We use micro-Raman spectroscopy to study strain in free-standing graphene monolayers anchored to SiN holes of non-circular geometry. We show that a uniform differential pressure load yields measurable deviations from hydrostatic strain, conventionally observed in radially symmetric microbubbles. A pressure load of 1 bar yields a top hydrostatic strain of ≈ 0.7% and a G± splitting of 10 cm−1 in graphene clamped to elliptical boundaries with axes 40 and 20 μm, in good agreement with the calculated anisotropy Δε ≈ 0.6% and consistently with recent reports on Grüneisen parameters. The implementation of arbitrary strain configurations by designing suitable boundary clamping conditions is discussed.

Graphene displays a range of remarkable properties that have catalyzed—since its discovery in 20041—an impressive interest in the scientific community.2 Its unique electronic behavior stems from the hexagonal honeycomb structure of the carbon lattice, which forces low-energy conducting electrons to assume linear dispersions that mimic massless relativistic fermions.3 In addition, graphene displays an unusual mechanical strength and strains up to 10% can be applied to it without damaging appreciably its structure.4 This factor, combined with its intrinsic two-dimensional nature, opens unique perspectives for the investigation of strain engineering5–7 and for the development of novel device concepts.8 In fact, it has been predicted,7,9 and in part experimentally demonstrated,10,11 that mechanical deformations in graphene can be used to tailor its electron properties. As a particularly inspiring possibility, it is known that a suitable deformation of the honeycomb lattice can be equivalent to the application of a pseudomagnetic field.6,12

Achieving a controlled strain profile in graphene poses non-trivial technical challenges and various alternative approaches have been explored during recent years. Hydrostatic configurations were obtained and studied using circular holes and a uniform differential pressure load,13–15 and the impact of strain was studied by micro-Raman spectroscopy.16–18 In this device architecture, local strain was also induced or measured taking advantage of scanning probe techniques.19,20 Differently, uniform biaxial strain configurations could be induced by taking advantage of piezoelectric effects in the substrate21 and by using pressure transmitting media.22 Concerning uniaxial strain, various studies have demonstrated the possibility to anisotropically deform graphene deposited on non-flat substrates,23 on polydimethylsiloxane (PDMS)24 or on similar stretchable substrates.25,26 An alternative promising approach consists in anchoring graphene layers to micro-electromechanical actuators.27 More elaborated strain profiles, in particular those giving rise to a pseudomagnetic field, have been hard to demonstrate so far. Interesting experimental evidence has been put forward in the context of random nanobubbles10 and fascinating results have been obtained in deformed artificial honeycomb structures mimicking the behavior of graphene.11 In practice, the achievement of custom strain profiles has generally proved to be rather elusive.

In the present work, we demonstrate that markedly non-isotropic strain profiles can be obtained in free-standing graphene membranes that are clamped on an edge that is not radially symmetric and are subject to a vertical uniform load using a pressure difference between the two opposite faces of the graphene flake. In particular, we show that loaded elliptical membranes display Raman features that demonstrate the presence of an anisotropic component in the induced strain profile, in good agreement with what is expected with the studied geometry. The possibility to achieve custom strain profiles by choosing a suitable graphene clamping geometry is discussed.

Figure 1 shows the device architecture and setup adopted for this work. Free-standing graphene areas of various shapes and dimensions were obtained using micropatterned SiN membranes as the mechanical support for the graphene layer. Starting from a Si wafer doubly coated in “pre-stressed” 300nm of Si3N4, a combination of dry and wet etching protocols (see the supplementary material for further details) was adopted to obtain suspended 500×500μm2 Si3N4 membranes with through holes of various geometries. In the present study, we investigated a set of elliptical holes with various major (a) and minor (b) axes: a×b=5μm×10μm, 10μm×20μm, and 20μm×40μm. Circular holes were also investigated, as reported in the literature.17,28 Large-scale monocrystalline graphene flakes used in the present work were obtained by CVD growth on Cu29 and transferred on the Si3N4/Si chips using a standard “bubbling transfer” technique.30 As sketched in Fig. 1(a), a differential pressure ΔP is applied orthogonally to the free-standing graphene region thanks to a 2mm-thick polydimethylsiloxane (PDMS) coupling layer placed on top of a modified microscope slide. As a result, the top side of the free-standing graphene region is subject to ambient pressure (conventionally P0=1bar) while the bottom space can be partially or completely evacuated with a scroll pump. Static vacuum tests were performed to verify the stability of ΔP, which was found to decay over a time scale of various hours; this ensures pressure values measured by the gauge are meaningful. Maps were always performed under active pumping conditions, at ΔP=1bar.

FIG. 1.

Straining graphene with a differential pressure load. (a) Sketch of the device architecture. Monolayer CVD graphene is transferred on a patterned SiN membrane. The bottom of the chip is coupled to a vacuum chamber using a polydimethylsiloxane layer, providing a sealing between the two. Deformed graphene is investigated by micro-Raman spectroscopy as a function of the applied differential pressure ΔP with respect to the reference pressure P0. Right picture: optical image of one of the CVD monocrystals deposited on the patterned Si3N4. (b) Raman spectrum measured at ΔP=0 (blue curve) and ΔP=0.9bar (red curve) for an elliptical SiN membrane with axes 20μm and 10μm.

FIG. 1.

Straining graphene with a differential pressure load. (a) Sketch of the device architecture. Monolayer CVD graphene is transferred on a patterned SiN membrane. The bottom of the chip is coupled to a vacuum chamber using a polydimethylsiloxane layer, providing a sealing between the two. Deformed graphene is investigated by micro-Raman spectroscopy as a function of the applied differential pressure ΔP with respect to the reference pressure P0. Right picture: optical image of one of the CVD monocrystals deposited on the patterned Si3N4. (b) Raman spectrum measured at ΔP=0 (blue curve) and ΔP=0.9bar (red curve) for an elliptical SiN membrane with axes 20μm and 10μm.

Close modal

Local graphene deformation is investigated by micro-Raman spectroscopy, using an inVia confocal system by Renishaw equipped with a polarized λ=532nm laser source. Raman signal was collected through a 50× objective with N.A. = 0.75 and analyzed by 1800 grooves/mm grating. Light collection was not polarized, except for a minor polarization selectivity intrinsic in our single-grating monochromator. Raman maps were collected using a laser power of 1mW to minimize the impact of local heating. Laser polarization was always set along the ellipse minor axis. In Fig. 1(b) we report the measured Raman spectra collected at the center of a 10μm×20μm elliptical graphene region, for ΔP=0 (blue curve) and ΔP=0.9bar (red curve); as expected, the G and 2D Raman peaks are significantly red shifted by strain in the suspended graphene region. Importantly, the modification of the Raman spectra was always found to be completely reversible upon removal of the pressure load. In addition, as visible from the maps of Fig. 2, no significant Raman shift was observed in the regions where graphene is supported by the SiN, even in the proximity of the hole. This proves that no measurable adjustment or sliding of graphene occurs during our experiments.

FIG. 2.

Impact of strain on the 2D peak. (a) Map of the position of the 2D peak for ΔP=0; the peak position is mostly uniform with a slight red shift over the suspended region (see discussion in the main text). (b) The application of a differential pressure ΔP=1bar leads to a dome-shaped shift, which is maximal at the center of the membrane. As further argued based on data presented in Fig. 3, this effect can be explained as the consequence of a hydrostatic strain ε¯ in the elliptical hole. (c) Evolution of the Raman shift as a function of the parameter η=(ΔP/P0)2/3, which is proportional to ε¯. A linear regression of the observed Raman shifts (excluding the value at η=0) and a comparison with numerical estimates of the strain yield a Grüneisen parameter consistent with most recent results reported in the literature. (d) Sketch of the D mode in graphene, whose second order causes the 2D Raman resonance.

FIG. 2.

Impact of strain on the 2D peak. (a) Map of the position of the 2D peak for ΔP=0; the peak position is mostly uniform with a slight red shift over the suspended region (see discussion in the main text). (b) The application of a differential pressure ΔP=1bar leads to a dome-shaped shift, which is maximal at the center of the membrane. As further argued based on data presented in Fig. 3, this effect can be explained as the consequence of a hydrostatic strain ε¯ in the elliptical hole. (c) Evolution of the Raman shift as a function of the parameter η=(ΔP/P0)2/3, which is proportional to ε¯. A linear regression of the observed Raman shifts (excluding the value at η=0) and a comparison with numerical estimates of the strain yield a Grüneisen parameter consistent with most recent results reported in the literature. (d) Sketch of the D mode in graphene, whose second order causes the 2D Raman resonance.

Close modal

The most evident impact of deformation upon ΔP is visible in Fig. 2, where we compare the map of the 2D peak Raman shift ω2D for ΔP=0 (Fig. 2(a)) and ΔP=1bar (Fig. 2(b)). For symmetry reason, its center frequency is sensitive to the hydrostatic component of the strain tensor εij, which we name ε¯=(εxx+εyy)/2. Experimentally, the 2D peak displays a maximal red shift at the center of the suspended region, similar to what was reported for inflated circular graphene membranes.17 The quantitative evolution of the shift versus ΔP is highlighted in Fig. 2(c) where we report ω2D measured at the maximal shift region at the center of the ellipse. The phononic mode giving rise to the higher order 2D peak is sketched in Fig. 2(d). Raman shifts at various pressure loads are compared with the adimensional parameter η=(ΔP/P0)2/3, where P0=1bar; all the components of the strain tensor εij are in fact expected to scale linearly with η, as highlighted for specific cases in the recent literature.16 The validity of such a scaling law is also formally demonstrated, for a general clamping geometry, in the supplementary material. The overall dependence of ω2D as a function of ΔP and of the position on the ellipse is found to be largely consistent with reports on the simpler case of radially symmetric graphene clamping and with the most recent estimates of the Grüneisen parameters.16,25,31 We also note that, consistently with reported data,16 the value of ω2D for ΔP0 is found to display a further surprising red shift. This effect was attributed to uncertainties in the exact determination of ΔP. We believe that an additional reason for the shift is possibly related to graphene adhesion on the vertical sidewalls of the SiN hole, which is known to occur in this kind of graphene drums16 and could be relevant in the low-ΔP regime. The verification of this hypothesis will likely require a combined Raman and atomic force microscopy study at low pressure loads, which goes beyond the scope of the present work. Further details regarding the numerical calculation of the εij tensor as a function of ΔP and scaling rules are reported in the supplementary material.

While hydrostatic deformations explain well the coarse evolution of the Raman spectra, the strain profiles in our elliptically clamped graphene membranes are expected to display a marked deviation from a uniform strain configuration and a larger strain can be expected along the shorter axis of the ellipse. More in general, the anisotropic component of the strain field can be expressed through the invariant

Δε=(εxxεyy)2+4εxy2
(1)

corresponding to the difference between two eigenvalues of the strain tensor ε=ε¯±Δε/2. It is well known31 that strain anisotropy, when sufficiently large, can be detected in Raman spectroscopy as a splitting of the degenerate phononic modes G±. In Fig. 3 we report a detailed study of the Raman spectrum of the G peak region as a function of ΔP. Data reported in Fig. 3 refer to the largest explored Si3N4 elliptical 20×40μm2 hole; larger suspended areas in fact correspond—for a given value of ΔP—to a larger anisotropic strain Δε.

FIG. 3.

Strain-induced shift and splitting of the G peak. (a) Map of the Raman shift ωG obtained by fitting the G peak with a single Lorentzian curve. (b) Simulated average strain map ε¯=(εxx+εyy)/2 at ΔP=1bar. (c) Map of the peak broadening ΓG obtained by fitting the G peak with a single Lorentzian curve. (d) Simulated strain anisotropy map Δε=(εxxεyy)2+4εxy2 at ΔP=1bar. (e) Multipeak fit of the G peak at ΔP=1bar measured at the center of the elliptical hole. The resulting positions of the G+ (red curve and markers) and G (green curve and markers) peak versus ΔP are reported in the inset along with a weighted linear fit. (f) Sketch of the G+ and G modes in the presence of an arbitrary anisotropic strain.

FIG. 3.

Strain-induced shift and splitting of the G peak. (a) Map of the Raman shift ωG obtained by fitting the G peak with a single Lorentzian curve. (b) Simulated average strain map ε¯=(εxx+εyy)/2 at ΔP=1bar. (c) Map of the peak broadening ΓG obtained by fitting the G peak with a single Lorentzian curve. (d) Simulated strain anisotropy map Δε=(εxxεyy)2+4εxy2 at ΔP=1bar. (e) Multipeak fit of the G peak at ΔP=1bar measured at the center of the elliptical hole. The resulting positions of the G+ (red curve and markers) and G (green curve and markers) peak versus ΔP are reported in the inset along with a weighted linear fit. (f) Sketch of the G+ and G modes in the presence of an arbitrary anisotropic strain.

Close modal

A first rough analysis was performed by fitting the Raman data with a single Lorentzian peak. As visible in Fig. 3(a), the resulting map of the Raman shift ωG at ΔP=1bar is found to be consistent with the ω2D map reported in Fig. 2(a). For comparison, we report in Fig. 3(b) the map of ε¯ calculated for the same pressure load. A top hydrostatic strain of 0.68% is expected, in agreement with the observed red shift of the G peak and known values of the corresponding Grüneisen parameter (see the supplementary material for further details). As argued in the following, on the other hand, hydrostatic strain is not sufficient to satisfactorily describe the evolution of the G peak as a function of the pressure load. Indeed, a non-trivial broadening is visible in Fig. 3(c), where we report the FWHM ΓG resulting from the same fitting procedure. The observed broadening displays a peculiar “saddle point” spatial evolution, with large ΓG values close to the central part of the ellipse, while a significantly smaller effect is observed at the top and bottom apexes. Remarkably, a very similar pattern is visible in Fig. 3(d), displaying the calculated Δε for the at ΔP=1bar. This suggests that, beyond mechanisms highlighted in recent works,16 broadening in our experiment is also in part connected to strain anisotropy.

A more detailed investigation of the broadening mechanism was performed through the analysis of the G peak measured at the center of the elliptical hole, which represents a good trade-off between the expected value of Δε and the minimization of the impact of the borders of the Si3N4 hole. In this position, numerical estimates indicate that a top anisotropic strain component Δε = 0.64% can be expected. The Raman spectrum in the G peak region for ΔP=1bar is reported in Fig. 3(e); the peak displays a lineshape which is clearly consistent with the superposition of two nearby Lorentzian peaks, which we interpret as corresponding to the G+ and G modes in uniaxially strained graphene (see Fig. 3(f)). A similar analysis (see the supplementary material for further information concerning the fitting procedure) was performed for various values of ΔP and the resulting peak positions are reported in the inset to Fig. 3(e). Two divergent peaks are obtained with a top splitting of about 10 cm−1, which is in very good agreement31 with what is expected for an anisotropy Δε=0.64%. A weighted linear regression of the two peak positions ωG± yields two remarkably linear trends in η that converge almost exactly for ΔP=0, as expected. It is important to note that, since light collection in our experiment is not polarized, both G+ and G are expected to be visible. In fact, only a minor amplitude difference is observed, as a likely consequence of the residual polarization sensitivity of our single-grating monochromator. Finally, we note that no evidence of a clear splitting of the 2D peak could be detected in the current experiments, differently from what was reported in other works.32 Since the 2D splitting strongly depends on the strain direction, this behavior could be linked to the orientation of the studied flakes with respect to the ellipse axes.

Our work demonstrates that non-trivial strain profiles can be obtained in Si3N4 holes with an elliptical shape and that a sizable anisotropic component in the graphene strain can be obtained. Similarly, we expect that more in general clamping geometries can be used to design even more advanced non-uniform and non-isotropic strain profiles, taking advantage of a relatively robust implementation with no free graphene edges. The relevant case of triaxial strain, which however also requires a correct crystalline orientation to give rise to a pseudo magnetic field, is reported in the supplementary material. In view of the possibility to impact the electronic states via the engineering of custom strain profiles, the observed G± splitting phenomenology has also been compared with a first-principle calculation on an atomistic model system mimicking the experimental setup. To this end, we have simulated an unstrained suspended graphene layer with an elliptic-shape depression (see the inset of Fig. 4). To reproduce the experimental configuration, we have fixed the position of the carbon atoms external to the ellipse to a “zero” height, as it happens for graphene on Si3N4 substrate, and the effect of the vertical load has been reproduced by fixing the two carbon atoms at the center of the ellipse (blue dots in the inset) at a lower vertical position and leaving all the other carbon atoms in the ellipse free to relax in order to reach the minimum energy structure. The simulation has been performed on a 7.4Å×12.8Å ellipse containing a total of 22 carbon atoms. The Raman spectra of the system have been calculated by means of density functional perturbation theory33 as implemented in QUANTUM-ESPRESSO code,34 with local density approximation and norm-conserving pseudopotential for the carbon atoms.35 We used a plane wave expansion up to 80Ry cutoff and 4×4×1 Monkhorst-Pack mesh36 for the sampling of Brillouin zone. The ellipse depression was created in the central part of a 5×5 graphene supercell with the minimum energy configuration lattice parameter a0=12.24Å.

FIG. 4.

Position of the G± peak as function of the major axis strain. The slope of the Raman shifts is ωG/εa35cm1/% for G and ωG+/εa23cm1/% for G+. The inset shows the simulation cell (green) containing the ellipse. The lowest fixed carbon atoms are indicated in blue.

FIG. 4.

Position of the G± peak as function of the major axis strain. The slope of the Raman shifts is ωG/εa35cm1/% for G and ωG+/εa23cm1/% for G+. The inset shows the simulation cell (green) containing the ellipse. The lowest fixed carbon atoms are indicated in blue.

Close modal

In the absence of strain, the frequency of the degenerate G phononic mode is ωG=1607.5 cm−1. For ΔP0, the value of the effective strain along the ellipse axes has been estimated from the depth of the two central carbon atoms, δ and the length of the principal axes of the ellipse, a and b; e.g., for the major axis we have εa(2La)/a where L(a/2)2+δ2 is the profile length for the depression along the a direction; the same holds for the minor axis. The results of the calculation are shown in Fig. 4; the asymmetry of the strain due to the elliptical geometry of the depression is responsible of an averaged splitting of the G peak mode that is found to be of the same order of magnitude as the one experimentally induced by uniaxial strain on the graphene layer.25 

In conclusion, we have provided evidence of an incipient splitting of the G mode in free-standing graphene regions clamped to elliptical holes in Si3N4 and subject to a uniform differential pressure load. Our results indicate a promising route to induce custom strain profiles, which can be controlled by the applied pressure load and designed according to the chosen geometry of the supporting Si3N4 frame. We also highlight that our experiment has been performed using large scale CVD monocrystalline graphene flakes, thus providing a route for scalable strain-engineered graphene devices. Finally, we would like to stress that present results have been obtained using a vacuum chamber to induce a maximal load ΔP=1bar. A recent report16 demonstrates that using pressurized gas as a load it is possible to reach ΔP=14bar, potentially leading to significantly increased achievable strain magnitude (about a factor six larger strain can be expected) and/or to less stringent limits on the minimal area of the Si3N4 holes.

See the supplementary material for more details about sample fabrication, numerical calculations, demonstration of the scaling law, and Raman measurements. An example of a hole geometry which induces a triaxial strain profile with the related pseudo-magnetic field is also given.

This work was supported by the EC under the Graphene Flagship Program (Contract No. CNECT-ICT-604391) and by the ERC advanced grant SoulMan (No. G.A. 321122). The authors acknowledge the “IT center” of the University of Pisa for the computational support and the allocation of computer resources from the CINECA, ISCRA C Project Nos. HP10C6H6O1 and HP10CAI9PV. F.C., A.P., and S.R. thank S. Gröblacher and R. Norte for their help in the initial development of this experiment. S.R. acknowledges the support of the CNR through the bilateral CNR-RFBR projects. F.C. acknowledges support from Fondazione Silvio Tronchetti Provera.

1.
K.
Novoselov
,
D.
Jiang
,
F.
Schedin
,
T.
Booth
,
V.
Khotkevich
,
S.
Morozov
, and
A.
Geim
, “
Two-dimensional atomic crystals
,”
Proc. Natl. Acad. Sci. U. S. A.
102
,
10451
10453
(
2005
).
2.
A. K.
Geim
and
K. S.
Novoselov
, “
The rise of graphene
,”
Nat. Mater.
6
,
183
191
(
2007
).
3.
A. C.
Neto
,
F.
Guinea
,
N.
Peres
,
K. S.
Novoselov
, and
A. K.
Geim
, “
The electronic properties of graphene
,”
Rev. Mod. Phys.
81
,
109
(
2009
).
4.
C.
Lee
,
X.
Wei
,
J. W.
Kysar
, and
J.
Hone
, “
Measurement of the elastic properties and intrinsic strength of monolayer graphene
,”
Science
321
,
385
388
(
2008
).
5.
V. M.
Pereira
and
A. C.
Neto
, “
Strain engineering of graphene electronic structure
,”
Phys. Rev. Lett.
103
,
046801
(
2009
).
6.
T.
Low
and
F.
Guinea
, “
Strain-induced pseudomagnetic field for novel graphene electronics
,”
Nano Lett.
10
,
3551
3554
(
2010
).
7.
F.
Guinea
, “
Strain engineering in graphene
,”
Solid State Commun.
152
,
1437
1441
(
2012
).
8.
Z.
Qi
,
D.
Bahamon
,
V. M.
Pereira
,
H. S.
Park
,
D.
Campbell
, and
A. C.
Neto
, “
Resonant tunneling in graphene pseudomagnetic quantum dots
,”
Nano Lett.
13
,
2692
2697
(
2013
).
9.
A. K.
Geim
, “
Graphene: Status and prospects
,”
Science
324
,
1530
1534
(
2009
).
10.
N.
Levy
,
S.
Burke
,
K.
Meaker
,
M.
Panlasigui
,
A.
Zettl
,
F.
Guinea
,
A. C.
Neto
, and
M.
Crommie
, “
Strain-induced pseudo–magnetic fields greater than 300 tesla in graphene nanobubbles
,”
Science
329
,
544
547
(
2010
).
11.
M.
Polini
,
F.
Guinea
,
M.
Lewenstein
,
H. C.
Manoharan
, and
V.
Pellegrini
, “
Artificial honeycomb lattices for electrons, atoms and photons
,”
Nat. Nanotechnol.
8
,
625
633
(
2013
).
12.
F.
Guinea
,
M.
Katsnelson
, and
A.
Geim
, “
Energy gaps and a zero-field quantum hall effect in graphene by strain engineering
,”
Nat. Phys.
6
,
30
33
(
2010
).
13.
J. S.
Bunch
,
S. S.
Verbridge
,
J. S.
Alden
,
A. M.
Van Der Zande
,
J. M.
Parpia
,
H. G.
Craighead
, and
P. L.
McEuen
, “
Impermeable atomic membranes from graphene sheets
,”
Nano Lett.
8
,
2458
2462
(
2008
).
14.
A. L.
Kitt
,
Z.
Qi
,
S.
Remi
,
H. S.
Park
,
A. K.
Swan
, and
B. B.
Goldberg
, “
How graphene slides: Measurement and theory of strain-dependent frictional forces between graphene and SiO2
,”
Nano Lett.
13
,
2605
2610
(
2013
).
15.
S. P.
Koenig
,
N. G.
Boddeti
,
M. L.
Dunn
, and
J. S.
Bunch
, “
Ultrastrong adhesion of graphene membranes
,”
Nat. Nanotechnol.
6
,
543
546
(
2011
).
16.
Y.
Shin
,
M.
Lozada-Hidalgo
,
J. L.
Sambricio
,
I. V.
Grigorieva
,
A. K.
Geim
, and
C.
Casiraghi
, “
Raman spectroscopy of highly pressurized graphene membranes
,”
Appl. Phys. Lett.
108
,
221907
(
2016
).
17.
J.
Zabel
,
R. R.
Nair
,
A.
Ott
,
T.
Georgiou
,
A. K.
Geim
,
K. S.
Novoselov
, and
C.
Casiraghi
, “
Raman spectroscopy of graphene and bilayer under biaxial strain: Bubbles and balloons
,”
Nano Lett.
12
,
617
621
(
2012
).
18.
J.-U.
Lee
,
D.
Yoon
, and
H.
Cheong
, “
Estimation of youngs modulus of graphene by raman spectroscopy
,”
Nano Lett.
12
,
4444
4448
(
2012
).
19.
S.
Zhu
,
Y.
Huang
,
N. N.
Klimov
,
D. B.
Newell
,
N. B.
Zhitenev
,
J. A.
Stroscio
,
S. D.
Solares
, and
T.
Li
, “
Pseudomagnetic fields in a locally strained graphene drumhead
,”
Phys. Rev. B
90
,
075426
(
2014
).
20.
R.
Beams
,
L. G.
Cançado
,
A.
Jorio
,
A. N.
Vamivakas
, and
L.
Novotny
, “
Tip-enhanced raman mapping of local strain in graphene
,”
Nanotechnology
26
,
175702
(
2015
).
21.
W.
Jie
,
Y. Y.
Hui
,
Y.
Zhang
,
S. P.
Lau
, and
J.
Hao
, “
Effects of controllable biaxial strain on the raman spectra of monolayer graphene prepared by chemical vapor deposition
,”
Appl. Phys. Lett.
102
,
223112
(
2013
).
22.
K.
Filintoglou
,
N.
Papadopoulos
,
J.
Arvanitidis
,
D.
Christofilos
,
O.
Frank
,
M.
Kalbac
,
J.
Parthenios
,
G.
Kalosakas
,
C.
Galiotis
, and
K.
Papagelis
, “
Raman spectroscopy of graphene at high pressure: Effects of the substrate and the pressure transmitting media
,”
Phys. Rev. B
88
,
045418
(
2013
).
23.
C.
Metzger
,
S.
Rémi
,
M.
Liu
,
S. V.
Kusminskiy
,
A. H.
Castro Neto
,
A. K.
Swan
, and
B. B.
Goldberg
, “
Biaxial strain in graphene adhered to shallow depressions
,”
Nano Lett.
10
,
6
10
(
2009
).
24.
M.
Huang
,
H.
Yan
,
C.
Chen
,
D.
Song
,
T. F.
Heinz
, and
J.
Hone
, “
Phonon softening and crystallographic orientation of strained graphene studied by raman spectroscopy
,”
Proc. Natl. Acad. Sci. U. S. A.
106
,
7304
7308
(
2009
).
25.
T.
Mohiuddin
,
A.
Lombardo
,
R.
Nair
,
A.
Bonetti
,
G.
Savini
,
R.
Jalil
,
N.
Bonini
,
D.
Basko
,
C.
Galiotis
,
N.
Marzari
 et al, “
Uniaxial strain in graphene by raman spectroscopy: G peak splitting, Grüneisen parameters, and sample orientation
,”
Phys. Rev. B
79
,
205433
(
2009
).
26.
Z. H.
Ni
,
T.
Yu
,
Y. H.
Lu
,
Y. Y.
Wang
,
Y. P.
Feng
, and
Z. X.
Shen
, “
Uniaxial strain on graphene: Raman spectroscopy study and band-gap opening
,”
ACS Nano
2
,
2301
2305
(
2008
).
27.
H. H.
Perez Garza
,
E. W.
Kievit
,
G. F.
Schneider
, and
U.
Staufer
, “
Controlled, reversible, and nondestructive generation of uniaxial extreme strains (>10%) in graphene
,”
Nano Lett.
14
,
4107
4113
(
2014
).
28.
C.
Si
,
Z.
Sun
, and
F.
Liu
, “
Strain engineering of graphene: A review
,”
Nanoscale
8
,
3207
(
2016
).
29.
V.
Miseikis
,
D.
Convertino
,
N.
Mishra
,
M.
Gemmi
,
T.
Mashoff
,
S.
Heun
,
N.
Haghighian
,
F.
Bisio
,
M.
Canepa
,
V.
Piazza
 et al, “
Rapid cvd growth of millimetre-sized single crystal graphene using a cold-wall reactor
,”
2D Mater.
2
,
014006
(
2015
).
30.
L.
Gao
,
W.
Ren
,
H.
Xu
,
L.
Jin
,
Z.
Wang
,
T.
Ma
,
L.-P.
Ma
,
Z.
Zhang
,
Q.
Fu
,
L.-M.
Peng
 et al, “
Repeated growth and bubbling transfer of graphene with millimetre-size single-crystal grains using platinum
,”
Nat. Commun.
3
,
699
(
2012
).
31.
Y.
Cheng
,
Z.
Zhu
,
G.
Huang
, and
U.
Schwingenschlögl
, “
Grüneisen parameter of the G mode of strained monolayer graphene
,”
Phys. Rev. B
83
,
115449
(
2011
).
32.
M.
Mohr
,
J.
Maultzsch
, and
C.
Thomsen
, “
Splitting of the raman 2 d band of graphene subjected to strain
,”
Phys. Rev. B
82
,
201409
(
2010
).
33.
S.
Baroni
,
S.
De Gironcoli
,
A.
Dal Corso
, and
P.
Giannozzi
, “
Phonons and related crystal properties from density-functional perturbation theory
,”
Rev. Mod. Phys.
73
,
515
(
2001
).
34.
P.
Giannozzi
,
S.
Baroni
,
N.
Bonini
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
G. L.
Chiarotti
,
M.
Cococcioni
,
I.
Dabo
 et al, “
Quantum espresso: A modular and open-source software project for quantum simulations of materials
,”
J. Phys.: Condens. Matter
21
,
395502
(
2009
).
35.
N.
Troullier
and
J. L.
Martins
, “
Efficient pseudopotentials for plane-wave calculations
,”
Phys. Rev. B
43
,
1993
(
1991
).
36.
H. J.
Monkhorst
and
J. D.
Pack
, “
Special points for brillouin-zone integrations
,”
Phys. Rev. B
13
,
5188
(
1976
).

Supplementary Material