Sei sulla pagina 1di 7

Letter

pubs.acs.org/NanoLett

van der Waals Heterostructures with High Accuracy Rotational


Alignment
Kyounghwan Kim, Matthew Yankowitz, Babak Fallahazad, Sangwoo Kang, Hema C. P. Movva,
Shengqiang Huang, Stefano Larentis, Chris M. Corbet, Takashi Taniguchi, Kenji Watanabe,
Sanjay K. Banerjee, Brian J. LeRoy, and Emanuel Tutuc*,

Microelectronics Research Center, Department of Electrical and Computer Engineering, The University of Texas at Austin, Austin,
Texas 78758, United States

Physics Department, University of Arizona, Tucson, Arizona 85721, United States

National Institute for Materials Science, 1-1 Namiki Tsukuba, Ibaraki 305-0044, Japan
*
S Supporting Information

ABSTRACT: We describe the realization of van der Waals


(vdW) heterostructures with accurate rotational alignment of
individual layer crystal axes. We illustrate the approach by
demonstrating a Bernal-stacked bilayer graphene formed using
successive transfers of monolayer graphene akes. The Raman
spectra of this articial bilayer graphene possess a wide 2D band,
which is best t by four Lorentzians, consistent with Bernal
stacking. Scanning tunneling microscopy reveals no moire
pattern on the articial bilayer graphene, and tunneling
spectroscopy as a function of gate voltage reveals a constant
density of states, also in agreement with Bernal stacking. In
addition, electron transport probed in dual-gated samples reveals
a band gap opening as a function of transverse electric eld. To
illustrate the applicability of this technique to realize vdW heterostructuctures in which the functionality is critically dependent on
rotational alignment, we demonstrate resonant tunneling double bilayer graphene heterostructures separated by hexagonal
boron-nitride dielectric.
KEYWORDS: Two-dimensional, heterostructure, graphene, boron-nitride, resonant tunneling

V an der Waals (vdW) heterostructures consisting of two-


dimensional materials such as graphene, hexagonal boron-
nitride (hBN), and transition metal dichalcogenides (TMDs)
as crystallographic references. For example, the straight edges of
mechanically exfoliated graphene akes are assumed to be
parallel to the zigzag or armchair crystal orientations, and they
have gained increased scrutiny recently owing to novel can be dierentiated from each other using polarized Raman
electronic states and applications that may be realized in spectroscopy. 14 Although this method may provide a
these heterostructures.13 Examples include the observation of straightforward and nondestructive way to determine the
moire patterns,47 Hofstadter buttery in graphenehBN crystal orientation of the graphene akes, the precision of this
heterostructures with small twist angles,810 and resonant method for accurate rotational alignments is limited because
tunneling in rotationally aligned double layer heterostruc- the edges are not always straight, nor purely zigzag or armchair,
tures.1113 To best explore the electronic properties of vdW but often a mixture of both zigzag and armchair segments,
heterostructures, two key ingredients are necessary: a layer-by- which complicates the crystal axes identication. In addition,
layer transfer or pick-up process that minimizes or eliminates the nite edge size combined with optical resolution challenges
the incorporation of residues at the interface, and control of the limits the angle determination accuracy to about 2 for typical
crystal axes rotational alignment and registration of successively ake sizes.
transferred layers. The latter ingredient, while ubiquitous in In this study, we introduce a transfer technique that provides
epitaxially grown heterostructures, is arguably an important a better control of the heterostructure fabrication and in
obstacle in realizing electronically coupled, crystallographically particular allows for accurate crystal axes alignment. The new
registered vdW heterostructures using a layer-by-layer transfer technique does not impose any limits on the size or shape of
approach.
The rotational alignment of the crystal axes in vdW Received: December 24, 2015
heterostructures realized by successive layer transfers has Revised: January 30, 2016
been achieved thus far by using the straight edges of the akes Published: February 9, 2016

2016 American Chemical Society 1989 DOI: 10.1021/acs.nanolett.5b05263


Nano Lett. 2016, 16, 19891995
Nano Letters Letter

the akes as it relies solely on well-dened metallic alignment hemispherical handle fabrication can be found in the
marks on the substrate instead of the straight edges of the akes Supporting Information Figures S1 and S2.
for the accurate rotational alignment. We illustrate the Figure 2, panels ad describe the fabrication process ow of
technique by realizing Bernal stacked bilayer graphene using a vdW heterostructure consisting of two rotationally aligned
successive transfers of two monolayer graphene akes with monolayer graphene akes along with the corresponding
rotationally aligned crystal axes and also demonstrate resonant optical micrographs (Figure 2fh). This heterostructure is
tunneling in rotational aligned double bilayer graphene chosen because the rotational alignment between the two layers
heterostructures separated by hBN. can be probed postfabrication by Raman spectroscopy,
The rotationally aligned transfer relies on a controlled two- scanning tunneling microscopy (STM), and electron transport,
dimensional ake pick-up technique, in which a micron size with distinct signatures between a rotationally aligned or
ake can be selectively detached from the exfoliation substrate twisted bilayer. To create this articial bilayer graphene, we rst
using a small contact area hemispherical handle substrate, exfoliate monolayer graphene and hBN on separate SiO2/Si
which leaves the akes in proximity intact. Figure 1, panel a substrates. The monolayer graphene is conrmed by optical
contrast and Raman spectroscopy. A hemispherical handle
consisting of PVA-coated epoxy is used to rst pick-up the hBN
ake, which will serve as a substrate of the vdW heterostructure.
The hBN is then aligned partially with a graphene monolayer
on SiO2/Si under an optical microscope, the hemispherical
handle is brought in contact with the substrate, and it is
subsequently detached. Thanks to a stronger vdW bonding
between graphene and hBN on one hand, and graphene and
PVA on the other, the monolayer graphene ake is sectioned,
and the area making contact with the hBN is selectively
detached (Figure 2b). Alternatively, the starting ake can also
be sectioned before the transfer using, for example, electron
beam lithography (EBL) and etching or using an atomic force
microscope (AFM) tip. Figure 2, panel g shows the section of
the graphene ake remaining on the SiO2/Si substrate. Next,
the hemispherical handle is translated laterally, brought in
contact with the substrate, and detached to pick up the
remaining graphene section (Figure 2d). The resulting double
graphene layer on hBN heterostructure has a region where the
two graphene monolayers overlap. Thanks to the single crystal
nature of the monolayer graphene, combined with spatial
translation of the transfer process, the two graphene
monolayers are expected to have rotationally aligned crystal
axes.
Figure 3 describes the results of the Raman spectroscopy
measurements on the vdW heterostructure of Figure 2 after it
was transferred onto a SiO2 substrate with the graphene layers
facing up (see details in the Supporting Information Figure
S3a). Figure 3, panel a shows the optical micrograph of the
graphene/graphene/hBN vdW heterostructure where the edges
Figure 1. (a) Illustration of the hemispherical handle substrate used
for selective ake pick-up and transfer. (b, c) Schematic and (d, e) of individual graphene akes are marked with dashed lines. We
corresponding optical micrographs of the ake pick-up procedure. (b, use the graphene Raman 2D band full width at half-maximum
d) The handle substrate is brought into contact with an exfoliated hBN (FWHM) in the overlapping region for an initial assessment of
ake. The contact area marked by black dashed line is approximately the rotational alignment between the two monolayer graphene
100 m 100 m. (c) Schematic and (e) optical micrograph of the akes. For monolayer graphene, the 2D band FWHM is 24
hBN transfer to a handle substrate. cm1, while for a Bernal stacked bilayer graphene, the 2D band
FWHM is approximately twice as large (50 cm 1).15
illustrates the hemispherical, bulb-like handle substrate that can Furthermore, because Bernal stacked bilayer graphene has
be fabricated from a small epoxy or polydimethylsiloxane two electronic energy bands near the neutrality point, four
(PDMS) droplet deposited onto a transparent base substrate distinct scattering processes contribute to the Raman 2D band,
such as glass or a planar PDMS mold. The contact area which is best t by a superposition of four Lorentzians, and is
between the handle substrate and the ake is determined by the blue-shifted in comparison to monolayer graphene. The Raman
hemisphere radius and contact force and can be as small as 50 2D band FWHM of a twisted bilayer graphene is expected to be
m 50 m. The hemisphere can be coated with a adhesion- close to the nominal value of a Bernal stacked bilayer for small
improving polymer such as poly(methyl methacrylate) twist angles, and decrease with increasing the twist angle, albeit
(PMMA), poly(vinyl alcohol) (PVA), or polypropylene not monotonically.16 Figure 3, panel b shows the Raman 2D
carbonate (PPC). Figure 1, panels b and c illustrate band FWHM map of panel a sample, measured using a 532 nm
schematically the selective ake pick-up using a hemispherical excitation wavelength with a spot size of 1 m and a power of 5
epoxy handle, along with an example of an hBN ake pick-up mW. Figure 3, panel b data show two distinct regions in the
(Figure 1d), and transfer (Figure 1e). More details on the area covered by graphene akes: one region characterized by a
1990 DOI: 10.1021/acs.nanolett.5b05263
Nano Lett. 2016, 16, 19891995
Nano Letters Letter

Figure 2. Rotationally aligned graphene double layer realized by successive transfers from a monolayer graphene using a hemispherical handle
substrate. (a, e) Schematic of layer pick-up. The red box represents a zoom-in view of the hemispherical handle substrate. (bd) Schematics and (f
h) corresponding optical micrographs of successive stacking steps. Panels b and f illustrate a partial contact of the handle with the bottom graphene.
Panels c and g show the handle substrate release with one graphene section detached. Panels d and h illustrate the second contact with the handle
translated laterally to create an overlap region of the two graphene layers. The two images in panels fh represent the same optical micrographs, with
the contrast enhanced and the graphene contour marked in the right-hand image. Because the two graphene layers are obtained from a single grain,
their principal crystal axes remain aligned as long as the handle is not rotated with respect to the bottom substrate.

Figure 3. Raman spectroscopy characterization of rotationally aligned double layer graphene. (a) Optical micrograph of rotationally aligned double
layer graphene on hBN heterostructure. The blue and red dashed lines mark the boundaries of two monolayer graphene akes. (b) Contour plot of
the Raman 2D band FWHM corresponding to the optical micrograph of panel a. The data show a signicantly wider 2D band in the overlap region
in comparison to the nonoverlapped regions. (c) Example of Raman spectra acquired at the locations marked in panel b. The spectra at locations 1
and 3 show a 2D band FWHM of 25 cm1, while the spectrum at location 2, in the overlap region, has a 2D band FWHM of 50 cm1. The 2D
band spectra at locations 1 and 3 can be t with a single Lorentzian, while the spectrum at location 2 is best t by four Lorentzians, consistent with
Bernal stacked bilayer graphene.

2D band FWHM of 25 cm1, corresponding to the dierentiate Bernal stacked bilayer graphene from twisted
nonoverlapping sections of the monolayer graphene akes, bilayer with rotational misalignment angles smaller than 3.16
and one region characterized by a FWHM of 50 cm1, To better quantify the alignment of our vdW hetero-
corresponding to the overlapping section. The 2D band structures, the rotationally aligned double monolayer graphene
FWHM in the overlapping region is consistent with that of of Figure 3 was investigated using scanning probe microscopy.
Bernal stacked bilayer graphene, suggesting that the rotational Metal contacts were patterned to allow a bias to be applied
misalignment between the two graphene layers is small. between the sample and the substrate, which served as back-
Figure 3, panel c shows sample representative Raman spectra gate. We characterized this sample on the atomic scale using
of the three regions as indicated in Figure 3, panel b. Each STM and spectroscopy. All the STM measurements were
Raman spectrum shows three district peaks at 1364 cm1, performed in ultrahigh vacuum at a temperature of 4.5 K.
corresponding to the hBN E2g band, at 1580 cm1, Figure 4, panel a shows a 200 nm 200 nm topographic map
corresponding to graphene G band, and at 2700 cm1, taken on the overlapping bilayer region of the sample. No
corresponding to graphene 2D band. The 2D band at spots 1 moire pattern is observed, indicating that the two layers of
and 3 can be t with a single Lorentizan peak, as expected for graphene are in essentially perfect rotational alignment. We
monolayer graphene, while the 2D band probed in the acquired atomically resolved topography to further characterize
overlapping region cannot be t with a single Lorentzian the dierence between the monolayer and bilayer regions in
peak. The best t to the 2D band in the overlapping region is Figure 4, panels b and c, respectively. The topography is easily
achieved with four Lorentzians, each with FWHM of 24 cm1 distinguishable between the two, with the monolayer region
consistent with the 2D band of Bernal stacked bilayer graphene. appearing hexagonal and the overlapped, bilayer region
We note that in spite of the good agreement between the appearing roughly triangular as anticipated for Bernal-stacked
Raman spectra in the overlapping region, and that of Bernal bilayer graphene.17 Finally, we examined the spectroscopic
stacked bilayer graphene, Raman spectroscopy cannot clearly signatures of both the monolayer and bilayer regions.
1991 DOI: 10.1021/acs.nanolett.5b05263
Nano Lett. 2016, 16, 19891995
Nano Letters Letter

Figure 4. Atomic-scale topographic and electronic characterization. (a) STM topography of the bilayer region of the sample. No moire pattern is
observed, indicating that the two graphene layers are perfectly aligned. The scale bar is 30 nm. Atomically resolved topography of the (b) monolayer
and (c) bilayer regions of the sample. The appearance of the atomic lattice is easily distinguishable and appears hexagonal in the monolayer region
and triangular in the bilayer region. The scale bar is 1 nm for both. The sample voltage is +300 mV and tunnel current is 150 pA for all images.
Normalized (dI/dV)/(I/V) spectroscopy as a function of gate voltage acquired on the (d) monolayer and (e) bilayer regions of the sample. The
white dotted lines roughly mark the position of the (d) Dirac point and (e) charge neutrality point. (f) Contour plot of resistivity () measured as a
function of VBG and VTG in a dual-gated bilayer graphene realized by successive transfers of rotationally aligned monolayer graphene. The density and
E-eld axes are indicated with dashed white lines. (g) versus E measured at charge neutrality and at dierent temperatures.

Dierential conductance (dI/dV) spectroscopy measurements graphene monolayers, we fabricated and characterized a
were acquired by turning o the feedback circuit and measuring separate, dual gated sample encapsulated in hBN dielectric.
the tunneling current (I) while adding a small 5 mV signal to The heterostructure is realized using a similar process to Figure
the sample voltage (V) at a 563 Hz frequency. Figure 4, panels 4, panels ae sample, but by adding an extra hBN pick-up after
d and e show the normalized (dI/dV)/(I/V) spectroscopy as a the two consecutive monolayer graphene, which will serve as
function of gate voltage in the monolayer and bilayer regions of top gate dielectric. To complete the device fabrication, a top
the sample. The white dotted lines roughly mark the sample gate electrode is dened on the overlap region by EBL followed
voltage of the minimum signal in the normalized dI/dV plot, by metallization of Cr (4 nm)/Au (40 nm). By using the top
representing the position of the charge neutrality point (CNP). gate electrode and PMMA as etch mask, a Hall bar is dened by
In monolayer graphene, the CNP (or equivalently Dirac point) CHF3 plasma etching. A sequence of EBL, metal deposition,
disperses as the square root of gate voltage due to the linear and lift o is used to create metal contacts to the graphene
band structure of graphene.18 However, the band structure of bilayer.
Bernal-stacked bilayer graphene is quadratic at low energy so The dual-gated bilayer device allows independent control of
the CNP is instead expected to disperse linearly with gate the carrier density (n) and transverse electric eld (E) using the
voltage. Indeed, this is exactly what we observe in the top (VTG) and bottom gate (VBG) biases, which in turn
experimental gate bias maps. Assuming a back gate capacitance provides further insight into the bilayer graphene stacking
of 9.5 nF cm2, corresponding to the 280 nm SiO2 dielectric in because the resistivity () dependence on total carrier density
series with a 52 nm thick hBN substrate, we extract a Fermi and E-eld is distinctly dierent for a Bernal stacked bilayer
velocity in the monolayer region of vF = 1.05 106 m/s. For graphene compared to a twisted bilayer graphene. Indeed, a
the bilayer region, the CNP shifts roughly linearly in sample Bernal stacked bilayer graphene has a parabolic energy-
voltage by 1.9 mV per volt on the back gate. This is in momentum dispersion with a tunable band gap as a function
reasonably good agreement with prior studies of bilayer of the applied E-eld, while a twisted bilayer graphene behaves
graphene1921 and corresponds to a Fermi velocity of vF = largely as two decoupled graphene monolayers in which the
0.95 106 m/s (assuming an interlayer hopping energy 1 = applied E-eld displaces the charge neutrality points of the two
0.38 eV) or equivalently to a constant density of states 2m*/ monolayers. Figure 4, panel f shows the contour plot of as a
(2) corresponding to a four-fold degenerate parabolic function of VBG and VTG measured at 1.5 K. When measured as
energy-momentum dispersion with eective mass m* = a function of n, the resistivity has one maximum, evincing a
0.037m0; m0 is the bare electron mass, and is the reduced single charge neutrality point, which depends linearly on VBG
Planck constant. Taken together, the topographic and and VTG. Along the charge neutrality line, the resistivity
electronic signatures of the bilayer region strongly suggest it increases markedly with the transverse E-eld and shows an
is in a perfectly aligned, Bernal-stacked conguration. insulating temperature dependence as illustrated in Figure 4,
To further verify the realization of Bernal stacked bilayer panel g consistent with the band gap opening in Bernal stacked
graphene by successive transfers of rotationally aligned bilayer graphene.2224 In contrast, a twisted bilayer graphene
1992 DOI: 10.1021/acs.nanolett.5b05263
Nano Lett. 2016, 16, 19891995
Nano Letters Letter

Figure 5. Rotationally aligned double bilayer graphene heterostructure with interlayer hBN. (a) Optical micrograph of a bilayer graphene ake
sectioned using AFM tip as indicated. (b) Optical micrograph of the rotationally aligned double bilayer graphene separated by interlayer hBN after
the pick-up and transfer sequence. The dashed lines mark the contours of the two layers. (c) Optical micrograph of the device after metal contact
deposition. (d) Device schematic and biasing scheme. (e, f) IIL versus VIL at dierent VBG for rotationally aligned double bilayer graphene with (e)
two and (f) four layer-thick interlayer hBN dielectric; the right y-axes show the tunneling current normalized to the active area (JIL). Panels e and f
data show distinct tunneling resonance with negative dierential resistance. Panel f inset: band diagrams corresponding to the primary and secondary
tunneling resonances. (g) Primary and secondary tunneling resonances as a function of VIL and VBG for a two layer (2L), four layer (4L), and ve
layer-thick (5L) interlayer hBN dielectric. The symbols (lines) are experimental (calculated) values.

does not exhibit a bandgap opening as a function of transverse To realize the rotationally aligned double bilayer graphene
electric eld,25 and the resistivity at charge neutrality decreases heterostructure using the process ow of Figure 2, we rst
as a function of an applied transverse electric eld.26 We note section a bilayer graphene ake using either an AFM tip (Figure
here that while the layers are rotationally aligned during 5a) or by EBL followed by O2 plasma etching. After sectioning
transfer, their translation respect to each other is controlled to the bilayer, we pick up one section of the bilayer graphene,
within 1 m accuracy, a value much larger than interatomic followed by a thin hBN ake, and the other section of the
spacing. Therefore, a natural question here is why the bilayers bilayer successively. Because the two graphene bilayers are
obtained by successive, rotationally aligned transfers are Bernal obtained from the same single crystal, their crystallographic
(AB) stacked, as opposed to, for example, AA stacking. We directions remain rotationally aligned as long as the handle
surmise that if the layers are rotationally aligned they will stack substrate is only translated during the pick-up sequence. The
vertically in the most energetically favorable, Bernal stacking27 entire stack in then transferred onto an hBN ake on a SiO2/Si
conguration, as opposed to AA stacking. substrate (Figure 5b). Finally, a sequence of EBL, CHF3+O2
To further illustrate the potential of the experimental plasma etching, e-beam evaporation of metal, and lift-o is used
technique described in Figure 2 to realize functional vdW to complete the device fabrication (more details are available in
heterostructures, we apply it to a double graphene bilayer the Supporting Information Figures S3b and S4). An optical
resonant tunneling eld-eect transistor, an electronic device in micrograph of the rotationally aligned double bilayer graphene
which the rotational alignment of the crystal axes is essential for heterostructure is shown in Figure 5, panel c. Figure 5, panel d
the device functionality. Such devices exhibit a gate-tunable shows the schematic of device conguration including bias
negative dierential resistance in their interlayer tunneling condition.
characteristics thanks to energy and momentum conservation. Figures 5, panels e and f show the interlayer current (IIL)
However, because the energy band minima in graphene and versus voltage (VIL) tunneling characteristics for two double
other two-dimensional semiconductors are located away from k bilayer graphene heterostructures with two (Figure 5e) and
= 0, to preserve the translation invariance required for four (Figure 5f) layer-thick interlayer hBN dielectric,
momentum conservation in resonant tunneling, a high degree respectively, measured at various VBG values and at a
of rotational alignment of the two layers is necessary. Indeed, a temperature T = 300 K. We determine the interlayer hBN
small twist angle between the crystal axes of the two layers can thickness using a combination of AFM, interlayer capacitance
yield a large momentum separation of the band minima at the data, and a comparison of the interlayer resistance (RC) near
corners of the Brillion zones and consequently suppress zero bias normalized by area, with previously reported RC as a
resonant tunneling between the two layers.13,28 In the function of number of hBN layers.13,29 The interlayer bias is
following, we demonstrate a resonant tunneling device using applied to the top bilayer graphene, while the bottom bilayer
rotationally aligned double bilayer graphene separated by hBN graphene is grounded. We note the IIL versus VIL traces show
fabricated using the technique of Figure 2. two sets of resonance with negative dierential resistance at |
1993 DOI: 10.1021/acs.nanolett.5b05263
Nano Lett. 2016, 16, 19891995
Nano Letters Letter

VIL| 0.10.2 V and |VIL| 0.70.9 V, with a position that Details of device fabrication methods (PDF)


shifts as a function of the applied VBG. These resonances stem
from momentum-conserving tunneling between the two bilayer
graphene akes and are explained by examining the bilayer AUTHOR INFORMATION
graphene band structure, which possesses two conduction Corresponding Author
(valence) bands: a lower energy band with a minimum *E-mail: etutuc@mer.utexas.edu.
(maximum) at the CNP, and an upper band located at 0.4
Present Address
eV from the CNP.30 The primary resonances at |VIL| 0.10.2
V are associated with tunneling between the lower energy (M.Y.) Department of Physics, Columbia University, New
bands of bilayer graphene,13 while the secondary resonances at | York, NY 10027
VIL| 0.70.9 V are associated with tunneling between the Author Contributions
lower energy band of one bilayer and the upper energy band of The manuscript was written through contributions of all
the opposite bilayer graphene. authors. All authors have given approval to the nal version of
A comparison of Figure 5, panels e and f data reveals an the manuscript.
interesting and counterintuitive dependence of the tunneling Funding
resonance on interlayer dielectric thickness, where the The work at The University of Texas was supported by the
secondary resonances occur at a larger VIL in the device with NRI-SWAN center, and Intel Corp. M.Y., S.H., and B.J.L. were
a thinner interlayer dielectric. Moreover, the primary supported by NSF CAREER Award No. DMR-0953784.
resonances are broader in the device with thinner interlayer
dielectric. These observations can be explained by considering Notes
the interplay between the larger interlayer capacitance at The authors declare no competing nancial interest.
reduced dielectric thickness and the nite quantum capacitance
of bilayer graphene. Indeed, the secondary resonances occur
when the lower energy band in one bilayer is energetically
REFERENCES
(1) Dean, C. R.; Young, A. F.; Meric, I.; Lee, C.; Wang, L.;
aligned with the upper energy band of the opposite bilayer, and Sorgenfrei, S.; Watanabe, K.; Taniguchi, T.; Kim, P.; Shepard, K. L.;
therefore the electrostatic potential drop across the interlayer Hone, J. Nat. Nanotechnol. 2010, 5 (10), 722726.
dielectric (VES) is equal to the separation between the lower (2) Geim, A. K.; Grigorieva, I. V. Nature 2013, 499 (7459), 419425.
and upper subbands (0.4 eV). Reducing the interlayer dielectric (3) Wang, L.; Meric, I.; Huang, P. Y.; Gao, Q.; Gao, Y.; Tran, H.;
thickness at a xed VES leads to an increase in the charge Taniguchi, T.; Watanabe, K.; Campos, L. M.; Muller, D. A.; Guo, J.;
density in the two layers accompanied by a corresponding Kim, P.; Hone, J.; Shepard, K. L.; Dean, C. R. Science 2013, 342
(6158), 614617.
increase of the layer chemical potential. Because the applied VIL (4) Marchini, S.; Gunther, S.; Wintterlin, J. Phys. Rev. B: Condens.
is the sum of the electrostatic potential drop and the layer Matter Mater. Phys. 2007, 76 (7), 075429.
chemical potentials, a thinner dielectric requires a larger VIL to (5) Varchon, F.; Mallet, P.; Magaud, L.; Veuillen, J. Y. Phys. Rev. B
observe the secondary resonances. 2008, 77 (16), 153412.
To quantify the above argument, we write VES = VIL (B (6) Sutter, E.; Acharya, D. P.; Sadowski, J. T.; Sutter, P. Appl. Phys.
T)/e,13,28 where T and B are the top and bottom layer Lett. 2009, 94 (13), 133101.
chemical potentials, respectively, measured with respect to their (7) Xue, J. M.; Sanchez-Yamagishi, J.; Bulmash, D.; Jacquod, P.;
CNP. The primary and secondary tunneling resonances occur Deshpande, A.; Watanabe, K.; Taniguchi, T.; Jarillo-Herrero, P.; Leroy,
when VES = 0 V and VES = 0.4 V, respectively. By using a B. J. Nat. Mater. 2011, 10 (4), 282285.
band structure model for double layers, and experimentally (8) Ponomarenko, L. A.; Gorbachev, R. V.; Yu, G. L.; Elias, D. C.;
Jalil, R.; Patel, A. A.; Mishchenko, A.; Mayorov, A. S.; Woods, C. R.;
measured chemical potential layer density dependence13,28 in Wallbank, J. R.; Mucha-Kruczynski, M.; Piot, B. A.; Potemski, M.;
Figure 5, panel g, we show the calculated (lines) and Grigorieva, I. V.; Novoselov, K. S.; Guinea, F.; Falko, V. I.; Geim, A.
experimental (symbols) VIL values for the primary and K. Nature 2013, 497 (7451), 594597.
secondary resonances as a function of VBG for dierent (9) Dean, C. R.; Wang, L.; Maher, P.; Forsythe, C.; Ghahari, F.; Gao,
interlayer dielectric thicknesses. A back-gate capacitance of Y.; Katoch, J.; Ishigami, M.; Moon, P.; Koshino, M.; Taniguchi, T.;
10.5 nF/cm2 and interlayer dielectric capacitance values of 3.10 Watanabe, K.; Shepard, K. L.; Hone, J.; Kim, P. Nature 2013, 497
F/cm2, 1.80 F/cm2, and 1.55 F/cm2 from the device with (7451), 598602.
two, four, and ve layer-thick interlayer hBN were used in the (10) Hunt, B.; Sanchez-Yamagishi, J. D.; Young, A. F.; Yankowitz,
calculations. M.; LeRoy, B. J.; Watanabe, K.; Taniguchi, T.; Moon, P.; Koshino, M.;
In summary, we describe a novel technique to realize Jarillo-Herrero, P.; Ashoori, R. C. Science 2013, 340 (6139), 1427
1430.
rotationally aligned vdW heterostructures and illustrate it by
(11) Britnell, L.; Gorbachev, R. V.; Geim, A. K.; Ponomarenko, L. A.;
creating a Bernal stacked bilayer graphene by successive Mishchenko, A.; Greenaway, M. T.; Fromhold, T. M.; Novoselov, K.
transfers of two monolayers, and we also demonstrate gate S.; Eaves, L. Nat. Commun. 2013, 4, 1794.
tunable resonant tunneling eld-eect transistors in double (12) Mishchenko, A.; Tu, J. S.; Cao, Y.; Gorbachev, R. V.; Wallbank,
bilayer graphene. The technique has the potential to be scaled J. R.; Greenaway, M. T.; Morozov, V. E.; Morozov, S. V.; Zhu, M. J.;
up using large area graphene single crystals,31 or TMDs, and Wong, S. L.; Withers, F.; Woods, C. R.; Kim, Y. J.; Watanabe, K.;
can enable various functional vdW heterostructures. Taniguchi, T.; Vdovin, E. E.; Makarovsky, O.; Fromhold, T. M.; Falko,

V. I.; Geim, A. K.; Eaves, L.; Novoselov, K. S. Nat. Nanotechnol. 2014,


ASSOCIATED CONTENT 9 (10), 808813.
(13) Fallahazad, B.; Lee, K.; Kang, S.; Xue, J. M.; Larentis, S.; Corbet,
*
S Supporting Information
C.; Kim, K.; Movva, H. C. P.; Taniguchi, T.; Watanabe, K.; Register, L.
The Supporting Information is available free of charge on the F.; Banerjee, S. K.; Tutuc, E. Nano Lett. 2015, 15 (1), 428433.
ACS Publications website at DOI: 10.1021/acs.nano- (14) You, Y. M.; Ni, Z. H.; Yu, T.; Shen, Z. X. Appl. Phys. Lett. 2008,
lett.5b05263. 93 (16), 163112.

1994 DOI: 10.1021/acs.nanolett.5b05263


Nano Lett. 2016, 16, 19891995
Nano Letters Letter

(15) Ferrari, A. C.; Meyer, J. C.; Scardaci, V.; Casiraghi, C.; Lazzeri,
M.; Mauri, F.; Piscanec, S.; Jiang, D.; Novoselov, K. S.; Roth, S.; Geim,
A. K. Phys. Rev. Lett. 2006, 97 (18), 187401.
(16) Kim, K.; Coh, S.; Tan, L. Z.; Regan, W.; Yuk, J. M.; Chatterjee,
E.; Crommie, M. F.; Cohen, M. L.; Louie, S. G.; Zettl, A. Phys. Rev.
Lett. 2012, 108 (24), 246103.
(17) Stolyarova, E.; Rim, K. T.; Ryu, S. M.; Maultzsch, J.; Kim, P.;
Brus, L. E.; Heinz, T. F.; Hybertsen, M. S.; Flynn, G. W. Proc. Natl.
Acad. Sci. U. S. A. 2007, 104 (22), 92099212.
(18) Zhang, Y. B.; Brar, V. W.; Wang, F.; Girit, C.; Yayon, Y.;
Panlasigui, M.; Zettl, A.; Crommie, M. F. Nat. Phys. 2008, 4 (8), 627
630.
(19) Deshpande, A.; Bao, W.; Zhao, Z.; Lau, C. N.; LeRoy, B. J. Appl.
Phys. Lett. 2009, 95 (24), 243502.
(20) Rutter, G. M.; Jung, S. Y.; Klimov, N. N.; Newell, D. B.;
Zhitenev, N. B.; Stroscio, J. A. Nat. Phys. 2011, 7 (8), 649655.
(21) Yankowitz, M.; Wang, J. I. J.; Li, S. C.; Birdwell, A. G.; Chen, Y.
A.; Watanabe, K.; Taniguchi, T.; Quek, S. Y.; Jarillo-Herrero, P.;
LeRoy, B. J. APL Mater. 2014, 2 (9), 092503.
(22) Zou, K.; Zhu, J. Phys. Rev. B: Condens. Matter Mater. Phys. 2010,
82 (8), 081407.
(23) Taychatanapat, T.; Jarillo-Herrero, P. Phys. Rev. Lett. 2010, 105
(16), 166601.
(24) Lee, K.; Fallahazad, B.; Min, H.; Tutuc, E. IEEE Trans. Electron
Devices 2013, 60 (1), 103108.
(25) Lopes dos Santos, J. M. B.; Peres, N. M. R.; Castro Neto, A. H.
Phys. Rev. Lett. 2007, 99 (25), 256802.
(26) Sanchez-Yamagishi, J. D.; Taychatanapat, T.; Watanabe, K.;
Taniguchi, T.; Yacoby, A.; Jarillo-Herrero, P. Phys. Rev. Lett. 2012, 108
(7), 076601.
(27) Kolmogorov, A. N.; Crespi, V. H. Phys. Rev. B: Condens. Matter
Mater. Phys. 2005, 71 (23), 235415.
(28) Kang, S.; Fallahazad, B.; Lee, K.; Movva, H.; Kim, K.; Corbet, C.
M.; Taniguchi, T.; Watanabe, K.; Colombo, L.; Register, L. F.; Tutuc,
E.; Banerjee, S. K. IEEE Electron Device Lett. 2015, 36 (4), 405407.
(29) Britnell, L.; Gorbachev, R. V.; Jalil, R.; Belle, B. D.; Schedin, F.;
Katsnelson, M. I.; Eaves, L.; Morozov, S. V.; Mayorov, A. S.; Peres, N.
M. R.; Castro Neto, A. H.; Leist, J.; Geim, A. K.; Ponomarenko, L. A.;
Novoselov, K. S. Nano Lett. 2012, 12 (3), 17071710.
(30) McCann, E.; Abergel, D. S. L.; Falko, V. I. Solid State Commun.
2007, 143 (12), 110115.
(31) Hao, Y. F.; Bharathi, M. S.; Wang, L.; Liu, Y. Y.; Chen, H.; Nie,
S.; Wang, X. H.; Chou, H.; Tan, C.; Fallahazad, B.; Ramanarayan, H.;
Magnuson, C. W.; Tutuc, E.; Yakobson, B. I.; McCarty, K. F.; Zhang,
Y. W.; Kim, P.; Hone, J.; Colombo, L.; Ruoff, R. S. Science 2013, 342
(6159), 720723.

NOTE ADDED AFTER ASAP PUBLICATION


This paper was published on the Web on February 15, 2016,
with the second authors last name misspelled. The corrected
version was reposted on February 16, 2016.

1995 DOI: 10.1021/acs.nanolett.5b05263


Nano Lett. 2016, 16, 19891995

Potrebbero piacerti anche