10
LETTERS PUBLISHED ONLINE: 20 NOVEMBER 2011 | DOI: 10.1038/NMAT3170 Atomic-scale transport in epitaxial graphene Shuai-Hua Ji * , J. B. Hannon, R. M. Tromp, V. Perebeinos, J. Tersoff and F. M. Ross * The high carrier mobility of graphene 1–4 is key to its applications, and understanding the factors that limit mobility is essential for future devices. Yet, despite significant progress, mobilities in excess of the 2 × 10 5 cm 2 V -1 s -1 demonstrated in free-standing graphene films 5,6 have not been duplicated in conventional graphene devices fabricated on substrates. Understanding the origins of this degradation is perhaps the main challenge facing graphene device research. Experiments that probe carrier scattering in devices are often indirect 7 , relying on the predictions of a specific model for scattering, such as random charged impurities in the substrate 8–10 . Here, we describe model-independent, atomic-scale transport mea- surements that show that scattering at two key defects— surface steps and changes in layer thickness—seriously de- grades transport in epitaxial graphene films on SiC. These measurements demonstrate the strong impact of atomic-scale substrate features on graphene performance. Our results are based on scanning tunnelling potentiometry to measure local electric potential as current flows through a graphene film. By measuring local perturbations caused by substrate steps and changes in graphene thickness, we demonstrate that such heterogeneity is critical to transport in graphene. Substrate steps alone can increase the resistivity several-fold relative to a perfect terrace, and direct calculation shows that resistance arising from the intrinsic wavefunction mismatch will always exist at junctions between monolayer and bilayer graphene. The performance of graphene devices on SiC surfaces is thus fundamentally limited by the ability to control both the layer thickness and substrate perfection. Figure 1a and b show low-energy electron microscopy (LEEM) images obtained immediately after graphene growth on two SiC(0001) substrates with different step densities (see Methods). The graphene thickness can be determined straightforwardly with LEEM from the reflectivity of the low-energy electrons, which depends on thickness through quantum confinement effects 11 . In Fig. 1a the sample consists of 80% monolayer graphene (grey regions labelled 1), 10% ‘buffer’ layer that has a C- rich 6 3 × 6 3 structure (white regions) and 10% bilayer graphene (dark regions). In Fig. 1b the fractions of monolayer and bilayer graphene (labelled 1 and 2 respectively) are almost equal. Identifying by atomic-resolution scanning tunnelling microscopy (STM) the same areas that were imaged by LEEM enables us to characterize monolayer, bilayer and buffer-layer graphene. Consistent with other work 12–14 , each has a distinctive appearance (Fig. 1c,d). Because substrate and graphene step configurations can be determined through STM height measurements, we can obtain a comprehensive picture of nanoscale topography. In Fig. 1e, we show the experimental approach used to obtain maps of the electrical potential through scanning tunnelling potentiometry 15,16 . Two static probes (1 and 3) contact the surface at a separation of 500 μm. A voltage applied between these probes IBM T. J. Watson Research Center, Yorktown Heights, New York 10598, United States. *e-mail: [email protected]; [email protected]. induces current flow through the graphene sheet, while a third, scanning probe (2) measures the local potential. The potential can be measured on the macroscale, by stepping probe 2 across the surface, or on the microscale, by scanning probe 2 over a smaller area, in which case the topography of the sample can be acquired simultaneously. (Note that traditional four-probe measurements require all probes to contact the surface.) Figure 1f shows the macroscale potential. The total resistance (including the contact resistance and the resistance of the graphene sheet) and total current passing through the graphene are also measured. The potential distribution in this two-dimensional system is then modelled 16 as a Laplace problem with fixed boundary conditions at the tips (see Methods). By fitting the potential acquired along the line between probes 1 and 3, macroscopic conductivities σ avg can be determined for the two samples (Table 1). In Fig. 2, we show microscale potential measurements over regions measuring hundreds of nanometres. The topography and graphene thicknesses are shown in Fig. 2a. Without applying a voltage between probes 1 and 3, the potential map (Fig. 2d) is almost featureless, as expected. But when a voltage is applied (Fig. 2b,e), the maps show two distinct features: dramatic potential jumps at the step edges, and a potential gradient on the terraces. These effects change sign when the applied voltage is reversed (Fig. 2c,f), showing that the measurement is directly related to transport. The terrace gradient demonstrates that graphene terraces have a finite conductivity, presumably due to random scattering sources at the terraces (such as defects 13 , long-range scattering 7–10 or phonons 6 ) and at the interface 17 , but the potential discontinuity at the step edges indicates additional scattering at these locations. Car- rier scattering seems to be particularly strong at the heterogeneous junctions between monolayer and bilayer graphene, and weaker but still visible at locations where a uniform graphene bilayer crosses a substrate step (top right corner of each map). Potential profiles across two terraces and a monolayer–bilayer junction are shown for a series of applied voltages in Fig. 2i,j. The linear relationships between the terrace gradient and monolayer–bilayer jump and the applied voltage are demonstrated in Fig. 2k. On the terraces, the linear dependence of slope on applied voltage suggests that the terraces are behaving Ohmically: the local electric field E (the potential change per unit length) is related to the local current density j by j = σ t E , where σ t is a constant (but local) terrace conductivity. We cannot measure the current density locally. However, we can estimate it by noting that all measurements were made approximately half way between the fixed probes, where the average current density can be calculated from the measured total current using the Laplace equation above. The local current density may differ somewhat from the average current density, but this approach enables us to estimate local terrace conductivities for monolayer graphene (Table 1). (Bilayer graphene shows similar resistivity to monolayer graphene, as can be inferred for example from Fig. 2i,j.) On the basis of 114 NATURE MATERIALS | VOL 11 | FEBRUARY 2012 | www.nature.com/naturematerials © 2012 Macmillan Publishers Limited. All rights reserved

Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

LETTERSPUBLISHED ONLINE: 20 NOVEMBER 2011 | DOI: 10.1038/NMAT3170

Atomic-scale transport in epitaxial grapheneShuai-Hua Ji*, J. B. Hannon, R. M. Tromp, V. Perebeinos, J. Tersoff and F. M. Ross*

The high carrier mobility of graphene1–4 is key to itsapplications, and understanding the factors that limit mobilityis essential for future devices. Yet, despite significant progress,mobilities in excess of the 2× 105 cm2 V−1 s−1 demonstratedin free-standing graphene films5,6 have not been duplicatedin conventional graphene devices fabricated on substrates.Understanding the origins of this degradation is perhaps themain challenge facing graphene device research. Experimentsthat probe carrier scattering in devices are often indirect7,relying on the predictions of a specific model for scattering,such as random charged impurities in the substrate8–10. Here,we describe model-independent, atomic-scale transport mea-surements that show that scattering at two key defects—surface steps and changes in layer thickness—seriously de-grades transport in epitaxial graphene films on SiC. Thesemeasurements demonstrate the strong impact of atomic-scalesubstrate features on graphene performance.

Our results are based on scanning tunnelling potentiometryto measure local electric potential as current flows througha graphene film. By measuring local perturbations caused bysubstrate steps and changes in graphene thickness, we demonstratethat such heterogeneity is critical to transport in graphene. Substratesteps alone can increase the resistivity several-fold relative toa perfect terrace, and direct calculation shows that resistancearising from the intrinsic wavefunction mismatch will alwaysexist at junctions between monolayer and bilayer graphene.The performance of graphene devices on SiC surfaces is thusfundamentally limited by the ability to control both the layerthickness and substrate perfection.

Figure 1a and b show low-energy electron microscopy (LEEM)images obtained immediately after graphene growth on twoSiC(0001) substrates with different step densities (see Methods).The graphene thickness can be determined straightforwardly withLEEM from the reflectivity of the low-energy electrons, whichdepends on thickness through quantum confinement effects11.In Fig. 1a the sample consists of ∼80% monolayer graphene(grey regions labelled 1), ∼10% ‘buffer’ layer that has a C-rich 6

√3 × 6

√3 structure (white regions) and ∼10% bilayer

graphene (dark regions). In Fig. 1b the fractions of monolayer andbilayer graphene (labelled 1 and 2 respectively) are almost equal.Identifying by atomic-resolution scanning tunnelling microscopy(STM) the same areas that were imaged by LEEM enables usto characterize monolayer, bilayer and buffer-layer graphene.Consistent with other work12–14, each has a distinctive appearance(Fig. 1c,d). Because substrate and graphene step configurations canbe determined through STM height measurements, we can obtain acomprehensive picture of nanoscale topography.

In Fig. 1e, we show the experimental approach used to obtainmaps of the electrical potential through scanning tunnellingpotentiometry15,16. Two static probes (1 and 3) contact the surfaceat a separation of∼500 µm. A voltage applied between these probes

IBM T. J. Watson Research Center, Yorktown Heights, New York 10598, United States. *e-mail: [email protected]; [email protected].

induces current flow through the graphene sheet, while a third,scanning probe (2) measures the local potential.

The potential can be measured on the macroscale, by steppingprobe 2 across the surface, or on the microscale, by scanning probe2 over a smaller area, in which case the topography of the samplecan be acquired simultaneously. (Note that traditional four-probemeasurements require all probes to contact the surface.) Figure 1fshows the macroscale potential. The total resistance (includingthe contact resistance and the resistance of the graphene sheet)and total current passing through the graphene are also measured.The potential distribution in this two-dimensional system is thenmodelled16 as a Laplace problem with fixed boundary conditions atthe tips (see Methods). By fitting the potential acquired along theline between probes 1 and 3, macroscopic conductivities σavg can bedetermined for the two samples (Table 1).

In Fig. 2, we show microscale potential measurements overregions measuring hundreds of nanometres. The topography andgraphene thicknesses are shown in Fig. 2a. Without applying avoltage between probes 1 and 3, the potentialmap (Fig. 2d) is almostfeatureless, as expected. But when a voltage is applied (Fig. 2b,e),the maps show two distinct features: dramatic potential jumps atthe step edges, and a potential gradient on the terraces. These effectschange sign when the applied voltage is reversed (Fig. 2c,f), showingthat themeasurement is directly related to transport.

The terrace gradient demonstrates that graphene terraces have afinite conductivity, presumably due to random scattering sourcesat the terraces (such as defects13, long-range scattering7–10 orphonons6) and at the interface17, but the potential discontinuity atthe step edges indicates additional scattering at these locations. Car-rier scattering seems to be particularly strong at the heterogeneousjunctions betweenmonolayer and bilayer graphene, and weaker butstill visible at locations where a uniform graphene bilayer crossesa substrate step (top right corner of each map). Potential profilesacross two terraces and a monolayer–bilayer junction are shownfor a series of applied voltages in Fig. 2i,j. The linear relationshipsbetween the terrace gradient and monolayer–bilayer jump and theapplied voltage are demonstrated in Fig. 2k.

On the terraces, the linear dependence of slope on appliedvoltage suggests that the terraces are behaving Ohmically: the localelectric field E (the potential change per unit length) is relatedto the local current density j by j = σtE, where σt is a constant(but local) terrace conductivity. We cannot measure the currentdensity locally. However, we can estimate it by noting that allmeasurements were made approximately half way between thefixed probes, where the average current density can be calculatedfrom the measured total current using the Laplace equationabove. The local current density may differ somewhat from theaverage current density, but this approach enables us to estimatelocal terrace conductivities for monolayer graphene (Table 1).(Bilayer graphene shows similar resistivity to monolayer graphene,as can be inferred for example from Fig. 2i,j.) On the basis of

114 NATURE MATERIALS | VOL 11 | FEBRUARY 2012 | www.nature.com/naturematerials

© 2012 Macmillan Publishers Limited. All rights reserved

Page 2: Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

NATURE MATERIALS DOI: 10.1038/NMAT3170 LETTERS

1

2

1

500 nm

Probe 2

200 µm

2 µm

2 nm

Probe 1 Probe 3¬400

¬300

¬400

¬100

¬200

100

0

300

400

200

¬200 0 200 400

Distance (µm)

Pote

ntia

l (m

V)

a

c e f

d

b

Figure 1 | LEEM, STM and macroscale potential measurement of graphene on SiC. a, Bright-field LEEM image of graphene on 0.06◦-miscut 4H SiC(0001),field of view 10 µm. b, Bright-field LEEM image of graphene on 0.5◦-miscut 6H SiC(0001), field of view 3 µm. c, STM of monolayer graphene (STM biasvoltage V=−0.05 V, tunnelling current I=0.1 nA) showing a honeycomb structure with a moiré pattern; roughness induced by the interface also evident.d, STM of bilayer graphene (V=−0.05 V, I=0.1 nA) showing a hexagonal structure and smoother surface. e, Scanning electron microscopy image of thescanning tunnelling potentiometry experiment. An a.c. voltage V13 (Vr.m.s.= 2 mV, frequency f= 2 kHz) is applied between fixed probes 1 and 3 to maintainscanning probe 2 in the tunnelling range, measuring through a feedback loop the potential when no net current flows at probe 2. f, Potential measured (at∼72 K) by stepping probe 2 across the surface to points along the dashed line in e with positions determined with scanning electron microscopy. Blacksquare data points were measured on the low-miscut sample at V13= 252 mV; the total current is 0.906 mA and hence resistance is 276�. Red triangulardata points are for the high-miscut sample, where the measured resistance is larger (722�) at the same probe spacing; V13=659 mV was used to achievethe same total current. Solid lines are the results of fitting a two-dimensional calculation of current flow using probe contact areas of 60 µm.

scanning tunnelling spectroscopy results also obtained on thesesamples (Supplementary Fig. S1) and angle-resolved photoemissionspectroscopymeasurements from the literature18,19, we estimate theelectron density in the monolayer graphene to be ∼1013 cm−2 andthe local mobility for monolayer graphene on terraces at 72 K to be∼3,000 cm2 V−1 s−1 in both samples.

The monolayer–bilayer graphene junction also obeys Ohm’s law(Fig. 2k): the linear dependence of voltage jump 1V on appliedvoltage and hence local electric field indicates that 1V is alsoproportional to the local current density, that is, 1V ∝ j. With thelocal current density j estimated as above, the monolayer–bilayerjunction resistanceρstep can be extracted (Fig. 3d) usingV = jρstep.

Where a single, continuous layer of graphene crosses a substratestep, the effect is weaker but still quantifiable. In Fig. 3a,b, single-layer graphene crosses 0.5-nm-high substrate steps. Although thepotential discontinuity at the steps is hard to discern in individualscan lines (Fig. 3b), averaging shows the magnitude of the effect(Fig. 3c). It is also clear from Fig. 3c that higher steps show a greaterpotential jump (for similar terrace gradient and hence local currentdensity). Ohm’s law is followed (Supplementary Fig. S2), so valuesfor the step resistance ρstep can be calculated. Although changes

Table 1 |Conductivity of graphene on SiC substrates withdifferent miscut angles.

Low miscut(∼0.06◦)

High miscut(0.5◦)

Average conductivity σavg

from macroscalemeasurement (mS)

4.32±0.09 1.46±0.03

Conductivity of monolayergraphene σt from microscalemeasurement (mS)

5.0±0.7 4.4±0.7

in graphene conductance near steps have been described20, thistechnique provides a quantitativemeasure of the extra resistance.

Figure 3d summarizes the results for monolayer graphenecrossing substrate steps, and formonolayer–bilayer junctions of dif-ferent configurations.Monolayer graphene crossing single (0.5 nm)substrate steps shows a resistance of 6.9±2.9� µm. The resistanceseems to increase linearly with step height, 14.9± 3.6� µm for

NATURE MATERIALS | VOL 11 | FEBRUARY 2012 | www.nature.com/naturematerials 115

© 2012 Macmillan Publishers Limited. All rights reserved

Page 3: Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

LETTERS NATURE MATERIALS DOI: 10.1038/NMAT3170

1

2

1

2

2

100 nm

0 nm 2.7 nm

s1

s2

s3

s2 = 0.69 nm

s1 = 1.10 nm

s3 = 1.55 nm

¬0.6 mV 0.6 mV

1,785

1,530

1,275

1,020

765

510

255

0

¬1,785

¬1,530

¬1,275

¬1,020

¬765

¬510

¬255

0

¬300 ¬200 ¬100 0Distance (nm)

¬300 ¬200 ¬100 0Distance (nm)

¬2 ¬1 0 1 2Voltage V13 (V)

0

0.5

1.0

1.5

2.0

2.5

3.0

Pote

ntia

l (m

V) 0

0.5

¬0.5

1.0

¬1.0

1.5

¬1.5

0

0.5

¬0.5

1.0

¬1.0

1.5

¬1.5

Ter

race

fiel

d (m

V µ

1 ) Jump at step (m

V)

a b c

d e f

g h

0

0.5

1.0

1.5

2.0

2.5

3.0

Pote

ntia

l (m

V)

kji

Figure 2 | Scanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography (a–c) and potential maps (d–f) recordedsimultaneously (a.c. voltage Va.c.= 2 mV,I= 50 pA; room temperature) with V13=0 (a,d), V13= 1.53 V and total current 5.73 mA (b,e) andV13=−1.53 V (c,f). The voltage range is shown on the colour scale; zero is arbitrary. The step heights and graphene thickness (labels in a and c) areidentified from STM. The local current density midway between the fixed probes is 4× 10−6 A µm−1. g,h, Simulated potential maps calculated using theexperimental boundary potential conditions. Step edges s1, s2 and s3 have resistances of 41, 13 and 69� µm respectively. i,j, Line profiles of the potentialaveraged from the rectangle in b, shown for the V13 values indicated (mV). Data are offset vertically for clarity. The terrace slopes on the monolayer andbilayer sides are similar for each voltage, suggesting that the resistivity of the bilayer graphene is similar to that of the monolayer graphene in this region.k, The electric field on the terraces (slopes in i and j; monolayer and bilayer terraces being similar) and the potential jump at the monolayer–bilayer junction(jump heights in i and j) as a function of V13.

1.0 nm steps and 24.7±4.3� µm for 1.5 nm steps. One example ofbilayer graphene crossing a step is also included, and it follows thesame trend. Monolayer–bilayer junctions have a higher resistance,20.9±5.7� µm and 28.4±7.0� µm for planar and stepped junc-tions respectively. Monolayer–bilayer junctions at a double-heightstep provide the highest resistance seen here, 88� µm.

We find that the upper graphene layer is continuous over themonolayer–bilayer junction (Supplementary Fig. S1b), consistentwith other reports14. Given the continuous nature of the graphenesheet, the substantial resistance of the junction is perhaps counter-intuitive, and might suggest the presence of defects or scatterers atthe graphene edge. To understand this, we calculated the resistance

116 NATURE MATERIALS | VOL 11 | FEBRUARY 2012 | www.nature.com/naturematerials

© 2012 Macmillan Publishers Limited. All rights reserved

Page 4: Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

NATURE MATERIALS DOI: 10.1038/NMAT3170 LETTERS

100 nm

0 nm 1.0 nm ¬0.99 mV 0.92 mV

s

s

s

1.59

1.07

1.01

1.34

1.03

1.5 nm

1.0 nm

0.5 nm

0.93 mV μm¬1

¬0.2

0

0

20

40

60

80

100

0.2

0.4

0.6

0.8

1.0

1.2

Resi

stan

ce (

Ω μ

m)

Pote

ntia

l (m

V)

0 50 100 150 200 250 300

Distance (nm)

0 0.5 1.0 1.5

Topographic step height s (nm)

a b

c d

Figure 3 | Single-layer graphene overlaying substrate steps. a,b, STM topography (a) and scanning tunnelling potentiometry (b) recorded simultaneouslyat V13=−1.53 V; data measured at low temperature (∼72 K) to reduce scan noise. The local current density is 6.4× 10−6 A µm−1. c, Potential profiles ofmonolayer graphene over single-, double- and triple-height substrate steps. The values of the terrace gradients are indicated to give an idea of variability inthe data and enable us to calculate that 0.5, 1.0 and 1.5 nm steps contribute resistance equivalent to∼40, 80 and 120 nm of monolayer terrace widthrespectively. d, The resistance of configurations with different numbers of layers and topographic step heights s. Inset diagrams show the configuration,with the buffer layer omitted for clarity. The bilayer is typically at the lower terrace. All data were obtained at low temperature except for the bilayer(triangular) data point, and all data are from the low-miscut sample except for the 88� µm data point, which is from a different but nominally identicalsample, and the bilayer data point, which is from the high-miscut sample. For monolayer graphene over 0.5, 1.0, 1.25 and 1.5 nm steps, seven, five, two andthree steps respectively were measured. For monolayer–bilayer junctions with 0 and 0.5 nm topographic height changes, two and five steps respectivelywere measured. Vertical and horizontal error bars are root mean squared values calculated from variations in resistance and in measured step heightrespectively. For the 88� µm data point only one step was measured and no error bar is shown.

using a standard tight-binding model of the monolayer and bilayerwavefunctions21. For energies near the Dirac point, a continuumapproximation is generally considered adequate22,23. However, inview of the considerable doping here, we use a full atomisticapproach with exact boundary conditions (effectively equivalent tothe non-equilibriumGreen function approach used for overlappingnanoribbons24; see Supplementary Information for details). Themodel has five parameters: in-plane and interlayer matrix elementst , bilayer bandgap ∆ and the respective doping levels EF, allof which are known for epitaxial graphene on SiC (refs 19,21).With these values (tin−plane = 3.1 eV,tinterlayer = 0.4 eV,EF = 0.45 eVin monolayer, 0.3 eV in bilayer, ∆ = 0.15 eV), we calculateresistances ∼25� µm and ∼14� µm for junctions with armchairand zigzag orientation, respectively. The junction measured haspredominantly armchair orientation. The agreement betweencalculated and measured values is striking, although this couldbe partly fortuitous given the uncertainties in doping level andbandgap. These results clearly show that a high resistance is anintrinsic property of an idealmonolayer–bilayer junction.

The 25� µm resistance corresponds to an average transmissionfactor of T ≈ 1

2 , consistent with that found at low doping22,23.We find that the poor transmission is largely a result of thewavefunction mismatch between monolayer and bilayer, unavoid-able because the bilayer wavefunctions have large amplitude onboth layers: wavefunction matching requires intermixing withan evanescent state from a higher-energy band of the bilayer.Thus, wavefunction mismatch is an inherent characteristic of thisinterface, and calculations using standard methods confirm that

this inherent mismatch is sufficient to account for the magnitudeof resistance observed experimentally. We also find consider-able interband (K–K′) scattering for the armchair orientation,so chirality is not conserved13. Even for the zigzag orientation,where by symmetry there is no K–K′ scattering, there is still astrong wavefunction mismatch; and the calculations suggest a largeresistance, although the conservation of chirality for the zigzagorientation contributes to the lower resistance when compared withthe armchair orientation.

For a continuous graphene layer going over a substrate step, wesuggest that the origin of the step-induced resistance may be in-trinsic, induced for example by σ–π hybridization arising from thecurvature of the graphene sheet near the top and bottomof the step.

The combination of potential mapping and modelling hasthus shown that substrate steps, terraces and thickness changesall contribute to the resistance. Each element can be treatedas following Ohm’s law at the nanoscale. Steps and junctionsstrongly affect transport. A 0.5 nm substrate step contributes extraresistance equivalent to a ∼40-nm-wide terrace. 1.0- and 1.5-nm-high substrate steps contribute resistance equivalent to ∼80 and∼120 nm of terrace respectively, and monolayer–bilayer junctionscontribute ∼100 nm or more. We can verify this understandingof current flow by simulating the potential distribution across thesample. In Fig. 2g,h, we show a simulation with fixed boundarypotentials taken from the experimental data in Fig. 2a–f. The resultsare in close agreement with the data, enabling resistances to beestimated that are consistent with the values obtained above fromanalysis of line profiles.

NATURE MATERIALS | VOL 11 | FEBRUARY 2012 | www.nature.com/naturematerials 117

© 2012 Macmillan Publishers Limited. All rights reserved

Page 5: Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

LETTERS NATURE MATERIALS DOI: 10.1038/NMAT3170

100 nm

¬0.71 mV 0.75 mV

0.5 nm

1.25 nm

0 nm 3.4 nm

a b

c

Figure 4 |Details of current flow around steps. a, Topography ofmonolayer graphene across substrate steps and a 1.25-nm-deep vacancyisland (V=0.2 V, I=0.1 nA). b, Potential map (385 nm×264 nm,V13= 1.53 V,Va.c.= 2 mV,I= 50 pA). c, Simulation of the same regionshowing equipotential contours and electron flow (arrows). The localcurrent density is 6.9× 10−6 A µm−1.

Macroscopic conductivity measurements thus provide only asmall part of the full picture required to understand transportthrough graphene on SiC. Miscut steps, islands formed duringthe growth process and thickness variations will all reduce themacroscopic conductivity. For example, the two samples in Table 1show similar (within 11%) local conductivities measured onsingle terraces. Yet in the high-miscut sample the macroscopicconductivity was ∼3× lower than the local value, whereas thelow-miscut sample showed a difference of only ×1.2. Steps fromeven a 0.7◦ miscut should double the resistance in the miscutdirection and lead to anisotropy in transport, as already seenby four-probe measurements25. We believe that these results arealso relevant to graphene on other substrates. Local variations inconductivity have already been seen within and between grapheneflakes on SiO2 (ref. 26). But even for a structurally perfect graphenelayer, substrate-induced steps may reduce mobility on substratessuch asmica27 and boron nitride28. As the properties and processingof dielectrics such as SiO2 are improved, the effects of thicknesschanges and distortions to the graphene sheet will become moreimportant in determining device properties.

We finally consider the possibilities suggested by local controlof current flow. The steps and boundaries analysed so far weregenerally perpendicular to the current. In Fig. 4a we show howcurrent flows in two dimensions when the resistive features havea more complex geometry, here a closed loop. Such features distortthe flow of current (Fig. 4b), as seen in the simulated equipotentialcontours and the small reduction in current density inside the island(Fig. 4c). It is interesting to speculate whether spatial control of stepconfigurations could be used to concentrate current into specificregions of a graphene sheet for new device designs.

MethodsGraphene growth and imaging. Growth of graphene layers took place in anultrahigh-vacuum low-energy electron microscope, electron-beam heating thesample to 1,300 ◦C while exposing undoped Si-terminated SiC(0001) wafers to5×10−6 torr disilane29,30. The instrument used was the IBM LEEM-II (refs 29,30).Growth in situ enabled the step density and configuration of the sample to bedetermined and the graphene thickness monitored during and after growth.Growth was terminated when samples showed large area fractions of monolayergraphene as well as smaller fractions of bilayer graphene. Some samples alsocontained lithographically defined patterns so that the same areas could be imagedin LEEM and STM. After graphene growth, samples were transferred throughair to an ultrahigh-vacuum system containing a four-probe, low-temperaturescanning tunnelling microscope and a scanning electron microscope. The scanningtunnelling microscope is a Unisoku UHV-LT four-probe, low-temperaturescanning tunnelling microscope operated with an RHK SPM-1000 controller.Scanning electron microscopy images were acquired using an FEI 2LE Schottky

column. Brief heating to 600 ◦C in a preparation chamber was used to desorbspecies such as water before loading into the STM stage. Scanning tunnellingmicroscopy and spectroscopy as well as scanning tunnelling potentiometry15,16were then carried out, either at room temperature or at 72 K. In scanning tunnellingpotentiometry, two probes set up a potential difference between two points on thesample, while a third probe measures the local potential between the two probes.The local potential is determined in a tunnelling experiment, by measuring thecondition in which no tunnelling current flows (that is zero potential differencebetween the tip and the location on the sample at which the tip is positioned).Contact potential thus does not play a role. Furthermore, the contact resistanceat the fixed probes only determines the potential drop at the contact regionsand has no effect on results from local potential measurement. The data wereacquired with a PtIr tip, cleaned in situ by approaching a metallic target whileapplying a high electric field, and using home-built electronics based on thecircuit described in ref. 15.

Macroscale conductivity measurement. We assume that the graphene sheet hasa uniform conductivity on the macroscale. In this two-dimensional system, thevoltage on the surface should satisfy ∇2V = 0. Assuming point contacts for thefixed probes, and taking the probe positions as (d/2,0) and (−d/2,0) and V = 0 atthe origin, the distribution of the potential is

V (x,y)=Aln

(√(x+ (d/2))2+y2√(x− (d/2))2+y2

)(1)

Integrating the current density along the y axis, we obtain the total current

I =∫+∞

−∞

∂xV (0,y)σavg dy =

∫+∞

−∞

Ad

(d/2)2+y2σavg dy = 2πAσavg

where σavg is the ‘macroscopic’ conductivity and A is a fitting constantwith units of voltage.

Finally, the ‘macroscopic’ current density at the origin is given by

j =4Aσavg

d=

2Iπd

We measure V in several places by stepping probe 2 along the line between probes1 and 3. Fitting the V values to equation (1) yields a value for A. We also measured and I directly. From A, d and I , we obtain the values for σavg and j used in theanalysis. In reality, the contact tips have finite size and irregular shape, withinwhich the potential is constant. The voltage difference between the two contacts isV13. To fit equation (1) we assume circular contacts of diameter∼60 µm estimatedfrom scanning electron microscopy.

Received 11 July 2011; accepted 13 October 2011; published online20 November 2011

References1. Novoselov, K. S. et al. Two-dimensional gas of massless Dirac fermions in

graphene. Nature 438, 197–200 (2005).2. Zhang, Y. B., Tan, Y-W., Stormer, H. L. & Kim, P. Experimental observation

of the quantum Hall effect and Berry’s phase in graphene. Nature 438,201–204 (2005).

3. Berger, C. et al. Electronic confinement and coherence in patterned epitaxialgraphene. Science 312, 1191–1196 (2006).

4. Avouris, P. Graphene: Electronic and photonic properties and devices.Nano Lett. 10, 4285–4294 (2010).

5. Du, X., Skachko, I., Barker, A. & Andrei, E. Y. Approaching ballistic transportin suspended graphene. Nature Nanotech. 3, 491–495 (2008).

6. Bolotin, K. I., Sikes, K. J., Hone, J., Stormer, H. L. & Kim, P.Temperature-dependent transport in suspended graphene. Phys. Rev. Lett. 101,096802 (2008).

7. Chen, J-H. et al. Intrinsic and extrinsic performance limits of graphene deviceson SiO2. Nature Nanotech. 3, 206–209 (2008).

8. Ando, T. Screening effect and impurity scattering in monolayer graphene.J. Phys. Soc. Jpn 75, 074716 (2006).

9. Hwang, E. H., Adam, S. & Das Sarma, S. Carrier transport in two-dimensionalgraphene layers. Phys. Rev. Lett. 98, 186806 (2007).

10. Cheianov, V. & Fal’ko, V. I. Friedel oscillations, impurity scattering, andtemperature dependence of resistivity in graphene. Phys. Rev. Lett. 97,226801 (2006).

11. Hibino, H. et al. Microscopic thickness determination of thin graphite filmsformed on SiC from quantized oscillation in reflectivity of low-energy electrons.Phys. Rev. B 77, 075413 (2008).

12. Mallet, P. et al. Electron states of mono- and bilayer graphene on SiC probedby scanning-tunnelling microscopy. Phys. Rev. B 76, 041403(R) (2007).

13. Rutter, G. M. et al. Scattering and interference in epitaxial graphene. Science317, 219–222 (2007).

118 NATURE MATERIALS | VOL 11 | FEBRUARY 2012 | www.nature.com/naturematerials

© 2012 Macmillan Publishers Limited. All rights reserved

Page 6: Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

NATURE MATERIALS DOI: 10.1038/NMAT3170 LETTERS14. Lauffer, P. et al. Atomic and electronic structure of few-layer graphene on

SiC(0001) studied with scanning tunnelling microscopy and spectroscopy.Phys. Rev. B 77, 155426 (2008).

15. Bannani, A., Bobisch, C. A. & Möller, R. Local potentiometry usinga multiprobe scanning tunnelling microscope. Rev. Sci. Intrum. 79,083704 (2008).

16. Homoth, J. et al. Electronic transport on the nanoscale: Ballistic transmissionand Ohm’s law. Nano Lett. 9, 1588–1592 (2009).

17. Rutter, G. M. et al. Imaging the interface of epitaxial graphene with siliconcarbide via scanning tunnelling microscopy. Phys. Rev. B 76, 235416 (2007).

18. Ohta, T. et al. Interlayer interaction and electronic screening in multilayergraphene investigated with angle-resolved photoemission spectroscopy.Phys. Rev. Lett. 98, 206802 (2007).

19. Zhou, S. Y. et al. Substrate-induced bandgap opening in epitaxial graphene.Nature Mater. 6, 770–775 (2007).

20. Nagase, M., Hibino, H., Kageshima, H. & Yamaguchi, H. Local conductancemeasurements of double-layer graphene on SiC substrate. Nanotechnology 20,445704 (2009).

21. Castro Neto, A. H., Guinea, F., Peres, N. M. R., Novoselov, K. S. & Geim, A. K.The electronic properties of graphene. Rev. Mod. Phys. 81, 109–162 (2009).

22. Nilsson, J., Castro Neto, A. H., Guinea, F. & Peres, N. M. R. Transmissionthrough a biased graphene bilayer barrier. Phys. Rev. B 76, 165416 (2007).

23. Nakanishi, T., Koshino, M. & Ando, T. Transmission through a boundarybetween monolayer and bilayer graphene. Phys. Rev. B 82, 125428 (2010).

24. González, J. W., Santos, H., Pacheco, M., Chico, L. & Brey, L. Electronictransport through bilayer graphene flakes. Phys. Rev. B 81, 195406 (2010).

25. Yakes, M. K. et al. Conductance anisotropy in epitaxial graphene sheetsgenerated by substrate interactions. Nano Lett. 10, 1559–1562 (2010).

26. Nirmalraj, P. N. et al. Nanoscale mapping of electrical resistivity andconnectivity in graphene strips and networks. Nano Lett. 11, 16–22 (2011).

27. Lui, C. H., Liu, L., Mak, K. F., Flynn, G. W. & Heinz, T. F. Ultraflat graphene.Nature 462, 339–341 (2009).

28. Dean, C. R. et al. Boron nitride substrates for high-quality graphene electronics.Nature Nanotech. 5, 722–726 (2010).

29. Tromp, R. M. & Hannon, J. B. Thermodynamics and kinetics of graphenegrowth on SiC(0001). Phys. Rev. Lett. 102, 106104 (2009).

30. Hannon, J. B. & Tromp, R. M. Pit formation during graphene synthesis onSiC(0001): In situ electron microscopy. Phys. Rev. B 77, 241404(R) (2008).

AcknowledgementsWe thank A. Ellis and M. C. Reuter of IBM for their assistance with experimental aspectsof this work, and R. Möller and X. Chen for discussions.

Author contributionsS-H.J. carried out scanning tunnelling potentiometry experiments, J.B.H. and R.M.T.grew the graphene and carried out LEEM; J.T. and V.P. carried out the calculations;S-H.J., F.M.R., J.B.H. and R.M.T. collaborated on equipment and experimental design;all authors wrote the paper.

Additional informationThe authors declare no competing financial interests. Supplementary informationaccompanies this paper on www.nature.com/naturematerials. Reprints and permissionsinformation is available online at http://www.nature.com/reprints. Correspondence andrequests for materials should be addressed to S-H.J. or F.M.R.

NATURE MATERIALS | VOL 11 | FEBRUARY 2012 | www.nature.com/naturematerials 119

© 2012 Macmillan Publishers Limited. All rights reserved

Page 7: Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

1

Supplementary information

Atomic Scale Transport in Epitaxial Graphene

Shuai-Hua Ji+, J. B. Hannon, R. M. Tromp, V. Perebeinos, J. Tersoff and F. M. Ross* IBM T. J. Watson Research Center, Yorktown Heights, NY 10598, USA

[email protected]; *[email protected] Electronic and atomic structure of graphene. We use scanning tunneling spectroscopy to characterize the electronic structure of the graphene. The differential conductance, dI/dV, measured by the lock-in technique, is related to the density of states of the graphene. The averaged dI/dV on the bilayer graphene shows a dip around -0.3V, Fig. S1(a), suggesting1 the Dirac point ED level is -0.3eV. This indicates the sample is electron doped, consistent with angle-resolved photoemission results2,3. For the monolayer graphene, the ED level can not be clearly resolved by dI/dV measurement, but the doping level of n≈1013cm-2 can be estimated from Refs. 2 and 3. High resolution STM imaging around the monolayer-bilayer junction shows a continuous lattice across the boundary, Fig. S1(b). Figure S1 Electronic and atomic structure of graphene. a, The density of states measured on bilayer graphene. Results are similar on the 0.5° and 0.06° samples. b, STM image of the monolayer-bilayer graphene junction (12nm×12nm, V=50mV, I=0.1nA). Monolayer graphene (1) shows the honeycomb structure while bilayer graphene (2) shows a hexagonal lattice due to the AB stacking. The continuity of the top graphene layer is visible. Inset is a schematic of the monolayer-bilayer graphene junction with buffer layer omitted for clarity.

SUPPLEMENTARY INFORMATIONDOI: 10.1038/NMAT3170

NATURE MATERIALS | www.nature.com/naturematerials 1

© 2011 Macmillan Publishers Limited. All rights reserved.

Page 8: Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

2

Potential profiles of monolayer graphene crossing steps. Fig S2 indicates the Ohmic behavior for monolayer graphene crossing substrate steps of different heights. Figure S2 Potential line profiles recorded for monolayer graphene across substrate steps. a, Step height 0.5nm; b, step height 1.5nm. The V13 values are indicated. The lines in a indicate the terrace gradient to highlight the small but measurable potential drop at the step. Calculation of transmission through a bilayer/monolayer junction. To find the transmission coefficient through a monolayer-bilayer junction we solve the scattering problem within a tight-binding approximation. The Hamiltonian equation H EΨ = Ψ describes the bilayer for atoms with coordinate x≤0 and the monolayer for x>0, according to Eqs. (36) and (5) of Ref. 4, respectively. We use the same notations as in Ref. 4, with tight-binding parameters t=tin-plane=3.1 eV, t’=0, γ1=tinterlayer=0.4 eV and γ3,4=0. In the scattering problem the component of wavevector parallel to the junction is conserved. The incoming wave is described by a wavevector whose component k0 normal to the junction has Im(k0)=0 and positive group velocity dE/dk0>0. Zig-zag edge bilayer/monolayer junction: For the zig-zag edge K and K′ states are not mixed, because a line with a constant ky can not cross both K and K′ points in the extended Brillouin zone. Therefore, there can only be two reflected x-components k1 and k2 in the bilayer, describing a wave in the first subband Im(k1)=0 with negative group velocity dE/dk1<0 and an evanescent wave (when the energy E is below the second subband as in our case) with Im(k2)>0, respectively. Note that k1≠-k0 because of trigonal warping. In the monolayer a transmitted wave with a positive group velocity has wavevector (k3, ky), where k3 is determined by having the same energy and ky as the incident wave, and by the condition Im(k3)=0 and dE/dk3>0. The general solution for the wavefunction has the form

3

( )( )( )( )

( )( )

( )( )( )( )

( )( )1 0 1

1

1 0 110

1,22 2 0 2

2 2 0 2

, ,( ), ,( )

exp exp( ) , ,( ) , ,

y i yB

y i yAx y y i i x y y

iA y i y

B y i y

b k k b k kRa k k a k kR

i k R k R r i k R k RR a k k a k kR b k k b k k

=

Ψ Ψ = + + + Ψ Ψ

(1a) for x≤0, and

( )( ) ( )( )3

3 3

3

,( )exp

( ) ,

yAx y y

B y

a k kRr i k R k R

R b k k

Ψ = + Ψ

(1b)

for x>0. Here R

is a primitive unit cell coordinate4 and for every atom in the bilayer, the

wavefunction at that atom is given by Eq. (1a), and for every atom in the monolayer, by Eq. (1b). The momentum-dependent coefficients (b1, a1, a2. b2) and (a,b) are fixed by satisfying the Hamiltonian equation H EΨ = Ψ for the bilayer and monolayer away from the boundary, as explained in Ref. 4. Applying the operator H-E to a wavefunction of the form Eq. (1) then gives zero contribution except at the interface, where there are terms depending on the scattering amplitudes ri=1-3. These amplitudes are found by requiring that these remaining interface contributions vanish, so that

0H EΨ − Ψ = everywhere, including the atoms at the boundary. Note that a similar approach applies for an incoming wave from a monolayer to bilayer, and results in an identical transmission coefficient. Armchair edge bilayer/monolayer junction: For the armchair edge, the propagation direction is along y and the conserved wavevector is kx. The K and K′ states are mixed now, because a constant kx line can cross both K and K′ points. The general solution for the wavefunction has the form:

( )( )( )( )

( )( )( )( )( )( )

( )( )1 0 11

1 0 110

1,42 0 22

2 0 22

, ,( ), ,( )

exp exp, ,( ), ,( )

x x iB

x x iAx x y i x x i y

ix x iA

x x iB

b k k b k kRa k k a k kR

i k R k R r i k R k Ra k k a k kRb k k b k kR

=

Ψ

Ψ = + + + Ψ Ψ

(2a) for y≤0, and

( )( ) ( )( )

5,6

,( )exp

( ) ,

x jAj x x j y

jB x j

a k kRr i k R k R

R b k k=

Ψ = + Ψ

(2b)

for y>0. Here for every atom in the bilayer, the wavefunction at that atom with coordinate y≤0 is given by Eq. (2a), and for every atom in monolayer with y>0 by Eq. (2b). The incoming

2 NATURE MATERIALS | www.nature.com/naturematerials

SUPPLEMENTARY INFORMATION DOI: 10.1038/NMAT3170

© 2011 Macmillan Publishers Limited. All rights reserved.

Page 9: Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

2

Potential profiles of monolayer graphene crossing steps. Fig S2 indicates the Ohmic behavior for monolayer graphene crossing substrate steps of different heights. Figure S2 Potential line profiles recorded for monolayer graphene across substrate steps. a, Step height 0.5nm; b, step height 1.5nm. The V13 values are indicated. The lines in a indicate the terrace gradient to highlight the small but measurable potential drop at the step. Calculation of transmission through a bilayer/monolayer junction. To find the transmission coefficient through a monolayer-bilayer junction we solve the scattering problem within a tight-binding approximation. The Hamiltonian equation H EΨ = Ψ describes the bilayer for atoms with coordinate x≤0 and the monolayer for x>0, according to Eqs. (36) and (5) of Ref. 4, respectively. We use the same notations as in Ref. 4, with tight-binding parameters t=tin-plane=3.1 eV, t’=0, γ1=tinterlayer=0.4 eV and γ3,4=0. In the scattering problem the component of wavevector parallel to the junction is conserved. The incoming wave is described by a wavevector whose component k0 normal to the junction has Im(k0)=0 and positive group velocity dE/dk0>0. Zig-zag edge bilayer/monolayer junction: For the zig-zag edge K and K′ states are not mixed, because a line with a constant ky can not cross both K and K′ points in the extended Brillouin zone. Therefore, there can only be two reflected x-components k1 and k2 in the bilayer, describing a wave in the first subband Im(k1)=0 with negative group velocity dE/dk1<0 and an evanescent wave (when the energy E is below the second subband as in our case) with Im(k2)>0, respectively. Note that k1≠-k0 because of trigonal warping. In the monolayer a transmitted wave with a positive group velocity has wavevector (k3, ky), where k3 is determined by having the same energy and ky as the incident wave, and by the condition Im(k3)=0 and dE/dk3>0. The general solution for the wavefunction has the form

3

( )( )( )( )

( )( )

( )( )( )( )

( )( )1 0 1

1

1 0 110

1,22 2 0 2

2 2 0 2

, ,( ), ,( )

exp exp( ) , ,( ) , ,

y i yB

y i yAx y y i i x y y

iA y i y

B y i y

b k k b k kRa k k a k kR

i k R k R r i k R k RR a k k a k kR b k k b k k

=

Ψ Ψ = + + + Ψ Ψ

(1a) for x≤0, and

( )( ) ( )( )3

3 3

3

,( )exp

( ) ,

yAx y y

B y

a k kRr i k R k R

R b k k

Ψ = + Ψ

(1b)

for x>0. Here R

is a primitive unit cell coordinate4 and for every atom in the bilayer, the

wavefunction at that atom is given by Eq. (1a), and for every atom in the monolayer, by Eq. (1b). The momentum-dependent coefficients (b1, a1, a2. b2) and (a,b) are fixed by satisfying the Hamiltonian equation H EΨ = Ψ for the bilayer and monolayer away from the boundary, as explained in Ref. 4. Applying the operator H-E to a wavefunction of the form Eq. (1) then gives zero contribution except at the interface, where there are terms depending on the scattering amplitudes ri=1-3. These amplitudes are found by requiring that these remaining interface contributions vanish, so that

0H EΨ − Ψ = everywhere, including the atoms at the boundary. Note that a similar approach applies for an incoming wave from a monolayer to bilayer, and results in an identical transmission coefficient. Armchair edge bilayer/monolayer junction: For the armchair edge, the propagation direction is along y and the conserved wavevector is kx. The K and K′ states are mixed now, because a constant kx line can cross both K and K′ points. The general solution for the wavefunction has the form:

( )( )( )( )

( )( )( )( )( )( )

( )( )1 0 11

1 0 110

1,42 0 22

2 0 22

, ,( ), ,( )

exp exp, ,( ), ,( )

x x iB

x x iAx x y i x x i y

ix x iA

x x iB

b k k b k kRa k k a k kR

i k R k R r i k R k Ra k k a k kRb k k b k kR

=

Ψ

Ψ = + + + Ψ Ψ

(2a) for y≤0, and

( )( ) ( )( )

5,6

,( )exp

( ) ,

x jAj x x j y

jB x j

a k kRr i k R k R

R b k k=

Ψ = + Ψ

(2b)

for y>0. Here for every atom in the bilayer, the wavefunction at that atom with coordinate y≤0 is given by Eq. (2a), and for every atom in monolayer with y>0 by Eq. (2b). The incoming

NATURE MATERIALS | www.nature.com/naturematerials 3

SUPPLEMENTARY INFORMATIONDOI: 10.1038/NMAT3170

© 2011 Macmillan Publishers Limited. All rights reserved.

Page 10: Atomic-scale transport in epitaxial graphene · 0.5 2.0 Potential (mV) i j k Figure 2 jScanning tunnelling potentiometry of terraces and monolayer–bilayer junctions. a–f, Topography

4

wave is described by the wavevector (kx, k0) with Im(k0)=0 and positive group velocity dE/dk0>0, and reflected waves in the first subband at K and K′ points have negative group velocities dE/dk1,2<0 with Im(k1,2)=0 and two evanescent modes in the second subbands at K and K′ with Im(k3,4)>0, correspondingly. In the monolayer, the two transmitted waves with a positive group velocity have wavevectors (kx, k5) and (kx, k6), where k5,6 are determined by having the same energy and kx as the incident wave, and by the condition: Im(k5,6)=0, dE/dk5,6>0. Applying the operator H-E to a wavefunction of the form Eq. (2) then gives zero contribution except at the interface, where there are terms depending on the scattering amplitudes ri=1-6. These amplitudes are found by requiring that these remaining interface contributions vanish, so that

0H EΨ − Ψ = everywhere. Note that a similar approach applies for scattering for an incoming wave from a monolayer to bilayer, and results in an identical transmission coefficient. References 1. Lauffer, P. et al. Atomic and electronic structure of few-layer graphene on

SiC(0001) studied with scanning tunneling microscopy and spectroscopy. Phys. Rev. B 77, 155426 (2008).

2. Zhou, S. Y. et al. Substrate-induced bandgap opening in epitaxial graphene. Nature Mat. 6, 770-775 (2007).

3. Ohta, T., Bostwick, A., Seyller, T., Horn, K. & Rotenberg, E. Controlling the electronic structure of bilayer graphene. Science 313, 951-954 (2006).

4. Castro Neto, A. H., Guinea F., Peres, N. M. R., Novoselov, K. S. & Geim, A. K. The electronic properties of graphene. Rev. Mod. Phys. 81, 109-162 (2009).

4 NATURE MATERIALS | www.nature.com/naturematerials

SUPPLEMENTARY INFORMATION DOI: 10.1038/NMAT3170

© 2011 Macmillan Publishers Limited. All rights reserved.