Jump to content

Graphene

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by Leadwind (talk | contribs) at 00:55, 23 April 2014 (topic sentence for last paragraph, simpler sentence structure and wording). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Graphene is an atomic-scale honeycomb lattice made of carbon atoms.

Graphene is pure carbon in the form of a very thin, nearly transparent sheet, basically one atom thick. It is remarkably strong for its very low weight, and it conducts heat and electricity with great efficiency. While scientists had theorized about graphene for decades, it was first produced in the lab in 2004.[1] Because it is virtually two-dimensional, it interacts oddly with light and with other materials. Researchers have identified the bipolar transistor effect, ballistic transport of charges and large quantum oscillations.

Technically, graphene is a crystalline allotrope of carbon with 2-dimensional properties. In graphene, carbon atoms are densely packed in a regular sp2-bonded atomic-scale chicken wire (hexagonal) pattern. Graphene can be described as a one-atom thick layer of graphite. It is the basic structural element of other allotropes, including graphite, charcoal, carbon nanotubes and fullerenes. It can also be considered as an indefinitely large aromatic molecule, the limiting case of the family of flat polycyclic aromatic hydrocarbons.

Graphene research has expanded quickly since the substance was first isolated in 2004. Research was informed by theoretical descriptions of graphene's composition, structure and properties, which had all been calculated decades earlier. High-quality graphene also proved to be surprisingly easy to isolate, making more research possible. Andre Geim and Konstantin Novoselov at the University of Manchester won the Nobel Prize in Physics in 2010 "for groundbreaking experiments regarding the two-dimensional material graphene".[2]

Definition

"Graphene" is a combination of graphite and the suffix -ene, named by Hanns-Peter Boehm,[3] who described single-layer carbon foils in 1962.[4]

The term graphene first appeared in 1987[5] to describe single sheets of graphite as a constituent of graphite intercalation compounds (GICs); conceptually a GIC is a crystalline salt of the intercalant and graphene. The term was also used in early descriptions of carbon nanotubes,[6] as well as for epitaxial graphene[7] and polycyclic aromatic hydrocarbons.[8]

The IUPAC compendium of technology states: "previously, descriptions such as graphite layers, carbon layers, or carbon sheets have been used for the term graphene... it is incorrect to use for a single layer a term which includes the term graphite, which would imply a three-dimensional structure. The term graphene should be used only when the reactions, structural relations or other properties of individual layers are discussed."[9]

Geim defined "isolated or free-standing graphene" as "graphene is a single atomic plane of graphite, which – and this is essential – is sufficiently isolated from its environment to be considered free-standing."[10] This definition is narrower than the IUPAC definition and refers to cleaved, transferred and suspended graphene.[citation needed] Other forms of graphene, such as graphene grown on various metals, can become free-standing if, for example, suspended or transferred to silicon dioxide (SiO
2
) or silicon carbide (after its passivation with hydrogen).[11]

Graphene can be considered an "infinite alternant" (only six-member carbon ring) polycyclic aromatic hydrocarbon (PAH).[12]

History

In 1859 Benjamin Collins Brodie was aware of the highly lamellar structure of thermally reduced graphite oxide.[13]

The structure of graphite was solved in 1916.[14] by the related method of powder diffraction,[15] It was studied in detail by V. Kohlschütter and P. Haenni in 1918, who also described the properties of graphite oxide paper.[16] Its structure was determined from single-crystal diffraction in 1924.[17]

The theory of graphene was first explored by P. R. Wallace in 1947 as a starting point for understanding the electronic properties of 3D graphite. The emergent massless Dirac equation was first pointed out by Gordon Walter Semenoff and David P. DeVincenzo and Eugene J. Mele.[18] Semenoff emphasized the occurrence in a magnetic field of an electronic Landau level precisely at the Dirac point. This level is responsible for the anomalous integer quantum Hall effect.[19][20][21]

The earliest TEM images of few-layer graphite were published by G. Ruess and F. Vogt in 1948.[22] Later, single graphene layers were also observed directly by electron microscopy.[23] Before 2004 intercalated graphite compounds were studied under a transmission electron microscope (TEM). Researchers occasionally observed thin graphitic flakes ("few-layer graphene") and possibly even individual layers. An early, detailed study on few-layer graphite dates to 1962.[24][25]

Starting in the 1970s single layers of graphite were grown epitaxially on top of other materials.[26] This "epitaxial graphene" consists of a single-atom-thick hexagonal lattice of sp2-bonded carbon atoms, as in free-standing graphene. However, there is significant charge transfer from the substrate to the epitaxial graphene, and, in some cases, hybridization between the d-orbitals of the substrate atoms and π orbitals of graphene, which significantly alters the electronic structure of epitaxial graphene.

Single layers of graphite were also observed by transmission electron microscopy within bulk materials, in particular inside soot obtained by chemical exfoliation. Efforts to make thin films of graphite by mechanical exfoliation started in 1990,[27] but nothing thinner than 50 to 100 layers was produced before 2004.

A lump of graphite, a graphene transistor and a tape dispenser. Donated to the Nobel Museum in Stockholm by Andre Geim and Konstantin Novoselov in 2010.

Initial attempts to make atomically thin graphitic films employed exfoliation techniques similar to the drawing method. Multilayer samples down to 10 nm in thickness were obtained.[28] Old papers were unearthed[24] in which researchers tried to isolate graphene starting with intercalated compounds. These papers reported the observation of very thin graphitic fragments (possibly monolayers) by transmission electron microscopy. Neither of the earlier observations was sufficient to "spark the graphene gold rush", which awaited macroscopic samples of extracted atomic planes.

One of the very first patents pertaining to the production of graphene was filed in October, 2002 (US Pat. 7071258).[29] Entitled, "Nano-scaled Graphene Plates", this patent detailed one of the very first large scale graphene production processes. Two years later, in 2004 Andre Geim and Kostya Novoselov at University of Manchester extracted single-atom-thick crystallites from bulk graphite.[30] They pulled graphene layers from graphite and transferred them onto thin SiO
2
on a silicon wafer in a process called either micromechanical cleavage or the Scotch tape technique. The SiO
2
electrically isolated the graphene and weakly interacted with it, providing nearly charge-neutral graphene layers. The silicon beneath the SiO
2
could be used as a "back gate" electrode to vary the charge density in the graphene over a wide range. They may not have been the first to use this technique— US 6667100 , filed in 2002, describes how to process commercially available flexible expanded graphite to achieve a graphite thickness of 0.01 thousandth of an inch. The key to success was high-throughput visual recognition of graphene on a properly chosen substrate, which provides a small but noticeable optical contrast.

The cleavage technique led directly to the first observation of the anomalous quantum Hall effect in graphene,[19][21] which provided direct evidence of graphene's theoretically predicted Berry's phase of massless Dirac fermions. The effect was reported soon after by Philip Kim and Yuanbo Zhang in 2005. These experiments started after the researchers observed colleagues who were looking for the quantum Hall effect[31] and Dirac fermions[32] in bulk graphite.

Even though graphene on nickel and on silicon carbide have both existed in the laboratory for decades, graphene mechanically exfoliated on SiO
2
provided the first proof of the Dirac fermion nature of electrons.[citation needed]

Andre Geim and Konstantin Novoselov, 2010

Geim and Novoselov received several awards for their pioneering research on graphene, notably the 2010 Nobel Prize in Physics.[33]

Properties

Structure

The atomic structure of isolated, single-layer graphene was studied by transmission electron microscopy (TEM) on sheets of graphene suspended between bars of a metallic grid.[23] Electron diffraction patterns showed the expected honeycomb lattice. Suspended graphene also showed "rippling" of the flat sheet, with amplitude of about one nanometer. These ripples may be intrinsic to the material as a result of the instability of two-dimensional crystals,[28][34][35] or may originate from the ubiquitous dirt seen in all TEM images of graphene. Atomic resolution real-space images of isolated, single-layer graphene on SiO
2
substrates are available[36][37] via scanning tunneling microscopy. Photoresist residue, which must be removed to obtain atomic-resolution images, may be the "adsorbates" observed in TEM images, and may explain the observed rippling. Rippling on SiO
2
is caused by conformation of graphene to the underlying SiO
2
, and is not intrinsic.[36]

Graphene sheets in solid form usually show evidence in diffraction for graphite's (002) layering. This is true of some single-walled nanostructures.[38] However, unlayered graphene with only (hk0) rings has been found in the core of presolar graphite onions.[39] TEM studies show faceting at defects in flat graphene sheets[40] and suggest a role for two-dimensional crystallization from a melt.

Graphene can self-repair holes in its sheets, when exposed to molecules containing carbon, such as hydrocarbons. Bombarded with pure carbon atoms, the atoms perfectly align into hexagons, completely filling the holes.[41][42]

Chemical

Graphene is the only form of carbon (and generally all solid materials) in which each single atom is in exposure for chemical reaction from two sides (due to the 2D structure). It is known that carbon atoms at the edge of graphene sheets have special chemical reactivity, and graphene has the highest ratio of edgy carbons (in comparison with similar materials such as carbon nanotubes). In addition, various types of defects within the sheet, which are very common, increase the chemical reactivity.[43] The onset temperature of reaction between the basal plane of single-layer graphene and oxygen gas is below 260 °C [44] and the graphene burns at very low temperature (e.g., 350 °C).[45] In fact, graphene is chemically the most reactive form of carbon, owing to the lateral availability of carbon atoms. Graphene is commonly modified with oxygen- and nitrogen-containing functional groups and analyzed by infrared spectroscopy and X-ray photoelectron spectroscopy. But, determination of structures of graphene with oxygen-[46] and nitrogen-[47] containing functional groups is a difficult task unless the structures are well controlled.

In 2013, Stanford University physicists reported that sheets of Graphene one atom thick are a hundred times more chemically reactive than thicker sheets.[48]

Physical

The carbon–carbon bond length in graphene is about 0.142 nanometers.[49] Graphene sheets stack to form graphite with an interplanar spacing of 0.335 nm.

first created with a pencil

Electronic

GNR band structure for zig-zag orientation. Tightbinding calculations show that zigzag orientation is always metallic.
GNR band structure for armchair orientation. Tightbinding calculations show that armchair orientation can be semiconducting or metallic depending on width (chirality).

Graphene is a semi-metal or zero-gap semiconductor. Four electronic properties separate it from other condensed matter systems.

Electronic spectrum

Electrons propagating through graphene's honeycomb lattice effectively lose their mass, producing quasi-particles that are described by a 2D analogue of the Dirac equation rather than the Schrödinger equation for spin-1/2 particles.[50][51]

Using a conventional tight-binding model the dispersion relation produces energy of the electrons with wave vector k is[52][53]

with the nearest-neighbor hopping energy γ02.8 eV and the lattice constant a ≈ 2.46 Å. The conduction and valence bands, respectively, correspond to the different signs; they touch each other at six points, the "K-values" of the two-dimensional hexagonal Brillouin zone. Two of these six points are independent, while the rest are equivalent by symmetry. In the vicinity of the K-points the energy depends linearly on the wave vector, similar to a relativistic particle.[52][54] Since an elementary cell of the lattice has a basis of two atoms, the wave function has an effective 2-spinor structure.

As a consequence, at low energies, even neglecting the true spin, the electrons can be described by an equation that is formally equivalent to the massless Dirac equation. Hence, the electrons and holes are called Dirac fermions and the six corners are called the Dirac points.[52] This pseudo-relativistic description is restricted to the chiral limit, i.e., to vanishing rest mass M0, which leads to interesting additional features:[52][55]

Here vF ~ 106 m/s (.003 c) is the Fermi velocity in graphene, which replaces the velocity of light in the Dirac theory; is the vector of the Pauli matrices, is the two-component wave function of the electrons, and E is their energy.[50]

The equation describing the electrons' linear dispersion relation is

;

where the wavevector k is measured from the Dirac points (the zero of energy is chosen here to coincide with the Dirac points). The equation uses a pseudospin matrix formula that describes two sublattices of the honeycomb lattice.[54]

Single-atom wave propagation

Electron waves in graphene propagate within a single-atom layer, making them sensitive to the proximity of other materials such as high-κ dielectrics, superconductors and ferromagnetics.

Conductance

Graphene electrons can cover micrometer distances without scattering, even at room temperature.[56]

Electrical resistance in 40-nanometer-wide nanoribbons of epitaxially-created graphene changes in discrete steps following quantum mechanical principles. Electrons travel ballistically, similar to those observed in cylindrical carbon nanotubes. The ribbons' theoretical conductance exceeds predictions by a factor of 10. Electrons can move tens or hundreds of microns without scattering. The ribbons can act more like optical waveguides or quantum dots, allowing electrons to flow smoothly along the ribbons' edges. By contrast, in conductors such as copper, resistance increases in proportion to the length as electrons encounter impurities while moving through the conductor.[57][58]

Transport is dominated by two modes. One is ballistic and temperature independent, while The other is thermally activated. Transport is protected from back-scattering, possibly reflecting ground-state properties of neutral graphene. At room temperature, the resistance of both modes is found to increase abruptly at a particular length—the ballistic mode at 16 micrometres and the other at 160 nanometres (1% of the former length).[57]

The ribbons were grown on the edges of three-dimensional structures etched into silicon carbide wafers. When the wafers are heated to approximately 1,000 °C (1,830 °F), silicon is preferentially driven off along the edges, forming nanoribbons whose structure is determined by the pattern of the three-dimensional surface. The nanoribbons had perfectly smooth edges, annealed by the fabrication process. Electron mobility measurements surpassing one million correspond to a sheet resistance of one ohm per square— two orders of magnitude lower than in two-dimensional graphene.[57]

Anomalous quantum Hall effect

The quantum Hall effect is a quantum mechanical version of the Hall effect, which is the production of transverse (perpendicular to the main current) conductivity in the presence of a magnetic field. The quantization of the Hall effect at integer multiples (the "Landau level") of the basic quantity (where e is the elementary electric charge and h is Planck's constant) It can usually be observed only in very clean silicon or gallium arsenide solids at temperatures around 3 K and very high magnetic fields.

Graphene shows the quantum Hall effect with respect to conductivity-quantization: the effect is anomalous in that the sequence of steps is shifted by 1/2 with respect to the standard sequence and with an additional factor of 4. Graphene's Hall conductivity is , where N is the Landau level and the double valley and double spin degeneracies give the factor of 4.[28] These anomalies are present at room temperature, i.e. at roughly 20 °C.[19]

This behavior is a direct result of graphene's massless Dirac electrons. In a magnetic field, their spectrum has a Landau level with energy precisely at the Dirac point. This level is a consequence of the Atiyah–Singer index theorem and is half-filled in neutral graphene,[52] leading to the "+1/2" in the Hall conductivity.[20] Bilayer graphene also shows the quantum Hall effect, but with only one of the two anomalies (i.e. ). In the second anomaly, the first plateau at N=0 is absent, indicating that bilayer graphene stays metallic at the neutrality point.[28]

Unlike normal metals, graphene's longitudinal esistance shows maxima rather than minima for integral values of the Landau filling factor in measurements of the Shubnikov–De Haas oscillations, whereby the term integral quantum Hall effect. These oscillations ahow a phase shift of π, known as Berry’s phase.[19][51] Berry’s phase arises due to the zero effective carrier mass near the Dirac points.[21] The temperature dependence of the oscillations reveals that the carriers have a non-zero cyclotron mass, despite their zero effective mass.[19]

Graphene samples prepared on nickel films, and on both the silicon face and carbon face of silicon carbide, show the anomalous effect directly in electrical measurements.[59][60][61][62][63][64] Graphitic layers on the carbon face of silicon carbide show a clear Dirac spectrum in angle-resolved photoemission experiments, and the effect is observed in cyclotron resonance and tunneling experiments.[65]

Strong magnetic fields

In magnetic fields above 10 Teslas or so additional plateaus of the Hall conductivity at with are observed.[66] A plateau at [67] and the fractional quantum Hall effect at were also reported.[67][68]

These observations with indicate that the four-fold degeneracy (two valley and two spin degrees of freedom) of the Landau energy levels is partially or completely lifted.

Casimir effect

The Casimir effect is an interaction between disjoint neutral bodies provoked by the fluctuations of the electrodynamical vacuum. Mathematically it can be explained by considering the normal modes of electromagnetic fields, which explicitly depend on the boundary (or matching) conditions on the interacting bodies' surfaces. Since graphene/electromagnetic field interaction is strong for a one-atom-thick material, the Casimir effect is of growing interest.[69][70]

van der Waals force

The van der Waals force (or dispersion force) is also unusual, obeying an inverse cubic, asymptotic power law in contrast to the usual inverse quartic.[71]

"Massive" electrons

Graphene's unit cell has two identical carbon atoms and two zero-energy states: one in which the electron resides on atom A, the other in which the electron resides on atom B. Both states exist at exactly zero energy. However, if the two atoms in the unit cell are not identical, the situation changes. Hunt et al. show that placing hBN in contact with graphene can alter the potential felt at atom A versus atom B enough that the electrons develop a mass and accompanying band gap of about 30 meV.[72]

The mass can be positive or negative. An arrangement that slightly raises the energy of an electron on atom A relative to atom B gives it a positive mass, while an arrangement that raises the energy of atom B produces a negative electron mass. The two versions behave alike and are indistinguishable via optical spectroscopy. An electron traveling from a positive-mass region to a negative-mass region must cross an intermediate region where its mass once again becomes zero. This region is gapless and therefore metallic. Metallic modes bounding semiconducting regions of opposite-sign mass is a hallmark of a topological phase and display much the same physics as topological insulators.[72]

If the mass in graphene can be controlled, electrons can be confined to massless regions by surrounding them with massive regions, allowing the patterning of quantum dots, wires, and other mesoscopic structures. It also produces one-dimensional conductors along the boundary. These wires would be protected against backscattering and could carry currents without dissipation.[72]

Electron transport

Experimental results from transport measurements show that graphene has a remarkably high electron mobility at room temperature, with reported values in excess of 15,000 cm2·V−1·s−1.[28] Additionally, the symmetry of the experimentally measured conductance indicates that hole and electron mobilities should be nearly the same.[51] The mobility is nearly independent of temperature between 10 K and 100 K,[19][73][74] which implies that the dominant scattering mechanism is defect scattering. Scattering by the acoustic phonons of graphene intrinsically limits room temperature mobility to 200,000 cm2·V−1·s−1 at a carrier density of 1012 cm−2.[74][75] The corresponding resistivity of the graphene sheet would be 10−6 Ω·cm. This is less than the resistivity of silver, the lowest known at room temperature.[76] However, for graphene on SiO
2
substrates, scattering of electrons by optical phonons of the substrate is a larger effect at room temperature than scattering by graphene’s own phonons. This limits mobility to 40,000 cm2·V−1·s−1.[74]

Despite zero carrier density near the Dirac points, graphene exhibits a minimum conductivity on the order of . The origin of this minimum conductivity is still unclear. However, rippling of the graphene sheet or ionized impurities in the SiO
2
substrate may lead to local puddles of carriers that allow conduction.[51] Several theories suggest that the minimum conductivity should be ; however, most measurements are of order or greater[28] and depend on impurity concentration.[77]

Graphene doped with various gaseous species (both acceptors and donors) can be returned to an undoped state by gentle heating in vacuum.[77][78] Even for dopant concentrations in excess of 1012 cm−2 carrier mobility exhibits no observable change.[78] Graphene doped with potassium in ultra-high vacuum at low temperature can reduce mobility 20-fold.[77][79] The mobility reduction is reversible on heating the graphene to remove the potassium.

Due to graphene's two dimensions, charge fractionalization (where the apparent charge of individual pseudoparticles in low-dimensional systems is less than a single quantum[80]) is thought to occur. It may therefore be a suitable material for constructing quantum computers[81] using anyonic circuits.[82]

Optical

Photograph of graphene in transmitted light. This one-atom-thick crystal can be seen with the naked eye because it absorbs approximately 2.3% of white light.

Graphene's unique optical properties produce an unexpectedly high opacity for an atomic monolayer in vacuum, absorbing πα ≈ 2.3% of white light, where α is the fine-structure constant.[83] This is a consequence of the "unusual low-energy electronic structure of monolayer graphene that features electron and hole conical bands meeting each other at the Dirac point... [which] is qualitatively different from more common quadratic massive bands".[84] Based on the Slonczewski–Weiss–McClure (SWMcC) band model of graphite, the interatomic distance, hopping value and frequency cancel when optical conductance is calculated using Fresnel equations in the thin-film limit.

Although confirmed experimentally, the measurement is not precise enough to improve on other techniques for determining the fine-structure constant.[85]

Graphene's band gap can be tuned from 0 to 0.25 eV (about 5 micrometre wavelength) by applying voltage to a dual-gate bilayer graphene field-effect transistor (FET) at room temperature.[86] The optical response of graphene nanoribbons is tunable into the terahertz regime by an applied magnetic field.[87] Graphene/graphene oxide systems exhibit electrochromic behavior, allowing tuning of both linear and ultrafast optical properties.[88]

A graphene-based Bragg grating (one-dimensional photonic crystal) has been fabricated and demonstrated its capability for excitation of surface electromagnetic waves in the periodic structure by using 633 nm He–Ne laser as the light source.[89]

Saturable absorption

Such unique absorption could become saturated when the input optical intensity is above a threshold value. This nonlinear optical behavior is termed saturable absorption and the threshold value is called the saturation fluence. Graphene can be saturated readily under strong excitation over the visible to near-infrared region, due to the universal optical absorption and zero band gap. This has relevance for the mode locking of fiber lasers, where fullband mode locking has been achieved by graphene-based saturable absorber. Due to this special property, graphene has wide application in ultrafast photonics. Moreover, the optical response of graphene/graphene oxide layers can be tuned electrically.[88][90] Saturable absorption in graphene could occur at the Microwave and Terahertz band, owing to its wideband optical absorption property. The microwave saturable absorption in graphene demonstrates the possibility of graphene microwave and terahertz photonics devices, such as microwave saturable absorber, modulator, polarizer, microwave signal processing and broad-band wireless access networks.[91]

Nonlinear Kerr effect

Under more intensive laser illumination, graphene could also possess a nonlinear phase shift due to the optical nonlinear Kerr effect. Based on a typical open and close aperture z-scan measurement, graphene possesses a giant non-linear Kerr coefficient of 10−7 cm2·W−1, almost nine orders of magnitude larger than that of bulk dielectrics.[92] This suggests that graphene may be a nonlinear Kerr medium, paving the way for graphene-based nonlinear Kerr photonics such as a soliton.

Excitonic

First-principle calculations with quasiparticle corrections and many-body effects are performed to study the electronic and optical properties of graphene-based materials. The approach is described as three stages.[93] With GW calculation, the properties of graphene-based materials are accurately investigated, including graphene,[94] graphene nanoribbons,[95] edge and surface functionalized armchair graphene nanoribbons,[96] hydrogen saturated armchair graphene nanoribbons,[97] Josephson effect in graphene SNS junctions with single localized defect[98] and scaling properties in armchair graphene nanoribbons.[99]

Thermal

Stability

Ab initio calculations show that a graphene sheet is thermodynamically unstable if its size is less than about 20 nm ("graphene is the least stable structure until about 6000 atoms") and becomes the most stable fullerene (as within graphite) only for molecules larger than 24,000 atoms.[100]

Conductivity

The near-room temperature thermal conductivity of graphene was measured to be between (4.84±0.44) × 103 to (5.30±0.48) × 103 W·m−1·K−1. These measurements, made by a non-contact optical technique, are in excess of those measured for carbon nanotubes or diamonds. The isotopic composition, the ratio of 12C to 13C, has a significant impact on thermal conductivity, where isotopically pure 12C graphene has higher conductivity than either a 50:50 isotope ratio or the naturally occurring 99:1 ratio.[101] It can be shown by using the Wiedemann–Franz law, that the thermal conduction is phonon-dominated.[102] However, for a gated graphene strip, an applied gate bias causing a Fermi energy shift much larger than kBT can cause the electronic contribution to increase and dominate over the phonon contribution at low temperatures. The ballistic thermal conductance of graphene is isotropic.[103][104]

Potential for this high conductivity can be seen by considering graphite, a 3D version of graphene that has basal plane thermal conductivity of over a 1000 W·m−1·K−1 (comparable to diamond). In graphite, the c-axis (out of plane) thermal conductivity is over a factor of ~100 smaller due to the weak binding forces between basal planes as well as the larger lattice spacing.[105] In addition, the ballistic thermal conductance of graphene is shown to give the lower limit of the ballistic thermal conductances, per unit circumference, length of carbon nanotubes.[106]

Despite its 2-D nature, graphene has 3 acoustic phonon modes. The two in-plane modes (LA, TA) have a linear dispersion relation, whereas the out of plane mode (ZA) has a quadratic dispersion relation. Due to this, the T2 dependent thermal conductivity contribution of the linear modes is dominated at low temperatures by the T1.5 contribution of the out of plane mode.[106] Some graphene phonon bands display negative Grüneisen parameters.[107] At low temperatures (where most optical modes with positive Grüneisen parameters are still not excited) the contribution from the negative Grüneisen parameters will be dominant and thermal expansion coefficient (which is directly proportional to Grüneisen parameters) negative. The lowest negative Grüneisen parameters correspond to the lowest transversal acoustic ZA modes. Phonon frequencies for such modes increase with the in-plane lattice parameter since atoms in the layer upon stretching will be less free to move in the z direction. This is similar to the behavior of a string, which, when it is stretched, will have vibrations of smaller amplitude and higher frequency. This phenomenon, named "membrane effect", was predicted by Lifshitz in 1952.[108]

Mechanical

The flat graphene sheet is unstable with respect to scrolling i.e. bending into a cylindrical shape, which is its lower-energy state.[109]

As of 2009, graphene appeared to be one of the strongest materials known with a breaking strength over 100 times greater than a hypothetical steel film of the same (thin) thickness,[110] with a Young's modulus (stiffness) of 1 TPa (150,000,000 psi).[111] The Nobel announcement illustrated this by saying that a 1 square meter graphene hammock would support a 4 kg cat but would weigh only as much as one of the cat's whiskers, at 0.77 mg (about 0.001% of the weight of 1 m2 of paper).[110]

However, the process of separating it from graphite, where it occurs naturally, requires technological development to be economical enough to be used in industrial processes.[112][113]

The spring constant of suspended graphene sheets has been measured using an atomic force microscope (AFM). Graphene sheets, held together by van der Waals forces, were suspended over SiO
2
cavities where an AFM tip was probed to test its mechanical properties. Its spring constant was in the range 1–5 N/m and the stiffness was 0.5 TPa, which differs from that of bulk graphite. These high values make graphene very strong and rigid. These intrinsic properties could lead to using graphene for NEMS applications such as pressure sensors and resonators.[114]

As is true of all materials, regions of graphene are subject to thermal and quantum fluctuations in relative displacement. Although the amplitude of these fluctuations is bounded in 3D structures (even in the limit of infinite size), the Mermin–Wagner theorem shows that the amplitude of long-wavelength fluctuations grows logarithmically with the scale of a 2D structure, and would therefore be unbounded in structures of infinite size. Local deformation and elastic strain are negligibly affected by this long-range divergence in relative displacement. It is believed that a sufficiently large 2D structure, in the absence of applied lateral tension, will bend and crumple to form a fluctuating 3D structure. Researchers have observed ripples in suspended layers of graphene,[23] and it has been proposed that the ripples are caused by thermal fluctuations in the material. As a consequence of these dynamical deformations, it is debatable whether graphene is truly a 2D structure.[28][34][35][115]

Spin transport

Graphene is claimed to be an ideal material for spintronics due to its small spin-orbit interaction and the near absence of nuclear magnetic moments in carbon (as well as a weak hyperfine interaction). Electrical spin current injection and detection has been demonstrated up to room temperature.[116][117][118] Spin coherence length above 1 micrometre at room temperature was observed,[116] and control of the spin current polarity with an electrical gate was observed at low temperature.[117]

Strong magnetic fields

Graphene's quantum Hall effect in magnetic fields above 10 Teslas or so reveals additional interesting features. Additional plateaus of the Hall conductivity at with are observed.[66] Also, the observation of a plateau at [67] and the fractional quantum Hall effect at were reported.[67][68]

These observations with indicate that the four-fold degeneracy (two valley and two spin degrees of freedom) of the Landau energy levels is partially or completely lifted. One hypothesis is that the magnetic catalysis of symmetry breaking is responsible for lifting the degeneracy.[citation needed]

Forms

Nanostripes

Graphene nanoribbons ("nanostripes" in the "zig-zag" orientation), at low temperatures, show spin-polarized metallic edge currents, which also suggests applications in the new field of spintronics. (In the "armchair" orientation, the edges behave like semiconductors.[50])

Oxide

Using paper-making techniques on dispersed, oxidized and chemically processed graphite in water, the monolayer flakes form a single sheet and create strong bonds. These sheets, called graphene oxide paper have a measured tensile modulus of 32 GPa.[119] The chemical property of graphite oxide is related to the functional groups attached to graphene sheets. These can change the polymerization pathway and similar chemical processes.[120] Graphene oxide flakes in polymers display enhanced photo-conducting properties.[121] Graphene-based membranes are impermeable to all gases and liquids (vacuum-tight). However, water evaporates through them as quickly as if the membrane was not present.[122]

Chemical modification

Photograph of single-layer graphene oxide undergoing high temperature chemical treatment, resulting in sheet folding and loss of carboxylic functionality, or through room temperature carbodiimide treatment, collapsing into star-like clusters.

Soluble fragments of graphene can be prepared in the laboratory[123] through chemical modification of graphite. First, microcrystalline graphite is treated with an acidic mixture of sulfuric acid and nitric acid. A series of oxidation and exfoliation steps produce small graphene plates with carboxyl groups at their edges. These are converted to acid chloride groups by treatment with thionyl chloride; next, they are converted to the corresponding graphene amide via treatment with octadecylamine. The resulting material (circular graphene layers of 5.3 angstrom thickness) is soluble in tetrahydrofuran, tetrachloromethane and dichloroethane.

Refluxing single-layer graphene oxide (SLGO) in solvents leads to size reduction and folding of individual sheets as well as loss of carboxylic group functionality, by up to 20%, indicating thermal instabilities of SLGO sheets dependant on their preparation methodology. When using thionyl chloride, acyl chloride groups result, which can then form aliphatic and aromatic amides with a reactivity conversion of around 70–80%.

Boehm titration results for various chemical reactions of single-layer graphene oxide, which reveal reactivity of the carboxylic groups and the resultant stability of the SLGO sheets after treatment.

Hydrazine reflux is commonly used for reducing SLGO to SLG(R), but titrations show that only around 20–30% of the carboxylic groups are lost, leaving a significant number available for chemical attachment. Analysis of SLG(R) generated by this route reveals that the system is unstable and using a room temperature stirring with HCl (< 1.0 M) leads to around 60% loss of COOH functionality. Room temperature treatment of SLGO with carbodiimides leads to the collapse of the individual sheets into star-like clusters that exhibited poor subsequent reactivity with amines (ca. 3–5% conversion of the intermediate to the final amide).[124] It is apparent that conventional chemical treatment of carboxylic groups on SLGO generates morphological changes of individual sheets that leads to a reduction in chemical reactivity, which may potentially limit their use in composite synthesis. Therefore, chemical reactions types have been explored. SLGO has also been grafted with polyallylamine, cross-linked through epoxy groups. When filtered into graphene oxide paper, these composites exhibit increased stiffness and strength relative to unmodified graphene oxide paper.[125]

Full hydrogenation from both sides of graphene sheet results in graphane, but partial hydrogenation leads to hydrogenated graphene.[126] Similarly, both-side fluorination of graphene (or chemical and mechanical exfoliation of graphite fluoride) leads to fluorographene (graphene fluoride), while partial fluorination (generally halogenation) provides fluorinated (halogenated) graphene.

Bilayer

Bilayer graphene displays the anomalous quantum Hall effect, a tunable band gap[127] and potential for excitonic condensation[128] –making it a promising candidate for optoelectronic and nanoelectronic applications. Bilayer graphene typically can be found either in twisted configurations where the two layers are rotated relative to each other or graphitic Bernal stacked configurations where half the atoms in one layer lie atop half the atoms in the other. Stacking order and orientation govern the optical and electronic properties of bilayer graphene.

One way to synthesize bilayer graphene is via chemical vapor deposition, which can produce large bilayer regions that almost exclusively conform to a Bernal stack geometry.[129]

Fiber

In 2011, Xinming Li and Hongwei Zhu from Tsinghua University reported a novel yet simple approach to fabricate graphene fibers from chemical vapor deposition grown graphene films.[130] The method was scalable and controllable, delivering tunable morphology and pore structure by controlling the evaporation of solvents with suitable surface tension. Flexible all-solid-state supercapacitors based on this graphene fibers were demonstrated in 2013.[131]

3D

In 2013, a three-dimensional honeycomb of hexagonally arranged carbon was termed 3D graphene, although self-supporting 3D graphene has not yet been produced.[132]

Production techniques

Isolated 2D crystals cannot be grown via chemical synthesis beyond small sizes even in principle, because the rapid growth of phonon density with increasing lateral size forces 2D crystallites to bend into the third dimension.[10] However, other routes to 2d materials exist:

Fundamental forces place seemingly insurmountable barriers in the way of creating [2D crystals]... The nascent 2D crystallites try to minimize their surface energy and inevitably morph into one of the rich variety of stable 3D structures that occur in soot. But there is a way around the problem. Interactions with 3D structures stabilize 2D crystals during growth. So one can make 2D crystals sandwiched between or placed on top of the atomic planes of a bulk crystal. In that respect, graphene already exists within graphite... One can then hope to fool Nature and extract single-atom-thick crystallites at a low enough temperature that they remain in the quenched state prescribed by the original higher-temperature 3D growth.[133]

The two basic approaches to producing graphene are to cleave multi-layer graphite into single layers or to grow it epitaxially by depositing one layer of carbon onto another material. The former was developed first, using adhesive tape to peel monolayers away. In either case, the graphite must then be bonded to some substrate to retain its 2d shape.[10] Other techniques have also been developed.

Exfoliation

Adhesive tape

Cleavage is also known as exfoliation. Achieving single layers typically requires multiple exfoliation steps, each producing a slice with fewer layers, until only one remains. Geim and Novosolev used adhesive tape to split their graphene.

After exfoliation the flakes are deposited on a silicon wafer using "dry deposition". Crystallites larger than 1 mm and visible to the naked eye can be obtained with the technique. It is often referred to as a "scotch tape" or "drawing" method. The latter name appeared because the dry deposition resembles drawing with a piece of graphite.[133]

Reduction of graphite oxide

Graphite oxide reduction was probably the first method of graphene synthesis. P. Boehm reported producing monolayer flakes of reduced graphene oxide in 1962.[134] Geim acknowledged Boehm's contribution.[135] Rapid heating of graphite oxide and exfoliation yields highly dispersed carbon powder with a few percent of graphene flakes. Reduction of graphite oxide monolayer films, e.g. by hydrazine with annealing in argon/hydrogen, was reported to yield graphene films. However, the quality is lower compared to scotch-tape graphene, due to incomplete removal of functional groups. Furthermore, the oxidation protocol introduces permanent defects due to over-oxidation. The oxidation protocol was enhanced to yield graphene oxide with an almost intact carbon framework that allows efficient removal of functional groups. The measured charge carrier mobility exceeded 1,000 centimetres (393.70 in)/Vs.[136] Spectroscopic analysis of reduced graphene oxide has been conducted.[137][138]

Sonication

Applying a layer of graphite oxide film to a DVD and burning it in a DVD writer produced a thin graphene film with high electrical conductivity (1738 siemens per meter) and specific surface area (1520 square meters per gram) that was highly resistant and malleable.[139]

Solvent-aided

Dispersing graphite in a proper liquid medium can produce graphene by sonication. Graphite is separated from graphene by centrifugation,[140] producing graphene concentrations initially up to 0.01 mg/ml in N-methylpyrrolidone (NMP) and later to 2.1 mg/ml in NMP,.[141] Using a suitable ionic liquid as the dispersing liquid medium produced concentrations of 5.33 mg/ml.[142] Graphene concentration produced by this method is very low, because nothing prevents the sheets from restacking due to van der Waals forces. The maximum concentrations achieved are the points at which the van der Waals forces overcome the interactive forces between the graphene sheets and the solvent molecules.

Solvent/surfactant-aided

Adding a surfactant to a solvent prior to sonication prevents restacking by adsorbing to the graphene's surface. This produces a higher graphene concentration, but removing the surfactant requires chemical treatments.[citation needed]

Immiscible liquids

Sonicating graphite at the interface of two immiscible liquids, most notably heptane and water, producing macro-scale graphene films. The graphene sheets are adsorbed to the high energy interface between the heptane and the water, where they are kept from restacking. The graphene remains at the interface even when exposed to force in excess of 300,000 g. The solvents may then be evaporated. The sheets are up to ~95% transparent and conductive.[143]

Epitaxy

Epitaxy refers to the deposition of a crystalline overlayer on a crystalline substrate, where there is registry between the two. In some cases epitaxial graphene layers are coupled to surfaces weakly enough (by Van der Waals forces) to retain the two dimensional electronic band structure of isolated graphene.[144][145] An example of weakly coupled epitaxial graphene is the one grown on SiC.[30]

Graphene monolayers grown on silicon carbide and iridium are weakly coupled to these substrates (how weakly remains debated) and the graphene–substrate interaction can be further passivated.[11]

Silicon carbide

Heating silicon carbide (SiC) to high temperatures (>1100 °C) under low pressures (~10−6 torr) reduces it to graphene.[146] This process produces epitaxial graphene with dimensions dependent upon the size of the wafer. The face of the SiC used for graphene formation, silicon- or carbon-terminated, highly influences the thickness, mobility and carrier density of the resulting graphene.

Graphene's electronic band-structure (so-called Dirac cone structure) was first visualized in this material.[147][148][149] Weak anti-localization is observed in this material, but not in exfoliated graphene produced by the drawing method.[150] Large, temperature-independent mobilities approach those in exfoliated graphene placed on silicon oxide, but lower than mobilities in suspended graphene produced by the drawing method. Even without transfer, graphene on SiC exhibits massless Dirac fermions.[59][60][61][62][63][151][64]

The weak van der Waals force that provides the cohesion of multilayer graphene stacks does not always affect the electronic properties of the individual layers. That is, while the electronic properties of certain multilayered epitaxial graphenes are identical to that of a single layer,[152] in other cases the properties are affected,[147][148] as they are in bulk graphite. This effect is well understood theoretically and is related to the symmetry of the interlayer interactions.[152]

Epitaxial graphene on SiC can be patterned using standard microelectronics methods. A band gap can be created and tuned by laser irradiation.[153]

Metal substrates

The atomic structure of a metal substrate can seed the growth of graphene.

Ruthenium

Graphene grown on ruthenium does not typically produce uniform layer thickness. Bonding between the bottom graphene layer and the substrate may affect layer properties.[154]

Iridium

Graphene grown on iridium is very weakly bonded, uniform in thickness and can be highly ordered. As on many other substrates, graphene on iridium is slightly rippled. Due to the long-range order of these ripples, minigaps in the electronic band-structure (Dirac cone) become visible.[155]

Nickel

High-quality sheets of few-layer graphene exceeding 1 cm2 (0.2 sq in) in area have been synthesized via chemical vapor deposition on thin nickel films with methane as a carbon source. These sheets have been successfully transferred to various substrates.[59][156][157]

Graphene has been demonstrated at temperatures compatible with conventional CMOS processing, using a nickel-based alloy with gold as catalyst.[158]

This process dissolves carbon atoms inside a transition metal melt at a certain temperature and then precipitates the dissolved carbon at lower temperatures as single layer graphene (SLG).[159]

The metal is first melted in contact with a carbon source, possibly a graphite crucible inside which the melt is carried out or graphite powder/chunks that are placed in the melt. Keeping the melt in contact with the carbon at a specific temperature dissolves the carbon atoms, saturating the melt based on the binary phase diagram of metal-carbon. Upon lowering the temperature, carbon's solubility decreases and the excess carbon precipitates atop the melt. The floating layer can be either skimmed or frozen for later removal. Using different morphology, including thick graphite, few layer graphene (FLG) and SLG were observed on metal substrate. Raman spectroscopy proved that SLG had grown on nickel substrate. The SLG Raman spectrum featured no D and D′ band, indicating its pristine nature. Since nickel is not Raman active, direct Raman spectroscopy of graphene layers on top of the nickel is achievable.[159]

Copper

An improvement of this technique employs copper foil; at very low pressure, the growth of graphene automatically stops after a single graphene layer forms. Arbitrarily large films can be created.[156][160] The single layer growth is also due to the low concentration of carbon in methane. Larger hydrocarbons such as ethane and propane produce bilayer coatings.[161] Atmospheric pressure CVD growth produces multilayer graphene on copper (similar to nickel).[162]

Sodium ethoxide pyrolysis

Gram-quantities of graphene were produced by the reduction of ethanol by sodium metal, followed by pyrolysis of the ethoxide product and washing with water to remove sodium salts.[163]

Silicon/germanium/hydrogen

A normal silicon wafer coated with a layer of germanium (Ge) dipped in dilute hydrofluoric acid strips the naturally forming germanium oxide groups, creating hydrogen-terminated germanium. Chemical vapor deposition deposits a layer of graphene on top. The graphene can be peeled from the wafer using a dry process and is then ready for use. The wafer can be reused. The graphene is wrinkle-free, high quality and low in defects.[164][165]

Nanotube slicing

Graphene can be created by cutting open carbon nanotubes.[166] In one such method multi-walled carbon nanotubes are cut open in solution by action of potassium permanganate and sulfuric acid.[167] In another method graphene nanoribbons were produced by plasma etching of nanotubes partly embedded in a polymer film.[168]

Carbon dioxide reduction

A highly exothermic reaction combusts magnesium in an oxidation-reduction reaction with carbon dioxide, producing a variety of carbon nanoparticles including graphene and fullerenes. The carbon dioxide reactant may be either solid (dry-ice) or gaseous. The products of this reaction are carbon and magnesium oxide. US patent 8377408  was issued for this process.[169]

Applications

While as of 2014, graphene is not used in commercial applications, many have been proposed and/or are under active development, in areas including electronics, biological engineering, filtration, lightweight/strong composite materials, photovoltaics and energy storage.[170]


See also

3

References

  1. ^ "This Month in Physics History: October 22, 2004: Discovery of Graphene". APS News. Series II. 18 (9): 2. 2009.
  2. ^ "The Nobel Prize in Physics 2010". The Nobel Foundation. Retrieved 2013-12-03.
  3. ^ Boehm, H. P.; Setton, R.; Stumpp, E. (1994). "Nomenclature and terminology of graphite intercalation compounds" (PDF). Pure and Applied Chemistry. 66 (9): 1893–1901. doi:10.1351/pac199466091893.
  4. ^ Boehm, H. P.; Clauss, A.; Fischer, G. O.; Hofmann, U. (1962). "Das Adsorptionsverhalten sehr dünner Kohlenstoffolien". Zeitschrift für anorganische und allgemeine Chemie (in German). 316 (3–4): 119–127. doi:10.1002/zaac.19623160303. {{cite journal}}: Cite has empty unknown parameter: |trans_title= (help)
  5. ^ Mouras, S.; et al. (1987). "Synthesis of first stage graphite intercalation compounds with fluorides". Revue de Chimie Minerale. 24: 572. {{cite journal}}: Explicit use of et al. in: |author2= (help)
  6. ^ Saito, R.; Fujita, Mitsutaka; Dresselhaus, G.; Dresselhaus, M. (1992). "Electronic structure of graphene tubules based on C60". Physical Review B. 46 (3): 1804. Bibcode:1992PhRvB..46.1804S. doi:10.1103/PhysRevB.46.1804.
  7. ^ Forbeaux, I.; Themlin, J.-M.; Debever, J.-M. (1998). "Heteroepitaxial graphite on 6H-SiC(0001): Interface formation through conduction-band electronic structure". Physical Review B. 58 (24): 16396. Bibcode:1998PhRvB..5816396F. doi:10.1103/PhysRevB.58.16396.
  8. ^ Wang, S.; Yata, S.; Nagano, J.; Okano, Y.; Kinoshita, H.; Kikuta, H.; Yamabe, T. (2000). "A new carbonaceous material with large capacity and high efficiency for rechargeable Li-ion batteries". Journal of the Electrochemical Society. 147 (7): 2498. doi:10.1149/1.1393559.
  9. ^ "graphene layer". IUPAC Gold Book. International Union of Pure and Applied Chemistry. Retrieved 2012-03-31.
  10. ^ a b c Geim, A. (2009). "Graphene: Status and Prospects". Science. 324 (5934): 1530–4. arXiv:0906.3799. Bibcode:2009Sci...324.1530G. doi:10.1126/science.1158877. PMID 19541989.
  11. ^ a b Riedl, C.; Coletti, C.; Iwasaki, T.; Zakharov, A.A.; Starke, U. (2009). "Quasi-Free-Standing Epitaxial Graphene on SiC Obtained by Hydrogen Intercalation". Physical Review Letters. 103 (24): 246804. arXiv:0911.1953. Bibcode:2009PhRvL.103x6804R. doi:10.1103/PhysRevLett.103.246804. PMID 20366220.
  12. ^ Simpson, C. D.; Brand, J. Diedrich; Berresheim, Alexander J.; Przybilla, Laurence; Räder, Hans Joachim; Müllen, Klaus (2002). "Synthesis of a Giant 222 Carbon Graphite Sheet". Chemistry – A European Journal. 6 (6): 1424. doi:10.1002/1521-3765(20020315)8:6<1424::AID-CHEM1424>3.0.CO;2-Z.
  13. ^ Brodie, B. C. (1859). "On the Atomic Weight of Graphite". Philosophical Transactions of the Royal Society of London. 149: 249–259. Bibcode:1859RSPT..149..249B. JSTOR 108699.
  14. ^ Debije, P; Scherrer, P (1916). "Interferenz an regellos orientierten Teilchen im Röntgenlicht I". Physikalische Zeitschrift (in German). 17: 277. {{cite journal}}: Cite has empty unknown parameter: |trans_title= (help)
  15. ^ Friedrich, W (1913). "Eine neue Interferenzerscheinung bei Röntgenstrahlen". Physikalische Zeitschrift (in German). 14: 317. {{cite journal}}: Cite has empty unknown parameter: |trans_title= (help)
    Hull, AW (1917). "A New Method of X-ray Crystal Analysis". Phys. Rev. 10 (6): 661. Bibcode:1917PhRv...10..661H. doi:10.1103/PhysRev.10.661.
  16. ^ Kohlschütter, V.; Haenni, P. (1919). "Zur Kenntnis des Graphitischen Kohlenstoffs und der Graphitsäure". Zeitschrift für anorganische und allgemeine Chemie (in German). 105 (1): 121–144. doi:10.1002/zaac.19191050109. {{cite journal}}: Cite has empty unknown parameter: |trans_title= (help)
  17. ^ Bernal, JD (1924). "The Structure of Graphite". Proc. R. Soc. Lond. A106 (740): 749–773. JSTOR 94336.
    Hassel, O; Mack, H (1924). "Über die Kristallstruktur des Graphits". Zeitschrift für Physik (in German). 25: 317. Bibcode:1924ZPhy...25..317H. doi:10.1007/BF01327534. {{cite journal}}: Cite has empty unknown parameter: |trans_title= (help)
  18. ^ DiVincenzo, D. P.; Mele, E. J. (1984). "Self-Consistent Effective Mass Theory for Intralayer Screening in Graphite Intercalation Compounds". Physical Review B. 295 (4): 1685. Bibcode:1984PhRvB..29.1685D. doi:10.1103/PhysRevB.29.1685.
  19. ^ a b c d e f Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Katsnelson, M. I.; Grigorieva, I. V.; Dubonos, S. V.; Firsov, A. A. (2005). "Two-dimensional gas of massless Dirac fermions in graphene". Nature. 438 (7065): 197–200. arXiv:cond-mat/0509330. Bibcode:2005Natur.438..197N. doi:10.1038/nature04233. PMID 16281030.
  20. ^ a b Gusynin, V. P.; Sharapov, S. G. (2005). "Unconventional Integer Quantum Hall Effect in Graphene". Physical Review Letters. 95 (14): 146801. arXiv:cond-mat/0506575. Bibcode:2005PhRvL..95n6801G. doi:10.1103/PhysRevLett.95.146801. PMID 16241680.
  21. ^ a b c Zhang, Y.; Tan, Y. W.; Stormer, H. L.; Kim, P. (2005). "Experimental observation of the quantum Hall effect and Berry's phase in graphene". Nature. 438 (7065): 201–204. arXiv:cond-mat/0509355. Bibcode:2005Natur.438..201Z. doi:10.1038/nature04235. PMID 16281031.
  22. ^ Ruess, G.; Vogt, F. (1948). "Höchstlamellarer Kohlenstoff aus Graphitoxyhydroxyd". Monatshefte für Chemie (in German). 78 (3–4): 222. doi:10.1007/BF01141527. {{cite journal}}: Cite has empty unknown parameter: |trans_title= (help)
  23. ^ a b c d Meyer, J.; Geim, A. K.; Katsnelson, M. I.; Novoselov, K. S.; Booth, T. J.; Roth, S. (2007). "The structure of suspended graphene sheets". Nature. 446 (7131): 60–63. arXiv:cond-mat/0701379. Bibcode:2007Natur.446...60M. doi:10.1038/nature05545. PMID 17330039.
  24. ^ a b Boehm, H. P.; Clauss, A.; Fischer, G.; Hofmann, U. (1962). "Surface Properties of Extremely Thin Graphite Lamellae". Pergamon Press http://graphenetimes.com/wp-content/uploads/1961/09/BoehmProcCarbon1962.pdf. {{cite book}}: Missing or empty |title= (help); Unknown parameter |booktitle= ignored (help)
  25. ^ This paper reports graphitic flakes that give an additional contrast equivalent of down to ~0.4 nm or 3 atomic layers of amorphous carbon. This was the best possible resolution for 1960 TEMs. However, neither then nor today it is possible to argue how many layers were in those flakes. Now we know that the TEM contrast of graphene most strongly depends on focusing conditions.[23] For example, it is impossible to distinguish between suspended monolayer and multilayer graphene by their TEM contrasts, and the only known way is to analyse relative intensities of various diffraction spots. The first reliable TEM observations of monolayers are probably given in refs. 24 and 26 of Geim & Novoselov 2007
  26. ^ Oshima, C.; Nagashima, A. (1997). "Ultra-thin epitaxial films of graphite and hexagonal boron nitride on solid surfaces". J. Phys.: Condens. Matter. 9: 1. Bibcode:1997JPCM....9....1O. doi:10.1088/0953-8984/9/1/004.
  27. ^ Geim, A. K.; Kim, P. (April 2008). "Carbon Wonderland". Scientific American. ... bits of graphene are undoubtedly present in every pencil mark
  28. ^ a b c d e f g Geim & Novoselov 2007.
  29. ^ "United States Patent: 7071258". US Patent Office. Retrieved 2014-01-12.
  30. ^ a b Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Dubonos, S. V.; Grigorieva, I. V.; Firsov, A. A. (2004). "Electric Field Effect in Atomically Thin Carbon Films" (PDF). Science. 306 (5696): 666–669. arXiv:cond-mat/0410550. Bibcode:2004Sci...306..666N. doi:10.1126/science.1102896. PMID 15499015.
  31. ^ Kopelevich, Y.; Torres, J.; Da Silva, R.; Mrowka, F.; Kempa, H.; Esquinazi, P. (2003). "Reentrant Metallic Behavior of Graphite in the Quantum Limit". Physical Review Letters. 90 (15): 156402. arXiv:cond-mat/0209406. Bibcode:2003PhRvL..90o6402K. doi:10.1103/PhysRevLett.90.156402. PMID 12732058.
  32. ^ Luk’yanchuk, Igor A.; Kopelevich, Yakov (2004). "Phase Analysis of Quantum Oscillations in Graphite". Physical Review Letters. 93 (16): 166402. arXiv:cond-mat/0402058. Bibcode:2004PhRvL..93p6402L. doi:10.1103/PhysRevLett.93.166402. PMID 15525015.
  33. ^ "Graphene pioneers bag Nobel prize". Institute of Physics, UK. 5 October 2010.
  34. ^ a b Carlsson, J. M. (2007). "Graphene: Buckle or break". Nature Materials. 6 (11): 801–2. Bibcode:2007NatMa...6..801C. doi:10.1038/nmat2051. PMID 17972931.
  35. ^ a b Fasolino, A.; Los, J. H.; Katsnelson, M. I. (2007). "Intrinsic ripples in graphene". Nature Materials. 6 (11): 858–61. arXiv:0704.1793. Bibcode:2007NatMa...6..858F. doi:10.1038/nmat2011. PMID 17891144.
  36. ^ a b Ishigami, Masa; et al. (2007). "Atomic Structure of Graphene on SiO2". Nano Lett. 7 (6): 1643–1648. Bibcode:2007NanoL...7.1643I. doi:10.1021/nl070613a. PMID 17497819. {{cite journal}}: Explicit use of et al. in: |author2= (help)
  37. ^ Stolyarova, Elena; et al. (2007). "High-resolution scanning tunneling microscopy imaging of mesoscopic graphene sheets on an insulating surface". Proceedings of the National Academy of Sciences. 104 (22): 9209–9212. arXiv:0705.0833. Bibcode:2007PNAS..104.9209S. doi:10.1073/pnas.0703337104. PMC 1874226. PMID 17517635. {{cite journal}}: Explicit use of et al. in: |author2= (help)
  38. ^ Kasuya, D.; Yudasaka, M.; Takahashi, K.; Kokai, F.; Iijima, S. (2002). "Selective Production of Single-Wall Carbon Nanohorn Aggregates and Their Formation Mechanism". J. Phys. Chem. B. 106 (19): 4947. doi:10.1021/jp020387n.
  39. ^ Bernatowicz; T. J.; et al. (1996). "Constraints on stellar grain formation from presolar graphite in the Murchison meteorite". Astrophysical Journal. 472 (2): 760–782. Bibcode:1996ApJ...472..760B. doi:10.1086/178105.
  40. ^ Fraundorf, P.; Wackenhut, M. (2002). "The core structure of presolar graphite onions". Astrophysical Journal Letters. 578 (2): L153–156. arXiv:astro-ph/0110585. Bibcode:2002ApJ...578L.153F. doi:10.1086/344633.
  41. ^ Zan, Recep; Ramasse, Quentin M.; Bangert, Ursel; Novoselov, Konstantin S. (2012). "Graphene re-knits its holes". Mesoscale and Nanoscale Physics. 12 (8): 3936. arXiv:1207.1487v1. Bibcode:2012NanoL..12.3936Z. doi:10.1021/nl300985q.
  42. ^ Puiu, Tibi (12 July 2012). "Graphene sheets can repair themselves naturally". ZME Science.
  43. ^ Denis, P. A.; Iribarne, F. (2013). "Comparative Study of Defect Reactivity in Graphene". Journal of Physical Chemistry C. 117 (37): 19048. doi:10.1021/jp4061945.
  44. ^ Yamada, Y.; Murota, K; Fujita, R; Kim, J; et al. (2014). "Subnanometer vacancy defects introduced on graphene by oxygen gas". Journal of American Chemical Society. 136 (6): 2232. doi:10.1021/ja4117268. {{cite journal}}: Explicit use of et al. in: |author5= (help)
  45. ^ Eftekhari, A.; Jafarkhani, P. (2013). "Curly Graphene with Specious Interlayers Displaying Superior Capacity for Hydrogen Storage". Journal of Physical Chemistry C. 117 (48): 25845. doi:10.1021/jp410044v.
  46. ^ Yamada, Y.; Yasuda, H.; Murota, K.; Nakamura, M.; Sodesawa, T.; Sato, S. (2013). "Analysis of heat-treated graphite oxide by X-ray photoelectron spectroscopy". Journal of Material Science. 48 (23): 8171. doi:10.1007/s10853-013-7630-0.
  47. ^ Yamada, Y.; Kim, J.; Murota, K.; Matsuo, S.; Sato, S. (2014). "Nitrogen-containing graphene analyzed by X-ray photoelectron spectroscopy". Carbon. 70: 59. doi:10.1016/j.carbon.2013.12.061.
  48. ^ "Thinnest graphene sheets react strongly with hydrogen atoms; thicker sheets are relatively unaffected". Phys.org. 1 February 2013.
  49. ^ Heyrovska, Raji (2008). "Atomic Structures of Graphene, Benzene and Methane with Bond Lengths as Sums of the Single, Double and Resonance Bond Radii of Carbon". arXiv:0804.4086 [physics.gen-ph].
  50. ^ a b c Neto, A Castro; Peres, N. M. R.; Novoselov, K. S.; Geim, A. K.; Geim, A. K. (2009). "The electronic properties of graphene" (PDF). Rev Mod Phys. 81: 109. arXiv:0709.1163. Bibcode:2009RvMP...81..109C. doi:10.1103/RevModPhys.81.109.
  51. ^ a b c d Charlier, J.-C.; Eklund, P.C.; Zhu, J.; Ferrari, A.C. (2008). Jorio, A.; Dresselhaus and, G.; Dresselhaus, M.S. (eds.). Electron and Phonon Properties of Graphene: Their Relationship with Carbon Nanotubes. Berlin/Heidelberg: Springer-Verlag. {{cite book}}: |work= ignored (help)
  52. ^ a b c d e Semenoff, G. W. (1984). "Condensed-Matter Simulation of a Three-Dimensional Anomaly". Physical Review Letters. 53 (26): 2449. Bibcode:1984PhRvL..53.2449S. doi:10.1103/PhysRevLett.53.2449.
  53. ^ Wallace, P.R. (1947). "The Band Theory of Graphite". Physical Review. 71 (9): 622. Bibcode:1947PhRv...71..622W. doi:10.1103/PhysRev.71.622.
  54. ^ a b Avouris, P.; Chen, Z.; Perebeinos, V. (2007). "Carbon-based electronics". Nature Nanotechnology. 2 (10): 605–15. Bibcode:2007NatNa...2..605A. doi:10.1038/nnano.2007.300. PMID 18654384.
  55. ^ Lamas, C.A.; Cabra, D.C.; Grandi, N. (2009). "Generalized Pomeranchuk instabilities in graphene". Physical Review B. 80 (7): 75108. arXiv:0812.4406. Bibcode:2009PhRvB..80g5108L. doi:10.1103/PhysRevB.80.075108.
  56. ^ Cite error: The named reference castro was invoked but never defined (see the help page).
  57. ^ a b c "new-form-of-graphene-allows-electrons-to-behave-like-photons".
  58. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1038/nature12952, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1038/nature12952 instead.
  59. ^ a b c Kim, Kuen Soo; et al. (2009). "Large-scale pattern growth of graphene films for stretchable transparent electrodes". Nature. 457 (7230): 706–10. Bibcode:2009Natur.457..706K. doi:10.1038/nature07719. PMID 19145232. {{cite journal}}: Explicit use of et al. in: |author2= (help)
  60. ^ a b Jobst, Johannes; Waldmann, Daniel; Speck, Florian; Hirner, Roland; Maude, Duncan K.; Seyller, Thomas; Weber, Heiko B. (2009). "How Graphene-like is Epitaxial Graphene? Quantum Oscillations and Quantum Hall Effect". Physical Review B. 81 (19): 195434. arXiv:0908.1900. Bibcode:2010PhRvB..81s5434J. doi:10.1103/PhysRevB.81.195434.
  61. ^ a b Shen, T.; Gu, J.J.; Xu, M; Wu, Y.Q.; Bolen, M.L.; Capano, M.A.; Engel, L.W.; Ye, P.D. (2009). "Observation of quantum-Hall effect in gated epitaxial graphene grown on SiC (0001)". Applied Physics Letters. 95 (17): 172105. arXiv:0908.3822. Bibcode:2009ApPhL..95q2105S. doi:10.1063/1.3254329.
  62. ^ a b Wu, Xiaosong; Hu, Yike; Ruan, Ming; Madiomanana, Nerasoa K; Hankinson, John; Sprinkle, Mike; Berger, Claire; de Heer, Walt A. (2009). "Half integer quantum Hall effect in high mobility single layer epitaxial graphene". Applied Physics Letters. 95 (22): 223108. arXiv:0909.2903. Bibcode:2009ApPhL..95v3108W. doi:10.1063/1.3266524.
  63. ^ a b Lara-Avila, Samuel; Kalaboukhov, Alexei; Paolillo, Sara; Syväjärvi, Mikael; Yakimova, Rositza; Fal'ko, Vladimir; Tzalenchuk, Alexander; Kubatkin, Sergey (Submitted 7 July 2009). "SiC Graphene Suitable For Quantum Hall Resistance Metrology". Science Brevia. arXiv:0909.1193. {{cite journal}}: Check date values in: |date= (help)CS1 maint: date and year (link)
  64. ^ a b Alexander-Webber, J.A.; Baker, A.M.R.; Janssen, T.J.B.M.; Tzalenchuk, A.; Lara-Avila, S.; Kubatkin, S.; Yakimova, R.; Piot, B. A.; Maude, D. K.; Nicholas, R.J. (2013). "Phase Space for the Breakdown of the Quantum Hall Effect in Epitaxial Graphene". Physical Review Letters. 111 (9): 096601. arXiv:1304.4897. Bibcode:2013PhRvL.111i6601A. doi:10.1103/PhysRevLett.111.096601. PMID 24033057.
  65. ^ Fuhrer, Michael S. (2009). "A physicist peels back the layers of excitement about graphene". Nature. 459 (7250): 1037. Bibcode:2009Natur.459.1037F. doi:10.1038/4591037e. PMID 19553953.
  66. ^ a b Zhang, Y.; Jiang, Z.; Small, J. P.; Purewal, M. S.; Tan, Y.-W.; Fazlollahi, M.; Chudow, J. D.; Jaszczak, J. A.; Stormer, H. L.; Kim, P. (2006). "Landau-Level Splitting in Graphene in High Magnetic Fields". Physical Review Letters. 96 (13): 136806. arXiv:cond-mat/0602649. Bibcode:2006PhRvL..96m6806Z. doi:10.1103/PhysRevLett.96.136806.
  67. ^ a b c d Du, X.; Skachko, Ivan; Duerr, Fabian; Luican, Adina; Andrei, Eva Y. (2009). "Fractional quantum Hall effect and insulating phase of Dirac electrons in graphene". Nature. 462 (7270): 192–195. arXiv:0910.2532. Bibcode:2009Natur.462..192D. doi:10.1038/nature08522. PMID 19829294.
  68. ^ a b Bolotin, K.; Ghahari, Fereshte; Shulman, Michael D.; Stormer, Horst L.; Kim, Philip (2009). "Observation of the fractional quantum Hall effect in graphene". Nature. 462 (7270): 196–199. arXiv:0910.2763. Bibcode:2009Natur.462..196B. doi:10.1038/nature08582. PMID 19881489.
  69. ^ Bordag, M.; Fialkovsky, I. V.; Gitman, D. M.; Vassilevich, D. V. (2009). "Casimir interaction between a perfect conductor and graphene described by the Dirac model". Physical Review B. 80 (24): 245406. arXiv:0907.3242. Bibcode:2009PhRvB..80x5406B. doi:10.1103/PhysRevB.80.245406.
  70. ^ Fialkovsky, I. V.; Marachevsky, V.N.; Vassilevich, D. V. (2011). "Finite temperature Casimir effect for graphene". Physical Review B. 84 (35446): 35446. arXiv:1102.1757. Bibcode:2011PhRvB..84c5446F. doi:10.1103/PhysRevB.84.035446.
  71. ^ Dobson, A.; Rubio, A. (2006). "Asymptotics of the dispersion interaction: analytic benchmarks for van der Waals energy functionals". Physical Review Letters. 96 (7): 073201. arXiv:cond-mat/0502422. Bibcode:2006PhRvL..96g3201D. doi:10.1103/PhysRevLett.96.073201.
  72. ^ a b c Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1126/science.1240317 , please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1126/science.1240317 instead.
  73. ^ Morozov, S.V.; Novoselov, K.; Katsnelson, M.; Schedin, F.; Elias, D.; Jaszczak, J.; Geim, A. (2008). "Giant Intrinsic Carrier Mobilities in Graphene and Its Bilayer". Physical Review Letters. 100 (1): 016602. arXiv:0710.5304. Bibcode:2008PhRvL.100a6602M. doi:10.1103/PhysRevLett.100.016602. PMID 18232798.
  74. ^ a b c Chen, J. H.; Jang, Chaun; Xiao, Shudong; Ishigami, Masa; Fuhrer, Michael S. (2008). "Intrinsic and Extrinsic Performance Limits of Graphene Devices on SiO2". Nature Nanotechnology. 3 (4): 206–9. doi:10.1038/nnano.2008.58. PMID 18654504.
  75. ^ Akturk, A.; Goldsman, N. (2008). "Electron transport and full-band electron–phonon interactions in graphene". Journal of Applied Physics. 103 (5): 053702. Bibcode:2008JAP...103e3702A. doi:10.1063/1.2890147.
  76. ^ Physicists Show Electrons Can Travel More Than 100 Times Faster in Graphene :: University Communications Newsdesk, University of Maryland. Newsdesk.umd.edu (2008-03-24). Retrieved on 2014-01-12.
  77. ^ a b c Chen, J. H.; Jang, C.; Adam, S.; Fuhrer, M. S.; Williams, E. D.; Ishigami, M. (2008). "Charged Impurity Scattering in Graphene". Nature Physics. 4 (5): 377–381. arXiv:0708.2408. Bibcode:2008NatPh...4..377C. doi:10.1038/nphys935.
  78. ^ a b Schedin, F.; Geim, A. K.; Morozov, S. V.; Hill, E. W.; Blake, P.; Katsnelson, M. I.; Novoselov, K. S. (2007). "Detection of individual gas molecules adsorbed on graphene". Nature Materials. 6 (9): 652–655. Bibcode:2007NatMa...6..652S. doi:10.1038/nmat1967. PMID 17660825.
  79. ^ Adam, S.; Hwang, E. H.; Galitski, V. M.; Das Sarma, S. (2007). "A self-consistent theory for graphene transport". Proc. Nat. Acad. Sci. USA. 104 (47): 18392–7. arXiv:0705.1540. Bibcode:2007PNAS..10418392A. doi:10.1073/pnas.0704772104. PMC 2141788. PMID 18003926.
  80. ^ Steinberg, Hadar; Barak, Gilad; Yacoby, Amir; et al. (2008). "Charge fractionalization in quantum wires (Letter)". Nature Physics. 4 (2): 116–119. arXiv:0803.0744. Bibcode:2008NatPh...4..116S. doi:10.1038/nphys810. {{cite journal}}: Explicit use of et al. in: |author4= (help)
  81. ^ Trisetyarso, Agung (2012). "Dirac four-potential tunings-based quantum transistor utilizing the Lorentz force". Quantum Information & Computation. 12 (11–12): 989. arXiv:1003.4590. Bibcode:2010arXiv1003.4590T.
  82. ^ Pachos, Jiannis K. (2009). "Manifestations of topological effects in graphene". Contemporary Physics. 50 (2): 375. arXiv:0812.1116. Bibcode:2009ConPh..50..375P. doi:10.1080/00107510802650507.
    Franz, M. (5 January 2008). "Fractionalization of charge and statistics in graphene and related structures" (PDF). University of British Columbia.
  83. ^ Kuzmenko, A. B.; Van Heumen, E.; Carbone, F.; Van Der Marel, D. (2008). "Universal infrared conductance of graphite". Physical Review Letters. 100 (11): 117401. arXiv:0712.0835. Bibcode:2008PhRvL.100k7401K. doi:10.1103/PhysRevLett.100.117401. PMID 18517825.
  84. ^ Nair, R. R.; Blake, P.; Grigorenko, A. N.; Novoselov, K. S.; Booth, T. J.; Stauber, T.; Peres, N. M. R.; Geim, A. K. (2008). "Fine Structure Constant Defines Visual Transparency of Graphene" (PDF). Science. 320 (5881): 1308. Bibcode:2008Sci...320.1308N. doi:10.1126/science.1156965. PMID 18388259. [1]
  85. ^ "Graphene Gazing Gives Glimpse Of Foundations Of Universe". ScienceDaily. 4 April 2008.
  86. ^ Zhang, Y.; Tang, Tsung-Ta; Girit, Caglar; Hao, Zhao; Martin, Michael C.; Zettl, Alex; Crommie, Michael F.; Shen, Y. Ron; Wang, Feng (11 June 2009). "Direct observation of a widely tunable bandgap in bilayer graphene". Nature. 459 (7248): 820–823. Bibcode:2009Natur.459..820Z. doi:10.1038/nature08105. PMID 19516337. {{cite journal}}: Invalid |display-authors=9 (help)
  87. ^ Liu, Junfeng; Wright, A. R.; Zhang, Chao; Ma, Zhongshui (29 July 2008). "Strong terahertz conductance of graphene nanoribbons under a magnetic field". Appl Phys Lett. 93 (4): 041106–041110. Bibcode:2008ApPhL..93d1106L. doi:10.1063/1.2964093.
  88. ^ a b Kurum, U.; Liu, Bo; Zhang, Kailiang; Liu, Yan; Zhang, Hao (2011). "Electrochemically tunable ultrafast optical response of graphene oxide". Applied Physics Letters. 98 (2): 141103. Bibcode:2011ApPhL..98b1103M. doi:10.1063/1.3540647.
  89. ^ Sreekanth, K.V.; Zeng, Shuwen; Shang, Jingzhi; Yong, Ken-Tye; Yu, Ting (2012). "Excitation of surface electromagnetic waves in a graphene-based Bragg grating". Scientific Reports. 2: 737. Bibcode:2012NatSR...2E.737S. doi:10.1038/srep00737. PMC 3471096. PMID 23071901.
  90. ^ Bao, Qiaoliang; Zhang, Han; Wang, Yu; Ni, Zhenhua; Yan, Yongli; Shen, Ze Xiang; Loh, Kian Ping; Tang, Ding Yuan (2009). "Atomic-Layer Graphene as a Saturable Absorber for Ultrafast Pulsed Lasers" (PDF). Advanced Functional Materials. 19 (19): 3077. doi:10.1002/adfm.200901007. Archived from the original (PDF) on 2011-07-17.
    Zhang, H.; Tang, D. Y.; Zhao, L. M.; Bao, Q. L.; Loh, K. P. (2009). "Large energy mode locking of an erbium-doped fiber laser with atomic layer graphene" (PDF). Optics Express. 17 (20): P17630. arXiv:0909.5536. Bibcode:2009OExpr..1717630Z. doi:10.1364/OE.17.017630. Archived from the original (PDF) on 2011-07-17.
    Zhang, H.; Bao, Qiaoliang; Tang, Dingyuan; Zhao, Luming; Loh, Kianping (2009). "Large energy soliton erbium-doped fiber laser with a graphene-polymer composite mode locker" (PDF). Applied Physics Letters. 95 (14): P141103. arXiv:0909.5540. Bibcode:2009ApPhL..95n1103Z. doi:10.1063/1.3244206. Archived from the original (PDF) on 2011-07-17.
    Zhang, H.; Tang, Dingyuan; Knize, R. J.; Zhao, Luming; Bao, Qiaoliang; Loh, Kian Ping (2010). "Graphene mode locked, wavelength-tunable, dissipative soliton fiber laser". Applied Physics Letters. 96 (11): 111112. arXiv:1003.0154. Bibcode:2010ApPhL..96k1112Z. doi:10.1063/1.3367743. Archived from the original (PDF) on 2010-05-21., Zhang (2009). "Graphene: Mode-locked lasers". NPG Asia Materials. doi:10.1038/asiamat.2009.52.
  91. ^ Zheng, Z.; Zhao, Chujun; Lu, Shunbin; Chen, Yu; Li, Ying; Zhang, Han; Wen, Shuangchun (2012). "Microwave and optical saturable absorption in graphene" (PDF). Optics Express. 20 (21): 23201–23214. Bibcode:2012OExpr..2023201Z. doi:10.1364/OE.20.023201. PMID 23188285.
  92. ^ Zhang, H.; Virally, Stéphane; Bao, Qiaoliang; Kian Ping, Loh; Massar, Serge; Godbout, Nicolas; Kockaert, Pascal (2012). "Z-scan measurement of the nonlinear refractive index of graphene". Optics Letters. 37 (11): 1856–1858. Bibcode:2012OptL...37.1856Z. doi:10.1364/OL.37.001856. PMID 22660052.
  93. ^ Onida, Giovanni; Rubio, Angel (2002). "Electronic excitations: Density-functional versus many-body Green's-function approaches". Rev. Mod. Phys. 74 (2): 601. Bibcode:2002RvMP...74..601O. doi:10.1103/RevModPhys.74.601.
  94. ^ Yang, Li; Deslippe, Jack; Park, Cheol-Hwan; Cohen, Marvin; Louie, Steven (2009). "Excitonic Effects on the Optical Response of Graphene and Bilayer Graphene". Physical Review Letters. 103 (18): 186802. arXiv:0906.0969. Bibcode:2009PhRvL.103r6802Y. doi:10.1103/PhysRevLett.103.186802. PMID 19905823.
  95. ^ Prezzi, Deborah; Varsano, Daniele; Ruini, Alice; Marini, Andrea; Molinari, Elisa (2008). "Optical properties of graphene nanoribbons: The role of many-body effects". Physical Review B. 77 (4): 041404. arXiv:0706.0916. Bibcode:2008PhRvB..77d1404P. doi:10.1103/PhysRevB.77.041404.
    Yang, Li; Cohen, Marvin L.; Louie, Steven G. (2007). "Excitonic Effects in the Optical Spectra of Graphene Nanoribbons". Nano Lett. 7 (10): 3112–5. arXiv:0707.2983. Bibcode:2007NanoL...7.3112Y. doi:10.1021/nl0716404. PMID 17824720.
    Yang, Li; Cohen, Marvin L.; Louie, Steven G. (2008). "Magnetic Edge-State Excitons in Zigzag Graphene Nanoribbons". Physical Review Letters. 101 (18): 186401. Bibcode:2008PhRvL.101r6401Y. doi:10.1103/PhysRevLett.101.186401. PMID 18999843.
  96. ^ Zhu, Xi; Su, Haibin (2010). "Excitons of Edge and Surface Functionalized Graphene Nanoribbons". J. Phys. Chem. C. 114 (41): 17257. doi:10.1021/jp102341b.
  97. ^ Wang, Min; Li, Chang Ming (2011). "Excitonic properties of hydrogen saturation-edged armchair graphene nanoribbons". Nanoscale. 3 (5): 2324–8. Bibcode:2011Nanos...3.2324W. doi:10.1039/c1nr10095e. PMID 21503364.
  98. ^ Bolmatov, Dima; Mou, Chung-Yu (2010). "Josephson effect in graphene SNS junction with a single localized defect". Physica B. 405 (13): 2896. arXiv:1006.1391. Bibcode:2010PhyB..405.2896B. doi:10.1016/j.physb.2010.04.015. {{cite journal}}: Cite has empty unknown parameter: |1= (help)
    Bolmatov, Dima; Mou, Chung-Yu (2010). "Tunneling conductance of the graphene SNS junction with a single localized defect". Journal of Experimental and Theoretical Physics (JETP). 110 (4): 613. arXiv:1006.1386. Bibcode:2010JETP..110..613B. doi:10.1134/S1063776110040084.
  99. ^ Zhu, Xi; Su, Haibin (2011). "Scaling of Excitons in Graphene Nanoribbons with Armchair Shaped Edges". Journal of Physical Chemistry A. 115 (43): 11998–12003. doi:10.1021/jp202787h.
  100. ^ Shenderova, O. B.; Zhirnov, V. V.; Brenner, D. W. (2002). "Carbon Nanostructures". Critical Reviews in Solid State and Materials Sciences. 27 (3–4): 227. Bibcode:2002CRSSM..27..227S. doi:10.1080/10408430208500497.
  101. ^ Chen, Shanshan; Wu, Qingzhi; Mishra, Columbia; Kang, Junyong; Zhang, Hengji; Cho, Kyeongjae; Cai, Weiwei; Balandin, Alexander A.; Ruoff, Rodney S. (2012). "Thermal conductivity of isotopically modified graphene". Nature Materials. 11 (3) (published 2012-01-10): 203. arXiv:1112.5752. Bibcode:2012NatMa..11..203C. doi:10.1038/nmat3207. {{cite journal}}: Invalid |display-authors=9 (help)
    Lay summary: Tracy, Suzanne (12 January 2012). "Keeping Electronics Cool". Scientific Computing. Advantage Business Media (published 2012-01-12). scientificcomputing.com.
  102. ^ Balandin, A. A.; Ghosh, Suchismita; Bao, Wenzhong; Calizo, Irene; Teweldebrhan, Desalegne; Miao, Feng; Lau, Chun Ning (20 February 2008). "Superior Thermal Conductivity of Single-Layer Graphene". Nano Letters ASAP. 8 (3): 902–907. Bibcode:2008NanoL...8..902B. doi:10.1021/nl0731872. PMID 18284217.
  103. ^ Saito, K.; Nakamura, J.; Natori, A. (2007). "Ballistic thermal conductance of a graphene sheet". Physical Review B. 76 (11): 115409. Bibcode:2007PhRvB..76k5409S. doi:10.1103/PhysRevB.76.115409.
  104. ^ Liang, Qizhen; Yao, Xuxia; Wang, Wei; Liu, Yan; Wong, Ching Ping (2011). "A Three-Dimensional Vertically Aligned Functionalized Multilayer Graphene Architecture: An Approach for Graphene-Based Thermal Interfacial Materials". ACS Nano. 5 (3): 2392–2401. doi:10.1021/nn200181e. PMID 21384860.
  105. ^ Delhaes, P. (2001). Graphite and Precursors. CRC Press. ISBN 90-5699-228-7.
  106. ^ a b Mingo, N.; Broido, D.A. (2005). "Carbon Nanotube Ballistic Thermal Conductance and Its Limits". Physical Review Letters. 95 (9): 096105. Bibcode:2005PhRvL..95i6105M. doi:10.1103/PhysRevLett.95.096105.
  107. ^ Mounet, N.; Marzari, N. (2005). "First-principles determination of the structural, vibrational and thermodynamic properties of diamond, graphite, and derivatives". Physical Review B. 71 (20): 205214. arXiv:cond-mat/0412643. Bibcode:2005PhRvB..71t5214M. doi:10.1103/PhysRevB.71.205214.
  108. ^ Lifshitz, I.M. (1952). Journal of Experimental and Theoretical Physics (in Russian). 22: 475. {{cite journal}}: Missing or empty |title= (help)
  109. ^ Braga, S.; Coluci, V. R.; Legoas, S. B.; Giro, R.; Galvão, D. S.; Baughman, R. H. (2004). "Structure and Dynamics of Carbon Nanoscrolls". Nano Letters. 4 (5): 881. Bibcode:2004NanoL...4..881B. doi:10.1021/nl0497272.
  110. ^ a b "2010 Nobel Physics Laureates" (PDF). nobelprize.org.
  111. ^ Lee, C.; Wei, X.; Kysar, J. W.; Hone, J. (2008). "Measurement of the Elastic Properties and Intrinsic Strength of Monolayer Graphene". Science. 321 (5887): 385–8. Bibcode:2008Sci...321..385L. doi:10.1126/science.1157996. PMID 18635798. {{cite journal}}: Unknown parameter |laysummary= ignored (help)
  112. ^ Sanderson, Bill (25 August 2008). "Toughest Stuff Known to Man: Discovery Opens Door to Space Elevator". nypost.com.
  113. ^ "Breakthrough in Developing Super-Material Graphene". ScienceDaily. 20 January 2010.
  114. ^ Frank, I. W.; Tanenbaum, D. M.; Van Der Zande, A.M.; McEuen, P. L. (2007). "Mechanical properties of suspended graphene sheets" (PDF). J. Vac. Sci. Technol. B. 25 (6): 2558–2561. Bibcode:2007JVSTB..25.2558F. doi:10.1116/1.2789446.
  115. ^ Bolmatov, Dima; Mou, Chung-Yu (2011). "Graphene-based modulation-doped superlattice structures". Journal of Experimental and Theoretical Physics (JETP). 112: 102. arXiv:1011.2850. Bibcode:2011JETP..112..102B. doi:10.1134/S1063776111010043.
    Bolmatov, Dima (2011). "Thermodynamic properties of tunneling quasiparticles in graphene-based structures". Physica C. 471 (23–24): 1651. arXiv:1106.6331. Bibcode:2011PhyC..471.1651B. doi:10.1016/j.physc.2011.07.008.
  116. ^ a b Tombros, Nikolaos; et al. (2007). "Electronic spin transport and spin precession in single graphene layers at room temperature". Nature. 448 (7153): 571–575. arXiv:0706.1948. Bibcode:2007Natur.448..571T. doi:10.1038/nature06037. PMID 17632544. {{cite journal}}: |format= requires |url= (help); Explicit use of et al. in: |author2= (help)
  117. ^ a b Cho, Sungjae; Chen, Yung-Fu; Fuhrer, Michael S. (2007). "Gate-tunable Graphene Spin Valve". Applied Physics Letters. 91 (12): 123105. arXiv:0706.1597. Bibcode:2007ApPhL..91l3105C. doi:10.1063/1.2784934.
  118. ^ Ohishi, Megumi; et al. (2007). "Spin Injection into a Graphene Thin Film at Room Temperature". Jpn J Appl Phys. 46: L605–L607. arXiv:0706.1451. Bibcode:2007JaJAP..46L.605O. doi:10.1143/JJAP.46.L605. {{cite journal}}: Explicit use of et al. in: |author2= (help)
  119. ^ "Graphene Oxide Paper". Northwestern University. Archived from the original on 2011-07-20. Retrieved 2011-02-28.
  120. ^ Eftekhari, Ali; Yazdani, Bahareh (2010). "Initiating electropolymerization on graphene sheets in graphite oxide structure". Journal of Polymer Science Part A: Polymer Chemistry. 48 (10): 2204. Bibcode:2010JPoSA..48.2204E. doi:10.1002/pola.23990.
  121. ^ Nalla, Venkatram; Polavarapu, L; Manga, KK; Goh, BM; Loh, KP; Xu, QH; Ji, W (2010). "Transient photoconductivity and femtosecond nonlinear optical properties of a conjugated polymer–graphene oxide composite". Nanotechnology. 21 (41): 415203. Bibcode:2010Nanot..21O5203N. doi:10.1088/0957-4484/21/41/415203. PMID 20852355.
  122. ^ Nair, R. R.; Wu, H. A.; Jayaram, P. N.; Grigorieva, I. V.; Geim, A. K. (2012). "Unimpeded permeation of water through helium-leak-tight graphene-based membranes". Science. 335 (6067): 442–4. arXiv:1112.3488. Bibcode:2012Sci...335..442N. doi:10.1126/science.1211694. PMID 22282806.
  123. ^ Niyogi, Sandip; Bekyarova, Elena; Itkis, Mikhail E.; McWilliams, Jared L.; Hamon, Mark A.; Haddon, Robert C. (2006). "Solution Properties of Graphite and Graphene". J. Am. Chem. Soc. 128 (24): 7720–7721. doi:10.1021/ja060680r. PMID 16771469.
  124. ^ Whitby, Raymond L.D.; Korobeinyk, Alina; Glevatska, Katya V. (2011). "Morphological changes and covalent reactivity assessment of single-layer graphene oxides under carboxylic group-targeted chemistry". Carbon. 49 (2): 722–725. doi:10.1016/j.carbon.2010.09.049.
  125. ^ Park, Sungjin; Dikin, Dmitriy A.; Nguyen, SonBinh T.; Ruoff, Rodney S. (2009). "Graphene Oxide Sheets Chemically Cross-Linked by Polyallylamine". J. Phys. Chem. C. 113 (36): 15801–15804. doi:10.1021/jp907613s.
  126. ^ Elias, D. C.; Nair, R. R.; Mohiuddin, T. M. G.; Morozov, S. V.; Blake, P.; Halsall, M. P.; Ferrari, A. C.; Boukhvalov, D. W.; Katsnelson, M. I.; Geim, A. K.; Novoselov, K. S. (2009). "Control of Graphene's Properties by Reversible Hydrogenation: Evidence for Graphane". Science. 323 (5914): 610–3. arXiv:0810.4706. Bibcode:2009Sci...323..610E. doi:10.1126/science.1167130. PMID 19179524.
  127. ^ Min, Hongki; Sahu, Bhagawan; Banerjee, Sanjay; MacDonald, A. (2007). "Ab initio theory of gate induced gaps in graphene bilayers". Physical Review B. 75 (15): 155115. arXiv:cond-mat/0612236. Bibcode:2007PhRvB..75o5115M. doi:10.1103/PhysRevB.75.155115.
  128. ^ Barlas, Yafis; Côté, R.; Lambert, J.; MacDonald, A. H. (2010). "Anomalous Exciton Condensation in Graphene Bilayers". Physical Review Letters. 104 (9): 96802. arXiv:0909.1502. Bibcode:2010PhRvL.104i6802B. doi:10.1103/PhysRevLett.104.096802.
  129. ^ Min, Lola; Hovden, Robert; Huang, Pinshane; Wojcik, Michal; Muller, David A.; Park, Jiwoong (2012). "Twinning and Twisting of Tri- and Bilayer Graphene". NanoLetters. 12 (3): 1609. Bibcode:2012NanoL..12.1609B. doi:10.1021/nl204547v.
  130. ^ Li, Xinming; Zhao, Tianshuo; Wang, Kunlin; Yang, Ying; Wei, Jinquan; Kang, Feiyu; Wu, Dehai; Zhu, Hongwei (29 August 2011). "Directly Drawing Self-Assembled, Porous, and Monolithic Graphene Fiber from Chemical Vapor Deposition Grown Graphene Film and Its Electrochemical Properties". Langmuir. 27 (19): 12164–71. doi:10.1021/la202380g. PMID 21875131.
  131. ^ "Flexible all solid-state supercapacitors based on chemical vapor deposition derived graphene fibers". 3 September 2013.
  132. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1002/ange.201303497, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1002/ange.201303497 instead.
    Wang, Hui; Sun, Kai; Tao, Franklin; Stacchiola, Dario J.; Hu, Yun Hang. "3D graphene could replace expensive platinum in solar cells". Angewandte Chemie. 125 (35). KurzweilAI: 9380. doi:10.1002/ange.201303497. Retrieved 2013-08-24.
  133. ^ a b Geim, A. K.; MacDonald, A. H. (2007). "Graphene: Exploring carbon flatland". Physics Today. 60 (8): 35–41. Bibcode:2007PhT....60h..35G. doi:10.1063/1.2774096.
  134. ^ "Boehm's 1961 isolation of graphene". Graphene Times. 7 December 2009.
  135. ^ "Many Pioneers in Graphene Discovery". Letters to the Editor. Aps.org. January 2010.
  136. ^ Dotzer, C.; Enzelberger-Heim, M.; Grimm, S.; Xiao, J.; Papp, C.; Lytken, O.; Steinrück, H.-P.; Müller, P.; Hirsch, A. (2013). "Wet Chemical Synthesis of Graphene". Advanced Materials. 25 (26): 3583–3587. doi:10.1002/adma.201300155. PMID 23703794. {{cite journal}}: |first4= missing |last4= (help)
  137. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1007/s10853-013-7630-0 , please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1007/s10853-013-7630-0 instead.
  138. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1039/C2CP40790F , please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1039/C2CP40790F instead.
  139. ^ "Laser Scribing of High-Performance and Flexible Graphene-Based Electrochemical Capacitors". Sciencemag.org. 16 March 2012.
    Marcus, Jennifer (15 March 2012). "Researchers develop graphene supercapacitor holding promise for portable electronics / UCLA Newsroom". Newsroom.ucla.edu.
  140. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1038/nnano.2008.215 , please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1038/nnano.2008.215 instead.
  141. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1039/C1JM11076D , please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1039/C1JM11076D instead.
  142. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1039/C0JM02461A , please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1039/C0JM02461A instead.
  143. ^ Woltornist, S. J., Oyer, A. J., Carrillo, J.-M. Y., Dobrynin, A. V, & Adamson, D. H. (2013). Conductive thin films of pristine graphene by solvent interface trapping. ACS nano, 7(8), 7062–6. doi:10.1021/nn402371c
  144. ^ Gall, N. R.; Rut'Kov, E. V.; Tontegode, A. Ya. (1997). "Two Dimensional Graphite Films on Metals and Their Intercalation". International Journal of Modern Physics B. 11 (16): 1865. Bibcode:1997IJMPB..11.1865G. doi:10.1142/S0217979297000976.
  145. ^ Gall, N. R.; Rut'Kov, E. V.; Tontegode, A. Ya. (1995). "Influence of surface carbon on the formation of silicon-refractory metal interfaces". Thin Solid Films. 266 (2): 229. Bibcode:1995TSF...266..229G. doi:10.1016/0040-6090(95)06572-5.
  146. ^ Sutter, P. (2009). "Epitaxial graphene: How silicon leaves the scene". Nature Materials. 8 (3): 171–2. Bibcode:2009NatMa...8..171S. doi:10.1038/nmat2392. PMID 19229263.
  147. ^ a b Ohta, T.; Bostwick, Aaron; McChesney, J.; Seyller, Thomas; Horn, Karsten; Rotenberg, Eli (2007). "Interlayer Interaction and Electronic Screening in Multilayer Graphene Investigated with Angle-Resolved Photoemission Spectroscopy". Physical Review Letters. 98 (20): 206802. Bibcode:2007PhRvL..98t6802O. doi:10.1103/PhysRevLett.98.206802. PMID 17677726.
  148. ^ a b Bostwick, A.; Ohta, Taisuke; McChesney, Jessica L; Emtsev, Konstantin V; Seyller, Thomas; Horn, Karsten; Rotenberg, Eli (2007). "Symmetry breaking in few layer graphene films". New Journal of Physics. 9 (10): 385. arXiv:0705.3705. Bibcode:2007NJPh....9..385B. doi:10.1088/1367-2630/9/10/385.
  149. ^ Zhou, S.Y.; Gweon, G.-H.; Graf, J.; Fedorov, A. V.; Spataru, C. D.; Diehl, R. D.; Kopelevich, Y.; Lee, D.-H.; Louie, Steven G.; Lanzara, A. (2006). "First direct observation of Dirac fermions in graphite". Nature Physics. 2 (9): 595–599. arXiv:cond-mat/0608069. Bibcode:2006NatPh...2..595Z. doi:10.1038/nphys393.
  150. ^ Morozov, S.V.; Novoselov, K. S.; Katsnelson, M. I.; Schedin, F.; Ponomarenko, L. A.; Jiang, D.; Geim, A. K. (2006). "Strong Suppression of Weak Localization in Graphene". Physical Review Letters. 97 (1): 016801. arXiv:cond-mat/0603826. Bibcode:2006PhRvL..97a6801M. doi:10.1103/PhysRevLett.97.016801. PMID 16907394.
  151. ^ Tzalenchuk, Alexander; Lara-Avila, Samuel; Kalaboukhov, Alexei; Paolillo, Sara; Syväjärvi, Mikael; Yakimova, Rositza; Kazakova, Olga; Janssen, T. J. B. M.; Fal'Ko, Vladimir; Kubatkin, Sergey (2010). "Towards a quantum resistance standard based on epitaxial graphene". Nature Nanotechnology. 5 (3): 186–9. arXiv:0909.1220. Bibcode:2010NatNa...5..186T. doi:10.1038/nnano.2009.474. PMID 20081845.
  152. ^ a b Hass, J.; Varchon, F.; Millán-Otoya, J.; Sprinkle, M.; Sharma, N.; De Heer, W.; Berger, C.; First, P.; Magaud, L.; Conrad, E. (2008). "Why multilayer graphene on 4H-SiC(000(1)over-bar) behaves like a single sheet of graphene". Physical Review Letters. 100 (12): 125504. Bibcode:2008PhRvL.100l5504H. doi:10.1103/PhysRevLett.100.125504.
  153. ^ Singh, Ram Sevak; Nalla, Venkatram; Chen, Wei; Wee, Andrew Thye Shen; Ji, Wei (2011). "Laser Patterning of Epitaxial Graphene for Schottky Junction Photodetectors". ACS Nano. 5 (7): 5969–75. doi:10.1021/nn201757j. PMID 21702443.
  154. ^ "A smarter way to grow graphene". PhysOrg.com. May 2008.
  155. ^ Pletikosić, I.; Kralj, M.; Pervan, P.; Brako, R.; Coraux, J.; n’Diaye, A.; Busse, C.; Michely, T. (2009). "Dirac Cones and Minigaps for Graphene on Ir(111)". Physical Review Letters. 102 (5): 056808. arXiv:0807.2770. Bibcode:2009PhRvL.102e6808P. doi:10.1103/PhysRevLett.102.056808.
  156. ^ a b Bae, S.; et al. (2010). "Roll-to-roll production of 30-inch graphene films for transparent electrodes". Nature Nanotechnology. 5 (8): 574–578. Bibcode:2010NatNa...5..574B. doi:10.1038/nnano.2010.132. PMID 20562870.
  157. ^ Rafiee, J.; Mi, X.; Gullapalli, H.; Thomas, A.V.; Yavari, F.; Shi, Y.; Ajayan, P.M.; Koratkar, N.A. (2012). "Wetting transparency of graphene". Nature Materials (11): 217–222.
  158. ^ Weatherup, R.S.; Bayer, Bernhard C.; Blume, Raoul; Ducati, Caterina; Baehtz, Carsten; Schlögl, Robert; Hofmann, Stephan (2011). "In Situ Characterization of Alloy Catalysts for Low-Temperature Graphene Growth". Nano Letters. 11 (10): 4154–60. Bibcode:2011NanoL..11.4154W. doi:10.1021/nl202036y. PMID 21905732.
  159. ^ a b Amini, Shaahin; Garay, Javier; Liu, Guanxiong; Balandin, Alexander A.; Abbaschian, Reza (2010). "Growth of Large-Area Graphene Films from Metal-Carbon Melts". Journal of Applied Physics. 108 (9): 094321. arXiv:1011.4081. Bibcode:2010JAP...108i4321A. doi:10.1063/1.3498815.
  160. ^ Li, Xuesong; et al. (2009). "Large-Area Synthesis of High-Quality and Uniform Graphene Films on Copper Foils". Science. 324 (5932): 1312–4. arXiv:0905.1712. Bibcode:2009Sci...324.1312L. doi:10.1126/science.1171245. PMID 19423775. {{cite journal}}: Explicit use of et al. in: |author2= (help)
  161. ^ Wassei, Jonathan K.; Mecklenburg, Matthew; Torres, Jaime A.; Fowler, Jesse D.; Regan, B. C.; Kaner, Richard B.; Weiller, Bruce H. (12 May 2012). "Chemical Vapor Deposition of Graphene on Copper from Methane, Ethane and Propane: Evidence for Bilayer Selectivity". Small. 8 (9): 1415–1422. doi:10.1002/smll.201102276. PMID 22351509.
  162. ^ Lenski, Daniel R.; Fuhrer, Michael S. (2011). "Raman and optical characterization of multilayer turbostratic graphene grown via chemical vapor deposition". Journal of Applied Physics. 110: 013720. arXiv:1011.1683. Bibcode:2011JAP...110a3720L. doi:10.1063/1.3605545.
  163. ^ Choucair, M.; Thordarson, P; Stride, JA (2008). "Gram-scale production of graphene based on solvothermal synthesis and sonication". Nature Nanotechnology. 4 (1): 30–3. Bibcode:2009NatNa...4...30C. doi:10.1038/nnano.2008.365. PMID 19119279.
  164. ^ "Samsung's graphene breakthrough could finally put the wonder material into real-world devices". ExtremeTech. 2014-04-07. Retrieved 2014-04-13.
  165. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1126/science.1252268 , please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1126/science.1252268 instead.
  166. ^ Brumfiel, G. (2009). "Nanotubes cut to ribbons New techniques open up carbon tubes to create ribbons". Nature. doi:10.1038/news.2009.367.
  167. ^ Kosynkin, D. V.; Higginbotham, Amanda L.; Sinitskii, Alexander; Lomeda, Jay R.; Dimiev, Ayrat; Price, B. Katherine; Tour, James M. (2009). "Longitudinal unzipping of carbon nanotubes to form graphene nanoribbons". Nature. 458 (7240): 872–6. Bibcode:2009Natur.458..872K. doi:10.1038/nature07872. PMID 19370030.
  168. ^ Liying, Jiao; Zhang, Li; Wang, Xinran; Diankov, Georgi; Dai, Hongjie (2009). "Narrow graphene nanoribbons from carbon nanotubes". Nature. 458 (7240): 877–80. Bibcode:2009Natur.458..877J. doi:10.1038/nature07919. PMID 19370031.
  169. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1039/C1JM11227A , please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1039/C1JM11227A instead.
  170. ^ "Graphene Uses & Applications". Graphenea. Retrieved 2014-04-13.

Sources

Template:Link FA