You are on page 1of 7

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/257238998

Few layer graphene based superlattices as efficient thermal insulators

Article in Applied Physics Letters · October 2013


DOI: 10.1063/1.4824013

CITATIONS READS
30 893

3 authors, including:

Yuxiang Ni Sebastian Volz


Southwest Jiaotong University French National Centre for Scientific Research
136 PUBLICATIONS 1,942 CITATIONS 410 PUBLICATIONS 10,242 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Yuxiang Ni on 28 May 2014.

The user has requested enhancement of the downloaded file.


Few layer graphene based superlattices as efficient thermal insulators
Yuxiang Ni, Yann Chalopin, and Sebastian Volz

Citation: Appl. Phys. Lett. 103, 141905 (2013); doi: 10.1063/1.4824013


View online: http://dx.doi.org/10.1063/1.4824013
View Table of Contents: http://apl.aip.org/resource/1/APPLAB/v103/i14
Published by the AIP Publishing LLC.

Additional information on Appl. Phys. Lett.


Journal Homepage: http://apl.aip.org/
Journal Information: http://apl.aip.org/about/about_the_journal
Top downloads: http://apl.aip.org/features/most_downloaded
Information for Authors: http://apl.aip.org/authors
APPLIED PHYSICS LETTERS 103, 141905 (2013)

Few layer graphene based superlattices as efficient thermal insulators


Yuxiang Ni, Yann Chalopin, and Sebastian Volza)
Laboratoire d’Energ
etique Mol
eculaire et Macroscopique, CNRS UPR 288, Ecole Centrale Paris,
Grande Voie des Vignes, 92295 Ch^
atenay-Malabry, France

(Received 11 June 2013; accepted 17 September 2013; published online 1 October 2013)
While graphene and few layer graphene (FLG) are considered as having the highest thermal
conductivity in their in-plane directions, our molecular dynamics (MD) simulations however show
that those systems are also characterized by a superior thermal contact resistance, which could be
largely tuned with the layer number when in contact with a silica substrate. Taking advantages of
such a resistive interface, MD simulations show that SiO2/FLG superlattices have a thermal
conductivity as low as 0.30 W/m K, exhibiting a promising prospect in nano-scale thermal
insulation. These findings pave the way for an improved thermal management of nanoscale
systems such as thermal barrier coatings and phase change memory materials with atomic-scale
C 2013 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4824013]
super-insulators. V

Few layer graphene (FLG) was proposed as a material could be turned into efficient thermal insulators in the form
for heat removal owing to its extremely high in-plane ther- of FLG/SiO2 superlattices.
mal conductivity.1–5 In most practical devices, graphene The thermal contact resistance between FLG and SiO2 was
layers will however be encased within or supported on studied by MD simulations with LAMMPS software pack-
dielectrics such as SiO2.6 FLG was, for instance, envisioned age.11 Bernal stacking graphene layers (3  3.7 nm) and SiO2
to serve as a heat duct from active nano-devices to heat sink (3  3.7  2 nm) with periodic boundary conditions in the
regions.7 But as is the case in most molecular systems, the x and y directions are used, as shown in Fig. 1. MD simulations
excellent properties of the isolated graphene was found to consist in the computation of the atomic trajectories, where the
fade when embedded in a matrix or encased on a substrate. classical Newtonian equations of motion are numerically inte-
Atomic systems are indeed extremely sensitive to their envi- grated.12 We implemented the adaptive intermolecular reactive
ronment, since the mechanisms related to the low number of empirical bond order (AIREBO) potential13 to simulate the
internal atomic interactions could be easily dominated by the C-C interactions, which includes a Lennard-Jones (LJ) poten-
ones due to the interactions with external bodies. The intrin- tial for the non-bonded Van der Waals C-C interactions
sic thermal properties of atomic-scale compounds therefore between layers. The parameters in the LJ potential are CC
become less relevant than the ones generated by their ¼ 2:39 meV for the potential well depth and rCC ¼ 3:41 Å
interfaces. for the distance at which the inter-particle potential is zero.14 A
The thermal contact between graphene and the substrate modified Tersoff potential15 was used to describe the Si-Si, the
is known to be impeded by the weak Van der Waals interac- Si-O, and the O-O interactions in SiO2. Although Tersoff
tions at play. Due to the high interfacial resistance produced potential for SiO2 does not take into account the long-range
by those interactions, a system of carbon sheets appears as a Coulombic interaction, this is less likely to matter for the prob-
unique means to design heat paths at atomic scale. lem at hand as the interfacial thermal transport depends only on
The contact resistance between graphene and SiO2 was the Van der Waals interaction between the graphene and the
previously studied by different measurement methods, includ- substrate atoms. Ong and Pop16 also used the same potential to
ing Raman and electrical,8 Pump-Probe,9 as well as 3-omega calculate the thermal conductance between carbon nanotubes
methods.7 A broad interval of contact resistance values and SiO2 and reasonable results were obtained. LJ potential
ranging from 5.6  109 m2 K W1 to 4  108 m2 K W1 was adopted for the interfacial interactions between FLG
were reported. Despite a large number of studies on the gra- and SiO2, with SiC ¼ 8:909 meV; OC ¼ 3:442 meV; rSiC
phene and FLG-SiO2 contact resistances, those latter have ¼ 3:326 Å; and rOC ¼ 3:001 Å. The cutoff distances Rc of
always been considered as independent to the layer number,9 the LJ potential for the Si-C and O-C interactions were set to
since no correlation between resistance and layer number 2.5  r.16 The sensitivity to cutoff radius was checked with Rc
could be observed due to experimental inaccuracy. equals 4  r and 5  r, and no substantial variations in the
In this paper, we provide atomic scale simulation dem- thermal resistance were observed.
onstrations not only of the unequally high thermal contact re- Periodic boundary conditions were applied in the
sistance generated by supported graphene and FLG systems x and y directions. The structure was relaxed using a
but also of the possibility to tune this resistance with the steepest descent energy minimization algorithm (sd). The
number of layers. Taking advantages of such a high contact simulation first ran in the isothermal-isobaric (NPT) en-
resistance, low thermal conductivity of SiO2/FLG superlat- semble with a temperature of 300 K and a pressure of
tice with repeating interfaces was obtained from molecular 0 bar for 400 ps, then in the canonical ensemble (NVT) for
dynamics (MD) simulations. These findings show that FLGs another 200 ps to equilibrate the system. Atomic trajecto-
ries were then calculated in the microcanonical ensemble
a)
Electronic mail: sebastian.volz@ecp.fr. (NVE) at 300 K.

0003-6951/2013/103(14)/141905/5/$30.00 103, 141905-1 C 2013 AIP Publishing LLC


V
141905-2 Ni, Chalopin, and Volz Appl. Phys. Lett. 103, 141905 (2013)

normalized ACF of the temperature difference fluctuations


and the integration time, i.e., the integral of the ACFs versus
time, of 1,3,5-layer graphene-SiO2 interface are presented in
Fig. 2(b). The resistance between single layer graphene (SLG)
and silica was calculated as 3.34  108 m2 K W1. The value
of the contact resistance was found to decrease against the
layer number, and converges to 1.7  108 m2 K W1 for the
six-layer graphene. As shown in Fig. 2(a), the MD simulation
data are in satisfactory agreement with the experimental
values from Ref. 9, especially for FLGs with layer number
from 1 to 4. When the layer number is larger than 4, the
measurement gave a range of the resistances due to the
limited experimental accuracy, but the averaged value of
2  108 m2 K W1 (red dashed line) is quite close to our MD
data 1.7  108 m2 K W1.
The contact resistance decrease may be due to the stron-
ger coupling between FLG and substrate when there are
FIG. 1. Spatial configuration of a 5-layer graphene system supported by a more graphene layers interacting with SiO2. The potential
silicon dioxide substrate. cutoff radius of SiO2-FLG interaction was set to 2:5  rSiC
in the simulation, and the substrate could interact directly
The thermal resistance RC between two subsystems with with utmost three graphene layers. Judging from Fig. 2(b),
temperature difference DT can be calculated by the following the running time integral between 1 and 3 layer graphene-
equation:17 SiO2 interface are fairly close, but the number of degrees of
ð þ1   freedom of FLG atoms interacting across the interface, i.e.,
hDTð0Þ  DTðtÞi 1 1 the Ni in Eq. (1), increased from one-layer graphene to three-
Rc kB ¼ dt þ ; (1)
1 hDTð0Þ2 i N1 N2 layer graphene. In the cases of layer number larger than
three, Ni remained unchanged while the integration time has
where kB is the Boltzmann constant and Ni¼1,2 refer to the decreased (see blue dashed line in Fig. 2(b)). As a result, the
number of degrees of freedom of the kinetic energy of the calculated contact resistance is decreasing against layer
atoms in one subsystem that are interacting with the other number.
subsystem. Thus, Ni ¼ 3  Na, where Na is the number of FLG layers have interactions with their neighboring
atoms. Accordingly, DT(t) here is calculated from only the layers within a cutoff radius of 3  rCC . For the third FLG
interacting atoms across the interface, which is playing a key layer, it could interact directly with its third neighbor, i.e., the
role in the phonon transmission and resistance.18 The ther- sixth layer. As shown in Eq. (1), the method that we used for
mal resistance between the two subsystems could then be calculating resistance involves a temperature difference
obtained from the normalized autocorrelation function DT ¼ T1  T2 , the 3 layers that interact with SiO2 are in one
(ACF) of the temperature difference fluctuations measured at subsystem with a mean temperature T1 and the Si, O atoms
equilibrium. More than 20 independent simulations were that interact with C are in another subsystem with mean tem-
performed for each resistance point, until the resistance perature T2. The sixth layer interacts directly with the third
(averaged from different simulations) converged. The error one, so it has a direct effect on T1, thus, affecting the calcu-
bars were then defined by looking at the variance of the aver- lated results. Other graphene sheets except these six layers do
aged values. not have direct effect on T1, and their contribution to the con-
The contact resistance between FLG and SiO2 as a func- tact resistance is negligible. This may be the reason that the
tion of graphene layer number is reported in Fig. 2(a). The contact resistance saturates when layer number reaches six.

FIG. 2. (a) Contact resistance between


FLG and SiO2 substrate versus layer
number at 300 K: a comparison
between MD simulations and measure-
ments from Ref. 9. The red dashed line
is the averaged value of the experimen-
tal results. The black dashed lines are
the typical range of measured contact
resistance between graphene and SiO2
summarized from Ref. 10.
(b) Temperature difference autocorre-
lation functions (solid lines) and the
running time integrals (dashed lines)
for 1,3,5-layer graphene-SiO2 interfa-
ces at 300 K.
141905-3 Ni, Chalopin, and Volz Appl. Phys. Lett. 103, 141905 (2013)

Koh et al.19 measured the thermal conductance of Au/ 91.6 nm. This means that an encased graphene layer with an
Ti/FLG/SiO2 interfaces, and found that the total conductance effective thickness less than 0.7 nm, i.e., about twice the
is independent on the FLG layer number, which seems con- value of rSiC , provides the same thermal resistance as a
tradictory to our findings in Fig. 2(a) at first glance. 91.6 nm thick SiO2 layer. Given that SiO2 is already a good
However, considering the low cross-plane thermal conduc- thermal insulator due to its low thermal conductivity,34 gra-
tivity of FLG calculated from MD simulations20 and Debye phene systems encased by SiO2 should behave more like effi-
model21 (less than 0.1 W/m K for 6-layer graphene), the total cient thermal insulators than conductors in the direction
resistance of FLG will increase with layer number, which is perpendicular to the interface. Materials with such structures,
likely to cancel the FLG/SiO2 contact resistance reduction, such as the one of SiO2/graphene superlattices, are therefore
thus, the measured total conductance is not affected by the proposed to be used as planar thermal insulators while
graphene layer number. requiring a much reduced volume. To verify the projection
The contact resistance between FLG and SiO2 reveals of SiO2/graphene thermal insulator, we calculated the ther-
the interface phonon scattering, mostly the scattering mal conductivity of SiO2/SLG and SiO2/FLG superlattices
between FLG flexural modes and the substrate. It is widely (Fig. 4(a)) from the well-known Green Kubo formula35
reported that the strong interfacial scattering of flexural ð þ1
modes leads to a decrease of the in-plane thermal conductiv- V
k¼ hqð0Þ  qðtÞi dt; (2)
ity in supported graphene.22–24 Moreover, the increasing 3kB T 2 0
in-plane thermal conductivity of FLG supported on SiO2
with layer number measured by Jang et al.6 indicates where the brackets denote the heat flux autocorrelation func-
decreased interface scatterings, which corresponds to a tion, V is the system volume, and T is the temperature. Both
decreasing contact resistance between FLG and SiO2. These crystal SiO2 (c-SiO2) and amorphous SiO2 (a-SiO2) were
findings are in accordance with our simulation data. It would studied, and the thermal conductivities of the corresponding
also be interesting to note that the substrate-induced disrup- bulk SiO2 were also calculated as benchmarks. For superlat-
tions of the outermost graphene layers penetrate a finite dis- tices, the thermal conductivity was only calculated along the
tance into the core of the flake, and this characteristic direction perpendicular to the interfaces, by using the cross-
distance is reported to be 2.5 nm (about 7 layers),6 which is plane component of the heat flux.
fairly close to the distance at which the contact resistance It seems that Tersoff potential is not sufficient for the
converges in this study. thermal conductivity calculation since the long-range
The weak Van der Waals interactions between FLG and Coulombic interaction contribution to the thermal conductiv-
SiO2 generate a relatively large contact resistance, compared ity should not be neglected. As a result, the potential model
with that between many solids, as is shown in Fig. 3. previously developed by B.W.H. Beest, G.J. Kramer and
According to Fig. 3, compared to the other solid-solid inter- R.A. Santen namely BKS36 was used to simulate SiO2 to
faces, SiO2-graphene could provide a large contact resistance include the columbic force in the thermal conductivity
without requiring a very large mass mismatch. The graphene calculation.
encased by SiO2 could provide twice the contact resistance Amorphous SiO2 was produced by annealing the crystal-
due to the doubled interface, and this value is as high as line structure at 6000 K16 and then cooling down rapidly to
6.64  108 m2 K W1 if the largest contact resistance room temperature. Periodic boundary conditions were
obtained in our MD simulation is considered. Taking the applied to x, y, and z directions to generate the superlattice
thermal conductivity of SiO2 as 1.38 W/m K,33 the thickness structure. The period thickness L and the layer number of
of SiO2 that provides the same amount of resistance is FLG were varied to attain the lowest thermal conductivity.
As is shown in Fig. 4(b), the thermal conductivities of
both c-SiO2/SLG and a-SiO2/FLG superlattice increase with
increasing period thickness and approach the corresponding
bulk silica thermal conductivity. The thermal conductivity
increase with period thickness is due to the increased silica
proportion or reduced interface density. Amorphous SiO2/
SLG superlattice has generally lower thermal conductivity
than crystal SiO2/SLG, due to the lack of periodicity of the
crystalline structure and the normal modes are no longer
plane-waves in such a way phonons are damped as in anhar-
monic crystals.37 Unlike the strong interactions between two
solids, such as covalent bonds28 and nanojunctions,38 which
increase the thermal conductance, the weak Van der Waals
interactions between graphene sheet and SiO2 generate large
thermal resistance and reduce the thermal conductivity of the
corresponding superlattice. The lowest thermal conductivity
FIG. 3. Compilation of the thermal contact resistance of solid-solid interfaces appeared at the shortest period thickness, where SiO2 has
at 300 K. Data are taken from Ref. 25 (Bi/H-diamond), Ref. 26 (Pb/diamond
only one unit cell with a length of 0.6 nm.
and Al/Al2O3), Ref. 27 (CNT/Ag), Ref. 28 (CNT/PEMA), Ref. 29 (Si/Ge
superlattice), Ref. 30 (CNT/CNT), Ref. 31 (TiN/MgO and TiN/Al2O3), Ref. 32 From Fig. 2(a), the contact resistance between SLG and
(FLG/FLG), and this work (SiO2/graphene). SiO2 is about 3.34  108 m2 K W1. If this is repeated every
141905-4 Ni, Chalopin, and Volz Appl. Phys. Lett. 103, 141905 (2013)

FIG. 4. (a) Spatial configuration of


SiO2/SLG and SiO2/FLG superlattices.
(b) Period thickness dependence of the
thermal conductivities of c-SiO2/SLG
and a-SiO2/SLG superlattices. The ther-
mal conductivities of the corresponding
bulk SiO2 (parallel to the c-axis) were
also plotted as benchmarks. (c) Thermal
conductivity of a-SiO2/FLG superlattice
as a function of graphene layer number.
(d) Thermal conductivity of a-SiO2/
5-layer graphene superlattice as a func-
tion of temperature.

1 nm with two such interfaces, then the corresponding found to be decreasing with layer number and converged for
thermal conductivity would be roughly approximated as six graphene layers. The contact resistance between SLG
0.017 W/m K. The corresponding value reported in Fig. 4(b) and SiO2 was found to be 3.34  108 m2 K W1, which is
is around 0.35 W/m K, larger by a factor of 20 and suggest- more than one order of magnitude larger than the inter-layer
ing that the behavior of these samples is more complicated resistance in graphite. Owing to such a resistive interface,
than simple addition of series resistances. This phenomenon MD simulations show that SiO2/FLG superlattices have a
might be due to the fact that the period thicknesses are within low thermal conductivity of 0.30 W/m K, exhibiting a prom-
the characteristic length in which the contributions of the ising prospect in nano-scale thermal insulation. These find-
evanescent modes arise. Here, we define an evanescent mode ings provide useful information in the thermal design of
as a non-propagative surface mode that exists at the interface small-scale devices.
and that is locally confined over a typical thickness. Similar
results were observed in Si/Ge superlattices,29 where higher
thermal conductivities were found at very small period 1
A. A. Balandin, S. Ghosh, W. Z. Bao, I. Calizo, D. Teweldebrhan, F.
lengths. Miao, and C. N. Lau, Nano Lett. 8, 902 (2008).
2
A. K. Geim and K. S. Novoselov, Nature Mater. 6, 183 (2007).
Thermal conductivity was then calculated from the mod- 3
Y. K. Koh, M.-H. Bae, D. G. Cahill, and E. Pop, ACS Nano 5, 269 (2011).
els with a fixed a-SiO2 length (0.6 nm) but with different 4
K. Saito, J. Nakamura, and A. Natori, Phys. Rev. B 76, 115409 (2007).
FLG layer numbers. According to Fig. 4(c), the thermal con- 5
K. M. F. Shahil and A. A. Balandin, Nano Lett. 12, 861 (2012).
6
ductivity of a-SiO2/FLG superlattice slightly decreases with W. Jang, Z. Chen, W. Bao, C. N. Lau, and C. Dames, Nano Lett. 10, 3909
(2010).
the layer number and saturates to 0.30 W/m K at four gra- 7
Z. Chen, W. Jang, W. Bao, C. N. Lau, and C. Dames, Appl. Phys. Lett. 95,
phene layers. We also investigated the a-SiO2/5-layer gra- 161910 (2009).
8
phene superlattice at different temperatures up to 1000 K and M. Freitag, M. Steiner, Y. Martin, V. Perebeinos, Z. Chen, J. C. Tsang,
found that the thermal conductivity was not temperature de- and P. Avouris, Nano Lett. 9, 1883 (2009).
9
K. F. Mak, C. H. Lui, and T. F. Heinz, Appl. Phys. Lett. 97, 221904
pendent, as shown in Fig. 4(d). The lowest thermal conduc- (2010).
tivity of 0.27 W/m K found in the a-SiO2/FLG superlattice is 10
A. A. Balandin, Proc. SPIE 8101, 810107 (2011).
11
close to that of a suspended 10-layer graphene in the cross- S. Plimpton, J. Comput. Phys. 117, 1 (1995).
12
plane direction.20,21 The a-SiO2/FLG superlattice low ther- D. C. Rapaport, The Art of Molecular Dynamics Simulation, 2nd ed.
(Cambridge University Press, New York, 2004).
mal conductivity is however conserved whatever its total 13
S. J. Stuart, A. B. Tutein, and J. A. Harrison, J. Chem. Phys. 112, 6472
thickness, while the one of thicker FLG will significantly (2000).
increase.20,21 Another advantage of a-SiO2/FLG superlattice 14
15
L. A. Girifalco, M. Hodak, and R. S. Lee, Phys. Rev. B 62, 13104 (2000).
is that it could endure high temperatures of more than S. Munetoh, T. Motooka, K. Moriguchi, and A. Shintani, Comput. Mater.
Sci. 39, 334 (2007).
1800 K, while most polymers with equivalent thermal con- 16
Z. Y. Ong and E. Pop, Phys. Rev. B 81, 155408 (2010).
ductivities melt below 400 K, which is the typical working 17
A. Rajabpour and S. Volz, J. Appl. Phys. 108, 094324 (2010).
18
temperature of today’s thermal barrier coatings and phase Y. Chalopin and S. Volz, Appl. Phys. Lett. 103, 051602 (2013).
19
Y. K. Koh, M. H. Bae, D. G. Cahill, and E. Pop, Nano Lett. 10, 4363
change memory materials.
(2010).
To summarize, by using molecular dynamics simula- 20
Z. Y. Wei, Z. H. Ni, K. D. Bi, M. H. Chen, and Y. F. Chen, Phys. Lett. A
tions, the contact resistance between FLG and SiO2 was 375, 1195 (2011).
141905-5 Ni, Chalopin, and Volz Appl. Phys. Lett. 103, 141905 (2013)

21 30
K. Sun, M. A. Stroscio, and M. Dutta, Superlattices Microstruct. 45, 60 J. Yang, S. Waltermire, Y. Chen, A. A. Zinn, T. T. Xu, and D. Li, Appl.
(2009). Phys. Lett. 96, 023109 (2010).
22 31
J. H. Seol, I. Jo, A. L. Moore, L. Lindsay, Z. H. Aitken, M. T. Pettes, X. R. M. Costescu, M. A. Wall, and D. G. Cahill, Phys. Rev. B 67, 054302
Li, Z. Yao, R. Huang, D. Broido, N. Mingo, R. S. Ruoff, and L. Shi, (2003).
32
Science 328, 213 (2010). Y. Ni, Y. Chalopin, and S. Volz, Appl. Phys. Lett. 103, 061906 (2013).
23 33
Z. Wang, R. Xie, C. T. Bui, D. Liu, X. Ni, B. Li, and J. T. L. Thong, Nano M. B. Kleiner, S. A. Kuhn, and W. Weber, IEEE Trans. Electron Devices
Lett. 11, 113 (2011). 43, 1602 (1996).
24 34
Z.-Y. Ong and E. Pop, Phys. Rev. B 84, 075471 (2011). C. Zhang and K. Najafi, J. Micromech. Microeng. 14, 769
25
H.-K. Lyeo and D. G. Cahill, Phys. Rev. B 73, 144301 (2006). (2004).
26 35
R. J. Stoner and H. J. Maris, Phys. Rev. B 48, 16373 (1993). R. Kubo, M. Toda, and N. Hashitsume, Statistical Physics II (Springer,
27
B. A. Cola, J. Xu, C. Cheng, X. Xu, T. S. Fisher, and H. Hu, J. Appl. Phys. Berlin, 1985).
36
101, 054313 (2007). B. W. H. Vanbeest, G. J. Kramer, and R. A. Vansanten, Phys. Rev. Lett.
28
Y. Ni, H. L. Khanh, Y. Chalopin, J. B. Bai, P. Lebarny, L. Divay, and S. 64, 1955 (1990).
37
Volz, Appl. Phys. Lett. 100, 193118 (2012). S. Hunklinger, J. Phys. Colloques 43, 461 (1982).
29 38
Y. Chalopin, K. Esfarjani, A. Henry, S. Volz, and G. Chen, Phys. Rev. B L. Jalabert, T. Sato, T. Ishida, H. Fujita, Y. Chalopin, and S. Volz, Nano
85, 195302 (2012). Lett. 12, 5213 (2012).

View publication stats

You might also like