transfer process, leaving areas of the substrate exposed. Turning to the process of graphene transfer to another substrate, I investigated an alternative method for removing PMMA lm once the transfer is performed. The rst few times I tried the procedure I removed the PMMA in

37

(a) Original drop-cast

(b) After 2.5h evaporation

Figure 22: Optical thickness measurement of drop-cast PMMA.

38

(a) Original spin-cast

(b) After 2.5h evaporation

Figure 23: Optical thickness measurement of spin-cast PMMA.

39

an Acetone bath overnight; this sample was used in the Raman data above. However, the Acetone bath seems to cause the graphene to oat, tear and fold. An alternative approach is to set the substrate on a hot plate at 300C and allow the PMMA to evaporate for many hours. To test this process, I deposited a PMMA lm on a fresh SiO2 substrate and measured its thickness before and after evaporation. I tried both a drop-cast of 4% PMMA in anisole and a spincast of 8% PMMA in anisole at 2000rpm (both annealed at 145C for 1 min). Using the optical lm thickness measurement described in the Theory section above, I measured the thicknesses of these lms to be 4385nm for the drop-cast and 856.5nm for the spin-cast, with uncertainties smaller than the signicant digits of the values (see Figures 22a and 23a). After setting these samples on the hot plate for 2.5 hours, I repeated the measurement (see Figures 22b and 23b). This yielded thicknesses of 258.0nm for the drop-cast and 47.8nm for the spincast. Clearly 2.5 hours is not long enough to completely evaporate the PMMA lm, so I repeated the measurement after 5.5 hours. At this point, however, the spectra were practically at and no useful thicknesses could be extracted, indicating that the lm had largely evaporated. To further investigate the thickness of PMMA over evaporation time, I used the optical microscope to observe thin-lm interference between the PMMA and the silicon dioxide layer. I observed both a drop-cast and a spin-cast sample using the same parameters as before. The drop-cast was too thick to see any thin-lm interference: it went from discolored at the beginning to appearing completely gone after 11 hours. The spin-cast sample however yielded some interesting (and somewhat pretty) images (see Figure 24). The four subsequent images were taken after 3.5, 5.4, 11 and 23 hours on the hot plate. The rst two show complete coverage, while the third and forth show some clean areas but still some PMMA spots. This indicates that even after almost a day on

40

Figure 24: Microscope images of SiO2 wafer with evaporated PMMA over time. the hot plate, spin-cast lms do not completely evaporate, while drop-cast lms probably do. Finally, optical images were taken of the transfered graphene onto a SiO2 substrate. A simple camera image shows a visible distinction where the lm was transfered (see Figure 25), but that could be due to residual PMMA as much as to graphene. Two optical microscope images were also taken (see Figures 26 and 27). Figure 26 shows a wrinkled and folded sheet which could either be a large graphene segment or PMMA. In both cases, this indicates that the graphenePMMA layer ripped and folded at some point during the transfer. Figure 27 shows a more common region of the sample. Patchy regions of discoloration match the observed color from exfoliated graphene on silicon dioxide wafers, suggesting that this is more graphene that has torn and folded.

41

Figure 25: Photo of SiO2 wafter after graphene transfer.

Figure 26: Microscope image of SiO2 wafer after graphene transfer.

42

Figure 27: Microscope image of SiO2 wafer after graphene transfer.

Discussion and Future Work

Many methods for observing graphene have conrmed that I have been able to grow large-area graphene lms on copper foils, and I have taken preliminary steps in transferring this graphene to a silicon dioxide substrate. In the future, the growth parameters of the CVD system should be optimized to create the most reliable and uniform graphene lms with the largest domain sizes possible. The transfer procedure should also be investigated. Various techniques can be used to remove PMMA from the nal substrate, for example, an acetone bath and a hot plate. They should be tried and tested, with the end goal of transfer to an organic solar cell in mind. Once successful transfer to a silicon wafer has been achieved, atomic force microscopy should be performed to measure the thickness of the lm. Electronic transport measurements should also be done with a silicon substrate (or on ITO templates used in the production of solar cells at Pomona) using a four-point probe setup to determine the graphenes

43

purity. Lastly, the graphene should be included as a transparent conducting electrode in the production of an organic solar cell and compared to cells made with ITO.

Acknowledgements

I would like to recognize primarily Prof. David Tanenbaum for his interest and dedication to my work over my entire time at Pomona College. He has provided for me opportunity after opportunity to experience real research both at Pomona and at large research centers, and guided me throughout my undergraduate education and beyond. Secondly I wish to acknowledge Matt Hasling for his help running the CVD system and taking SEM images of our copper foils. I wish him the best of luck continuing this research and look forward to working with him for a bit longer before he becomes the student in charge! Jenna deBoisblanc has also been a superb lab colleague and an excellent resource on organic solar cells. Thanks to Desalegne Teweldebrhan and Prof. Alexander Baladins group at UC Riverside, and to Prof. Tyler Moersch in Pomonas chemistry department, for their assistance in taking Raman measurements of my graphene samples. I relied heavily on the work of Dr. Paul McEuen and his group at Cornell University, especially Arend van der Zande. Lastly, there have been many people connected to the Pomona Physics department who have aided me in my work. Glenn Flohr, Tony Grigsby and David Haley assisted me with machining and repairing lab equipment, while

44

Profs. Alfred Kwok and Dwight Whitaker provided me with advice in lieu of my primary advisor. And nally, thanks to Ian Frank and Scott Berkley, now Pomona alumni who were my student mentors during my rst two years in the lab and whose work served as the foundation for my own.

References
[1] Sukang S. Bae. Roll-to-roll production of 30-inch graphene lms for transparent electrodes. Nature nanotechnology, 5(8):574578, 2010. [2] Hctor A. Becerril, Jie Mao, Zunfeng Liu, Randall M. Stoltenberg, Zhenan Bao, and Yongsheng Chen. Evaluation of solution-processed reduced graphene oxide lms as transparent conductors. ACS Nano, 2(3):463470, 2008. http://pubs.acs.org/doi/pdf/10.1021/nn700375n. [3] Scott Berkley, Ian Frank, Arend van der Zande, David Tanenbaum, and Paul McEuen. Mechanical properties of suspended graphene sheets, 2008. [4] Scott Berkley, Arend van der Zande, David Tanenbaum, and Paul McEuen. Gold nanoparticles on graphene surfaces, 2007. [5] Sreekar Bhaviripudi, Xiaoting Jia, Mildred S. Dresselhaus, and Jing Kong. Role of kinetic factors in chemical vapor deposition synthesis of uniform large area graphene using copper catalyst. Nano Letters, 10(10):41284133, 2010. http://pubs.acs.org/doi/pdf/10.1021/nl102355e. [6] J. S. Bunch, Scott S. Verbridge, Jonathan S. Alden, der Zande van, Jeevak M. Parpia, Harold G. Craighead, and Paul L. McEuen. Impermeable atomic membranes from graphene sheets. Nano Letters, 8(8):24582462, 08/01 2008. [7] J. Scott Bunch, Arend M. van der Zande, Scott S. Verbridge, Ian W. Frank, David M. Tanenbaum, Jeevak M. Parpia, Harold G. Craighead, and Paul L. McEuen. Electromechanical resonators from graphene sheets. Science, 2007. [8] P. P. Edwards, A. Porch, M. O. Jones, D. V. Morgan, and R. M. Perks. Basic materials physics of transparent conducting oxides. Dalton Trans., (19):29953002, 2004. [9] A. C. Ferrari, J. C. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri, F. Mauri, S. Piscanec, D. Jiang, K. S. Novoselov, S. Roth, and A. K. Geim. Raman spectrum of graphene and graphene layers. Physical Review Letters, 97(18):187401, 10/30 2006.

45

[10] Phillip N. First, Walt A. De Heer, Thomas Seyller, Claire Berger, Joseph A. Stroscio, and Jeong-Sun Moon. Epitaxial graphenes on silicon carbide. MRS Bulletin, 35(April):135, 2010. [11] Anthony C. Fischer-Cripps. The materials physics companion. Taylor & Francis, New York, 2008. A.C. Fischer-Cripps.; Includes index. [12] The Class for Physics of the Royal Swedish Academy of Sciences. Graphene: Scientic background on the nobel prize in physics 2010, October 5, 2010 2010. [13] Ian Frank. Shaping graphene, an alternative approach. Undergraduate thesis, Pomona College Physics, 5/7 2008. [14] Michael S. Fuhrer. An introduction to graphene electronic structure. University of Maryland, www.physics.umd.edu/mfuhrer/Presentations/GrapheneIntro.pps. [15] A. A. K. Geim. The rise of graphene. Nature materials, 6(3):183191, 2007. [16] Alexander Grneis, Kurt Kummer, and Denis V. Vyalikh. Dynamics of graphene growth on a metal surface: a time-dependent photoemission study. New Journal of Physics, 11(7):073050, 2009. [17] O. S. Heavens. Optical properties of thin solid lms. Dover Publications, 1991. 91031278. [18] Inhwa Jung, Matthew Pelton, Richard Piner, Dmitriy A. Dikin, Sasha Stankovich, Supinda Watcharotone, Martina Hausner, and Rodney S. Ruo. Simple approach for high-contrast optical imaging and characterization of graphene-based sheets. Nano Letters, 7(12):35693575, 2007. http://pubs.acs.org/doi/pdf/10.1021/nl0714177. [19] Keun Soo Kim, Yue Zhao, Houk Jang, Sang Yoon Lee, Jong Min Kim, Kwang S. Kim, Jong-Hyun Ahn, Philip Kim, Jae-Young Choi, and Byung Hee Hong. Large-scale pattern growth of graphene lms for stretchable transparent electrodes. Nature, 457(7230):706710, 02/05 2009. [20] Xuesong Li, Weiwei Cai, Jinho An, Seyoung Kim, Junghyo Nah, Dongxing Yang, Richard Piner, Aruna Velamakanni, Inhwa Jung, Emanuel Tutuc, Sanjay K. Banerjee, Luigi Colombo, and Rodney S. Ruo. Large-area synthesis of high-quality and uniform graphene lms on copper foils, 05/11 2009. [21] Chhowalla M., Teo K.B.K., Ducati C., Rupesinghe N.L., Amaratunga G.A.J., Ferrari A.C., Roy D., Robertson J., and Milne W.I. Growth process conditions of vertically aligned carbon nanotubes using plasma enhanced chemical vapor deposition. Journal of Applied Physics, 90:53085317, nov 2001. Provided by the SAO/NASA Astrophysics Data System.

46

[22] Tadatsugu Minami. Present status of transparent conducting oxide thinlm development for indium-tin-oxide (ito) substitutes. Thin Solid Films, 516(17):58225828, 7/1 2008. [23] E. Moreau, F. J. Ferrer, D. Vignaud, S. Godey, and X. Wallart. Graphene growth by molecular beam epitaxy using a solid carbon source. physica status solidi (a), 207(2):300303, 2010. [24] R. R. Nair, P. Blake, A. N. Grigorenko, K. S. Novoselov, T. J. Booth, T. Stauber, N. M. R. Peres, and A. K. Geim. Fine structure constant denes visual transparency of graphene. Science, 2008. [25] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos, and A. A. Firsov. Two-dimensional gas of massless dirac fermions in graphene. Nature, 438(7065):197200, 11/10 2005. M3: 10.1038/nature04233; 10.1038/nature04233. [26] Phi Pham, Arend van der Zande, Xiaodong Xu, and Paul McEuen. Chemical vapor deposition growth of graphene, 2009. Cornell CNS REU Presentation. [27] Dhiraj Prasai, Kirill Bolotin, Juan Tuberquia, Robert Harl, and Kane Jennings. Graphene: Atomically thin protective coating. In Session A30: Graphene: Growth, Properties and Devices, pages Volume 56, Number 1. American Physical Society March Meeting, March 21 2011. [28] Heyrovska R. Atomic structures of graphene, benzene and methane with bond lengths as sums of the single, double and resonance bond radii of carbon. ArXiv e-prints, apr 2008. 0804.4086; Provided by the SAO/NASA Astrophysics Data System. [29] Christian Schonenberger. Bandstructure of Graphene and Carbon Nanotubes: An Exercise in Condensed Matter Physics. [30] Arend van der Zande and Paul McEuen. Graphene fabrication, mceuen group wiki. http://www.lassp.cornell.edu/lassp_ data/mceuen/homepage/dokuwiki-2008-05-05/doku.php?id=wiki: graphene_fabrication.

47

Das könnte Ihnen auch gefallen