You are on page 1of 12

Available online at www.sciencedirect.

com
Available online at www.sciencedirect.com
Available online at www.sciencedirect.com

ScienceDirect
Structural Integrity Procedia 00 (2019) 000–000
Structural
Procedia Integrity
Structural Procedia
Integrity 26 00 (2019)
(2020) 000–000
187–198 www.elsevier.com/locate/procedia
www.elsevier.com/locate/procedia

The 1 stst Mediterranean Conference on Fracture and Structural Integrity, MedFract1


The 1 Mediterranean Conference on Fracture and Structural Integrity, MedFract1
A
A review
review on
on thermophysical
thermophysical properties
properties of
of flexible
flexible graphite
graphite
E. Solfitiaa , F. Bertoaa
a Department
E. Solfiti , F. Berto
of Mechanical and Industrial Engineering, Norwegian University of Science and Technology, Richard Birkelands vei 2B, Trondheim,
a Department of Mechanical and Industrial Engineering, Norwegian University
7491, Norwayof Science and Technology, Richard Birkelands vei 2B, Trondheim,
7491, Norway

Abstract
Abstract
Flexible graphite (FG) is the last stage of compression of expanded graphite. Natural and highly crystalline graphite flakes
areFlexible
expanded graphite (FG)together
and rolled is the last stageany
without of binder
compression
in orderoftoexpanded
shape foils graphite. Natural
of different and highly
thicknesses andcrystalline graphite
densities. The resultflakes
is an
are expanded
anisotropic butand rolled
highly together material,
conductive without anywithbinder in orderand
low strength to shape foils
stiffness of different
compared thicknesses
to common and densities.
graphite, The aresult
but gathering is an
collection
anisotropic but highly conductive material, with low strength and stiffness compared to common graphite, but gathering
of different peculiar properties. About mechanical field, FG is mostly used in sealing and gasketing applications due to its resilience a collection
of
anddifferent peculiar
capability properties.
of dissipate About
energy mechanical
during vibration.field, FG shows
It also is mostly
high used in sealing
resistance and gasketing
against chemicalapplications due retain
agents and can to its resilience
its struc-
and
turalcapability
integrity atoftemperature
dissipate energy
up toduring
2500◦◦ Cvibration. It also showsAll
in inert atmosphere. hightheresistance
mechanical,against chemical
electrical agents and
and thermal can retain
properties are its struc-
strongly
tural integrity at temperature up to 2500 C in inert atmosphere. All the mechanical, electrical and thermal properties
affected by the inherent anisotropy that make it exploitable in thermal cooling, interface insulation, electromagnetic shielding and are strongly
affected by the
in electrical inherent anisotropy
conduction componentsthat make
such it exploitable
as fuel in thermal
cell electrodes. In thiscooling, interface
work the insulation,process
manufacturing electromagnetic shielding
is summarized and
together
in electrical
with conduction
the effect of densitycomponents
on strengthsuch
and as fuel celland
modulus electrodes. In this work
the experimental the found
results manufacturing process
in literature amongis summarized together
anisotropy, stiffness,
with theand
thermal effect of density
electrical on strength
conductivity, and modulus
specific heat and and the experimental
thermal expansion areresults
reviewedfound in literature
together with someamong anisotropy,
modeling stiffness,
approaches.
thermal and electrical conductivity, specific heat and thermal expansion are reviewed together with some modeling approaches.
© 2020The
© 2020 TheAuthors.
Authors. Published
Published by Elsevier
by Elsevier B.V. B.V.
© 2020an
This The Authors. Published by Elsevier B.V.
This isisan open
open access
access article
article under
under the the CC BY-NC-ND
CC BY-NC-ND licenselicense (http://creativecommons.org/licenses/by-nc-nd/4.0/)
(http://creativecommons.org/licenses/by-nc-nd/4.0/)
This is an open
Peer-review under
Peer-review access article
underresponsibility under the CC
of MedFract1
responsibility BY-NC-ND
organizers
of MedFract1 organizers.license (http://creativecommons.org/licenses/by-nc-nd/4.0/)
Peer-review under responsibility of MedFract1 organizers.
Keywords: Thermal conductivity;Thermal exspansion;Specific heat capacity
Keywords: Thermal conductivity;Thermal exspansion;Specific heat capacity

1. Introduction
1. Introduction
Flexible graphite (FG) is a porous anisotropic material which substantially differs from graphite from various
Flexible
aspects graphite
including ore(FG)
originisplace,
a porous anisotropicprocess
manufacturing material which
and substantially
microstructure. Asdiffers from
it will be graphite
shown from in
hereafter, various
some
aspects including ore origin place, manufacturing process and microstructure. As it will be shown hereafter,
cases its properties resemble those of regular graphite but in others they can remarkably deviate. FG development in some
cases its properties resemble those of regular graphite but in others they can remarkably deviate. FG development
dates back to the end of the 60’s [Shane et al. (1968)] and it results from a rolling compression of expanded graphite
dates back to the
(EG) particles end ofany
without theadditive
60’s [Shane et Commercial
binder. al. (1968)] and
FGitcommonly
results from a rolling
appears compression
in the of expanded
form of sheets (or foils, graphite
at lower
(EG) particles without any additive binder. Commercial FG commonly appears in the form
thicknesses) ranging from 0.076 to 3 mm thickness and from 0.5 to 1.8 g/cm33 in density; other shapes of sheets (or foils, at lower
are available
thicknesses) ranging from 0.076 to 3 mm thickness and from 0.5 to 1.8 g/cm in density; other shapes
such as yarns and composite stacked arrangements. In some cases, mostly for scientific purposes, simple uniaxial are available
such as yarns and composite stacked arrangements. In some cases, mostly for scientific purposes, simple uniaxial

∗ Corresponding author. E.Solfiti Tel.: +39-3405863109


∗ Corresponding
E-mail address:author. E.Solfiti Tel.: +39-3405863109
emanuele.solfiti@ntnu.no (E. Solfiti)
E-mail address: emanuele.solfiti@ntnu.no (E. Solfiti)

2210-7843 © 2020 The Authors. Published by Elsevier B.V.


2210-7843
This © 2020
is an open The Authors.
access Published by Elsevier B.V.
2452-3216 © 2020 Thearticle under
Authors. the CC BY-NC-ND
Published license (http://creativecommons.org/licenses/by-nc-nd/4.0/)
by Elsevier B.V.
This is an open
Peer-review access
under article under
responsibility the CC BY-NC-ND
of MedFract1 license (http://creativecommons.org/licenses/by-nc-nd/4.0/)
organizers.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/)
Peer-review under responsibility of MedFract1 organizers.
Peer-review under responsibility of MedFract1 organizers
10.1016/j.prostr.2020.06.022
188 E. Solfiti et al. / Procedia Structural Integrity 26 (2020) 187–198
2 E. Solfiti and F. Berto / Structural Integrity Procedia 00 (2019) 000–000

compression of expanded particles can be applied giving lower density material which is often referred to as graphite
compact. Finally, one can point at all the density range material as compressed expanded graphite. It is clarified right
now that in the following review, concern will be about sheets, foils and compacts and not about different arrangements
of compressed expanded graphite. The exfoliated graphite, i.e. the expanded particles, is instead a granular compound
obtained by expansion of natural graphite flakes (purely crystalline, figure 1a) which show uncommon worm-like or
accordion-like shape once expanded upon heating (figure 1b). These stem from the mono-directional expansion of
flakes along the crystalline c-axis, perpendicular to the crystal basal planes. Because of that, sometimes the terms
expanded and exfoliated graphite are used in interchangeable manner. The exfoliated graphite itself is sometimes used
in industrial application as fire extinguisher, thermal insulator,lubricant addition or as a blending material in polimeric
composites [Chung (1987)]. Expandable or intercalated graphite instead represent the stage just before the expansion
in which the natural flakes have only been intercalated with the proper chemical specie. The ”worms” deform under

(a) (b)
Fig. 1: (a) SEM pictures of natural graphite flakes and (b) worm-like particles of exfoliated graphite. [By courtesy of Asbury Carbons Inc.]

compression and they are mainly arranged on the direction parallel to the bedding plane (perpendicular to the pressure
axis), retaining a certain content of air voids which decrease along with the increase of compaction pressure. In the
early stage of compression, the overall material can be defined as a heterogeneous disordered material with an higher
degree of inherent isotropy [Celzard et al. (2005)] whereas, after a certain value of density, the anisotropy comes out
to a large extent affecting the elasticity, strength, electrical conductivity, thermal conductivity and so on. Porosity and
resistance to chemical agent are involved in applications such as fuel cells and batteries electrodes [Luo et al. (2002);
Bhattacharya et al. (2004)] and working temperature range goes from -240◦ C to +2500◦ C [Sigraflex®], depending
on inert or oxidizing atmospheres: oxidation appears up to approximately +450◦ C on most commercial products
[Sigraflex® and Grafoil®]. From a mechanical point of view, such a microstructure results also on high resilience and
viscous response [Gu et al. (2002) and Luo and Chung (2000)]: graphite layers indeed show a typical hexagonal lattice
that is retained after expansion in the form of somewhat regular honeycomb cells which allow a certain amount of
sliding among layers and hence friction dissipative phenomena [Chen and Chung (2012)]. This capability of dissipate
energy together with the chemical resistance in a large range of temperature (up to beyond 2500◦ C) and low gas
permeability, make FG excellent in gaskets and sealing applications, often in sandwiched structures together with
stainless steel foils or as impregnated yarns. Because of that, the research trend often moved towards the investigation
in static compression and recovery loading [Toda et al. (2013), Wang et al. (2015) and Kobayashi et al. (2012)], often
at room temperature or even in order to observe the relation between the electrical properties variation as in Xi and
Chung (2019), Wei et al. (2010) and Luo and Chung (2000). A few mechanical tests have been reported about high
temperatures [Dowell and Howard (1986)] or fracture behavior [Gu et al. (2002) and Leng et al. (1998)] and no data
are available about fatigue life and damage mechanism. Because of its conformability (high capability to conform
to rough surface), FG is effective as thermal interface material for cooling and insulation [Marotta et al. (2005)]
and furthermore, having thermal and electrical conductivity comparable to or higher than metals such as copper and
E. Solfiti et al. / Procedia Structural Integrity 26 (2020) 187–198 189
E. Solfiti and F. Berto / Structural Integrity Procedia 00 (2019) 000–000 3

aluminum, main applications can range from satellites where air convection is not available [Gandhi and Pathak
(2012)] or smartphones, tablet, LED lightning and electrical vehicles cells cooling where the low weight and volume
are of crucial importance [NeoGraf Solutions, as viewed from website1 ], and going on, resistive heating [Chugh
and Chung (2002)], electromagnetic field shielding [Luo et al. (2002) and Sykam and Rao (2018)] and plasma facing
[Song et al. (2005)]. In a recent work Xi and Chung (2019) also observed the electret, piezoelectret and piezoeresistive
properties for mechanical sensing and electric powering.

2. Microstructure and mechanical properties

In this section some notions about the microstructure and the related mechanical properties are summarized. As
introduced in the previous section, increasing the pressure of compaction means to increase the density and the degree
of FG anisotropy. The scheme in figure 2 shows a summary of such a relationship. The expanded graphite particles

Density Graphite
(single crystal)
2.26 g/cm
3

Graphite compact Flexible graphite

0 g /cm
3 3 3
0.4 - 0.6 g/cm 1.7 - 1.8 g/cm

Natural flakes Uniaxial Rolling


compression Commercial FG foils
Expansion
Anisotropy

Isotropic Anisotropic Anisotropic - transversal isotropic

Fig. 2: Manufacturing process along with density and anisotropy behavior (pictures partially rearranged from Celzard et al. (2005) and explanatory
commercial FG from Sigraflex®.

(before the compression) own a rather organized honeycomb structure with an average thickness of ≈60 layers each
wall [Chen and Chung (2013)]. An example of experimental testing on a single work-like particle can be found in Gu
et al. (2002). Once compressed, the worms are in a deformed state that inherits in an extent some properties from the
expanded state. In particular, the sliding among the cell walls is still allowed resulting on outstanding visco-elastic
properties. Dissipative and visco-elastic phenomena in low density compacts have been largely studied from Luo
and Chung (2000), Chen and Chung (2012), Chung (2014), Chen and Chung (2015) and Xiao and Chung (2016).
Deformed worms are also the fundamental structural units that determine the overall mechanical properties like com-
pression strength and recovery [Toda et al. (2013)]; hereafter they will referred to as structural units. The interlocking
happens by mean of the only friction forces among their sharp edges and mandate a preferred orientation in the direc-
tion parallel to the bedding plane. The alignment in the two directions of the plane is random thus giving a transversal
isotropy with the increasing stages of compression [Celzard et al. (2005)]. Therefore properties like elasticity and
conductivity are usually observed in the two directions in-plane and out-of-plane. In rolled sheets, some differences
can be found also in the plane directions (parallel-to-rolling and perpendicular-to-rolling) as described by Dowell and
Howard (1986). Such structural units participate in two ways to the overall strength of the material: first, they under-
goes bending and twisting due to their stuck edges and second, mostly in compression loading, the trapped air inside

1 https://neograf.com/
190 E. Solfiti et al. / Procedia Structural Integrity 26 (2020) 187–198
4 E. Solfiti and F. Berto / Structural Integrity Procedia 00 (2019) 000–000

the expanded compressed closed regions gives a significant contribution to unloading recovery, as reported in Toda
et al. (2013). FG mechanical properties not only depend on the degree of worms interlocking but also on prior process
parameters such as original flake sizes, ash content, chemical intercalant species and exfoliated volume. Larger flakes
for example can result on higher exfoliated volume and therefore bigger structural units in which an eventual pull-out
stress would be hindered [Yoshida et al. (1991)]. In figure 3a such effect is outlined. Fully exfoliated worms (up to 300

8 12

7
10
6
8
5

4 6

3
4
2
2
1

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

(a) (b)

Fig. 3: (a) Effect on tensile strength (in-plane) due to the average flake size: 1. Leng et al. (1998), 2. Dowell and Howard (1986), 3. Wei et al.
(2010), 4. Gu et al. (2002); and (b) effect on tensile strength (in-plane) due to the density: 1. Leng et al. (1998), 5. Reynolds (1965). The intercalant
species used are H2 SO4 ,HNO3 and HClO4 .

times the original flake thickness) assure a tensile strength increase [Leng et al. (1998)]. Pure carbon content without
any residual metal oxides could both avoid the burn-out at high working temperature and also improve the overall
strength [Dowell and Howard (1986), Savchenko et al. (2010)]. Finally, a linear relation was found among density
and in-plane tensile strength as in figure 3b. An increasing trend is valid also for compression values against density
increment [Leng et al. (1998)]. Scarce data are available to properly define the main mechanical properties, also due
to a large variation of these against density, different manufacturing technique and parameters: a summary of values
for tensile and compression strength are reported in table 1.

Reference Density Tensile Young’s Compressive Compressive Poisson’s ratio


[g/cm3 ] strength modulus strength [MPa] modulus [GPa]
[MPa] [GPa]

Seldin (1966) (polycr. 1.72 - 1.94 13.44 - 30.33 8.96 - 17.7 0.07 - 0.22
graph.)
Dowell and Howard 1 ≈4 2.29 150 0.366 - 0.597
(1986)
Leng et al. (1998) 1.1 ≈4
Gu et al. (2002) 1 ≈3.5
Toda et al. (2013) 1 0.04
Xi and Chung (2019) 1 1.335
Sigraflex® 1 >4
Grafoil® 1.12 4.5 - 5.2 165 0.190

Table 1: Range of main mechanical properties (tensile properties in-plane, compression properties out-of-plane)
E. Solfiti et al. / Procedia Structural Integrity 26 (2020) 187–198 191
E. Solfiti and F. Berto / Structural Integrity Procedia 00 (2019) 000–000 5

Dowell and Howard (1986) reported a reduction on the ultimate tensile strain after heat treatment at 1750◦ C by a
factor of 2 whereas strength increases by 1.2. It is also worth to be reported the attempt to fit the stress-strain curve
with Jenkins (1962) relation used for polycrystalline graphite, that is
ε = Aσ + Bσ2 (1)
where A and B are two fitting coefficients. Such fitting model is found to be valid up to a half of the tensile strength.
Due to its low mechanical strength FG is quite complicated to test and some attempts of nanoindentation has been
carried out first from Chen and Chung (2015) and then from Khelifa et al. (2018) with good results.

3. Thermophysical properties

In the following sections the thermal conductivity, the coefficient of thermal expansion and the specific heat ca-
pacity of FG will be reviewed. Data from literature result limited and sometimes absent especially when dependence
on temperature is observed. Polycrystalline graphite and pyrolytic graphite without any additive will be the main ref-
erence for FG comparison due to their particular structure: pyrolytic graphite indeed have highly oriented graphite
crystals and due to its grade of purity it can be considered as the most similar material to pure cystalline graphite
[Kelly (1981)]. Data from commercial FG datasheet will be shown as well, in particular Sigraflex®, Grafoil® and
Papyex®.

3.1. Thermal conductivity

Thermal conductivity has been employed to describe the material compression mechanism during density growth
and thus the microstructure modeling. In the first section, the anisotropy has been introduced but there is not a unique
definition of it since depends on the property observed. In such sense, the thermal conductivity can be a good indicator
of the thermal response of the material:
λ
Conductivity anisotropy = (2)
λ⊥
where λ and λ⊥ are the thermal conductivities in the in-plane and out-of-plane directions respectively. The electrical
conductivity can be employed as well, since it is strictly related to the thermal conductivity, often holding the same
trend against density increasing. Celzard et al. (2005) measured values for the electrical conductivity anisotropy up to
≈70 around the conductivity threshold and for elasticity anisotropy (dynamic elastic modulus) beyond 10. In figure
4, the behavior of the thermal conductivity with the bulk density increase is displayed. The dashed lines indicate data
from commercial FG datasheets. At low density, all the materials tested are graphite compacts, whereas at high density
are either compacts or rolled foils. In figure 4a the thermal conductivity in the in-plane direction increases in a pretty
linear manner all along the wide density range as expected by the gradual orientation of the structural units along
the plane directions. Figure 4b instead deserves more discussion: almost all the data, except Chen and Chung (2014),
increase at the early density but decrease when overcome a certain value oscillating within 0.4 - 0.6 g/cm3 interval.
Such fact can be attributed to a still random orientation of the units whose interlocking enhances the conductivity in
both in-plane and out-of-plane direction indistinctly. After this region the conductivity decreases, as expected, being
hindered from already oriented particles. It is underlined that data from Chen and Chung (2014) seems to follow a quite
different trend in the first interval, that is decreasing with density, probably due to a different measurement technique.
Value of thermal conductivity anisotropy are calculated up to ≈32 - 35 at 1 g/cm3 , around a half of the electrical
conductivity anisotropy. If the same approach is used in order to calculate the tensile strength anisotropy, one can
obtain values beyond ≈100 at 1 g/cm3 [out-of-plane tensile data from Gu et al. (2002)]. The thermal conductivity,
moreover, is temperature dependent and despite FG can reach high working temperature in absence of oxygen, data
from literature are not available except those ones from datasheets of commercial FG. In figure 5 the comparison is
shown among various graphitic materials.
It should be noticed that commercial FG fall in the range delimited by the pyrolytic graphite (PG) thermal con-
ductivity in-plane and out-of-plane, which can be viewed as the upper and lower bounds of graphite-based material
Conductivity anisotropy for PG shows values above 300. FG conductivity in the in-plane direction also fits very well
192 E. Solfiti et al. / Procedia Structural Integrity 26 (2020) 187–198
6 E. Solfiti and F. Berto / Structural Integrity Procedia 00 (2019) 000–000

700 25

600
20
500

400 15
300

200 10

100
5
0

-100 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

(a) (b)

Fig. 4: (a) Thermal conductivity in the in-plane direction against bulk density and (b) thermal conductivity in the out-of-plane direction against
bulk density: 1. Py et al. (2001),2. Wei et al. (2010), 3. Liu et al. (2013), 4. Olives and Mauran (2001), 5. Bonnissel et al. (2001), 6. Papyex®,
7.Grafoil®, 8. Chen and Chung (2014).

10 5

4
10

10 3

2
10

10

0.1

0.01
0 500 1000 1500 2000 2500 3000

Fig. 5: Temperature dependence of thermal conductivity: 1. Ho et al. (1972), 2. Chen et al. (2011a), 3. Grafoil®, ρ = 1.12 g/cm3 , 4. Papyex®, ρ =
0.7 - 1.1 g/cm3 , 5. Sigraflex®, ρ = 1 g/cm3 .

the polycrystalline trend and follow the same behavior on the whole range of temperature. In the out-of-plane direc-
tion instead appears more similar to PG perpendicular to the main orientation of crystals. Copper and aluminum own
values higher than FG although in the same order of magnitude but, more fundamental, they cannot hold the same
working temperature of FG. For example, their melting point results indeed equal to 1357◦ C and 933◦ C respectively
[Ho et al. (1972)], which is considerably lower than 2500◦ C. The effect of increasing temperature acts on lattice
vibrations of graphite being useful for the conductivity at low temperature but having the opposite effect at high tem-
perature. FG in the in-plane direction seems to inherit the same effect whereas in the out-of-plane direction it keeps
a constant trend with small variations around a stable average. Despite the microstructure of PG is highly crystalline
and the manufacturing process is totally different from flexible graphite foils, it is possible to intuitively capture not
only that FG owns a mixed behavior among the two extreme bounds of PG but it holds such feature along the whole
working range of temperature. To design the conductivity in a material has a fundamental relevance in order to im-
prove the heat transfer capability and to better understand the heat transfer mechanism of FG. The atomic mechanism
of heat transfer is not well-known indeed and only Chen and Chung (2014) suggested a contribution of the crystalline
E. Solfiti et al. / Procedia Structural Integrity 26 (2020) 187–198 193
E. Solfiti and F. Berto / Structural Integrity Procedia 00 (2019) 000–000 7

a-axis (trough the graphene layers) also in the out-of-plane direction of FG. This would mean a phonon-based heat
transfer mechanism in all the directions. The mixing concept described above has often been applied when modeling
of the material conductivity has been attempted. The approaches present in literature are various but show often the
common starting point: the observation of micromechanism of material formation at very low densities. The material
is assumed to be a binary mixtures made by air and a solid part and some hypothesis or deductions [as in Bonnissel
et al. (2001) and Celzard et al. (2005)] are made about the pore shape. At higher densities, corresponding to rolled
commercial FG range, the conclusions are obtained by regression and compared with experimental data which some-
times result scarce. Celzard et al. (2005) observed the conductivity threshold at very low densities, that is the critical
density corresponding to a sudden increase of the inherent conductivity. Around such threshold the percolation theory
was found to be adequate on modeling the conductivity and elasticity behavior using power laws of density such as

λ ∝ (ρ − ρc )t (3)

where ρc is the critical density corresponding to the threshold and t is a fitting exponent. Another attempt of
modeling has been carried out by Bonnissel et al. (2001) by mean of the Hashin-Shtrikman model for two phases
compound [upper bound, Hashin and Shtrikman (1963)]. The thermal conductivity of the overall compact can be
expressed as a function of the thermal conductivity of the solid content λ s and the porosity P = 1 − ρ/ρG , being ρG =
2.26 g/cm3 the density of crystalline graphite:
 P 
λ = λs 1 − (4)
2+P
λ s is considered as a function mixing the contributions to the thermal conductivity from the in-plane direction and
out-of-plane direction of the solid content. This mixing approach seems to fit the experimental data with a low error
in a wide range of density, up to the density of commercial FG foils. Finally, Chen and Chung (2014) compared the
two approaches of Hashin-Shtrikman and the Rule of Mixtures:
λ = v s λ s + v a λa (5)
in the out-of-plane direction. v s and va are the volume fractions of the solid part and air, whereas λa ≈ 0 is the
conductivity of the air. The Rule of Mixtures does not consider the solid content orientation changes during the
compaction considering the conduction path as uniformly oriented in the same direction but still captures the trend
with increasing density.

3.2. Specific heat capacity

The specific heat capacity c p (T) at constant pressure is dependent on the temperature but, for example, usually a
constant reference value is employed to calculate the thermal conductivity from the thermal diffusivity α when using
the flash method [Parker et al. (1961)], by mean of the following equation:
λ(T ) = ρ · ĉ p · α(T ) (6)
No data are available in literature about the FG specific heat against temperature trend except data from datasheet
(see figure 6). Only one measured valued at room temperature is given by Bonnissel et al. (2001) and it results c p =
850 J/kgK. Grafoil® declares a value of 711 J/kgK at 21◦ C which is quite similar to that of polycrystalline graphite
[704 to 720 J/kgK when T = 19 - 25◦ C, Picard et al. (2006)]. The increase of graphite c p around room temperature
can be considered as linear whereas a polynomial fit is proposed when the range of T is extended such as in Butland
and Maddison (1973). The tendency of the copper and aluminum curve is markedly different from that of graphitic
materials, with a milder increment for copper and a positive concavity for aluminum.

3.3. Coefficient of thermal expansion

About the coefficient of linear thermal expansion αT , no data are available in literature. Only data from datasheets
are shown in figure 7 and compared with thermal expansion of copper and aluminum. Due to the manufacturing
process, FG foils result on a very low expansion, or even negative, along the in-plane direction which has been already
8194 E. Solfiti
E. Solfiti and F. Bertoet/ al. / Procedia
Structural Structural
Integrity Integrity
Procedia 26 (2020)
00 (2019) 187–198
000–000

2500

2000

1500

1000

500

0
0 500 1000 1500 2000 2500 3000

Fig. 6: Temperature dependence of FG specific heat capacity c p : 1. Butland and Maddison (1973), 2. Picard et al. (2006), 3. Lutcov et al. (1970), 4.
Bonnissel et al. (2001), 5. Brooks and Bingham (1968), 6. Davis et al. (2001), 7. Grafoil®, 8. Sigraflex®.

stressed in a large extent. αT along the out-of-plane direction instead reaches values beyond 30·10−6 K−1 . In Grafoil®
is suggested that in a first range of temperature up to 1000◦ C the foil shrinks (negative αT ) due to relaxation of residual
stresses of compression and then slightly expands when the temperature effect dominates. αT of both polycrystalline
graphite and PG has a stable plateau approximately from room temperature up to beyond 1000◦ C and this might be
resembled by FG in the same range. The similarities already shown among FG and other graphites induce to think
about a monotonic increase of FG thermal expansion along with temperature. Copper and aluminum are isotropic and
their thermal expansion is markedly higher than that of FG in the in-plane direction. Indeed, it has been shown in
comparison with αT in the out-of-plane direction in figure 7b. The trend is more similar to that of graphite with a fast
increase at low temperatures followed by a relatively lower slope in the region where graphite (and apparently FG)
shows the plateau.

2 60

1.5 50
1
40
0.5
30
0
20
-0.5
10
-1

-1.5 0

-2 -10
0 500 1000 1500 2000 2500 3000 0 500 1000 1500 2000 2500 3000

(a) (b)

Fig. 7: Temperature dependence of coefficient of thermal expansion αT : 1. Kellett and Richards (1964), 2. Harrison (1977), 3. Bailey and Yates
(1970), 4. Papyex®, 5. Sigraflex®, 6. Grafoil®, 7. Morgan (1972), 8. Hahn (1970), 9. Otte et al. (1963). For polycrystalline graphite the symbol ⊥
is intended as perpendicular to extrusion direction.
E. Solfiti et al. / Procedia Structural Integrity 26 (2020) 187–198 195
E. Solfiti and F. Berto / Structural Integrity Procedia 00 (2019) 000–000 9

3.4. Other relevant properties

As suggested in the previous sections, electrical and thermal conductivity have a strong relationship. For metals
this is usually described by the well-known Wiedemann-Franz law:
λ = LT σ (7)

where σ is the electrical conductivity and L the Lorentz number whose values for metals is equal to 2.44·10−8 WΩ/K2 .
Similar relations have been carried out for different types of graphites, for example

λ−1 = 2.93 · 10−3 σ−1 + 0.34 (8)

that works for graphitised coked based materials and PG [Kelly (1981) and Mason and Knibbs (1962)]. Typical value
of L for polycrystalline graphites is 1.2·10−6 WΩ/K2 [Powell (1937)] whereas for PG in the in-plane and out-of-plane
directions is respectively 2.9·10−6 WΩ/K2 and 5.4·10−5 WΩ/K2 [Chen and Chung (2014)]. With regard to FG, results
in Wei et al. (2010) gave L values from 5.6 to 6.2·10−6 WΩ/K2 and a final fit in the shape of a sigmoidal regression at
room temperature as follow

λ = 1168.4e−1/(3.5σ) + 102.2. (9)

Chen and Chung (2014) finally found a linear relation among the electrical and thermal conductivity in the out-
of-plane direction corresponding to L = 7.3·10−6 WΩ/K2 . Typical values for the electrical conductivity are reported
in figure 8, both against density and temperature. As expected, the values of σ in the in-plane direction increase
together with the compaction pressure i.e. the density, in the same monotonical manner as the thermal conductivity
does (figure 4a). The effect of temperature instead seems to be more effective, making the conductivity to increase for
a larger range of temperature. Despite effects due to different intercalant species were not found [Ionov et al. (2000)],
other effects can take action and give quite different values that remains anyway in the same order of magnitude.
Theoretically, both the thermal and electrical conductivity do not depend on the thickness of the specimen tested but
actually a slight difference is found in the σ results from Chugh and Chung (2002) at very low thicknesses (0.13 -
0.38mm). FG also inherits from graphite properties both a certain degree of thermoelectric power and piezoresistivity.

0.35 0.18

0.3 0.16

0.25 0.14

0.2 0.12

0.15 0.1

0.1 0.08

0.05 0.06

0 0.04
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 500 1000 1500 2000 2500 3000

(a) (b)

Fig. 8: (a) Density and (b) temperature dependence of FG electrical conductivity: 1. Wei et al. (2010), 2. Sigraflex®, ρ = 1 g/cm3 , 3. Ionov et al.
(2000), ρ = 1 g/cm3 , 4. Chugh and Chung (2002), ρ = 1.1 g/cm3 .

The first was quantified by Hoi and Chung (2002) in the out-of-plane direction. The voltage at the outer surfaces raises
approximately in a linear manner with a slope equal to -2.6 µV/◦ C when reported to the absolute scale. The second
instead were investigated recently from Xi and Chung (2019) who pointed out an increment of electrical resistivity
over 30% when the material is stressed up to 3.18 MPa. In the latter work moreover the behavior of FG as electret and
piezoelectret element has been studied for the first time. Finally, due to good compliance that stems from the porous
196 E. Solfiti et al. / Procedia Structural Integrity 26 (2020) 187–198
10 E. Solfiti and F. Berto / Structural Integrity Procedia 00 (2019) 000–000

microstructure, FG foils are good thermal interface materials. It was underlined by Gandhi and Pathak (2012) how
the contact pressure can modify it depending strongly on the application: increasing the contact pressure leads to a
minimization of the resistance, but overcoming such value can deform the surface and hence leading to an increment of
that. A modeling proposal were done by Marotta et al. (2005) in order to describe such trend of the contact resistance
against the contact resistance of FG foils with a percentage error of approximately 20%.

4. Summary

A short review of FG microstructure and related mechanical properties has been given in the first part and then the
thermal conductivity and the strong relation with the electrical conductivity has been observed both against density
and temperature. Their fundamental role on material modeling has been underlined. A large lack of data exists about
the coefficient of thermal expansion and specific heat capacity, especially their dependence on temperature. FG owns
a collection of properties due to the graphitical nature, but all of them are often partially investigated or involve
applications that do not require any further in-depth investigation. The heat transfer mechanism has not been well
clarified and nor heating rate effect is reported neither coupling of thermal effect and mechanical loading. Modulus and
strength changes are unknown at high temperature (a rough mention can be found in Dowell and Howard (1986) and
thermal cycling effect has never been reported. In an industrial world where the technological applications are looking
for enhanced performances at very low weight and costs, the ease of production process, even in large volumes, make
FG suitable for a more and more large variety of applications that could request an higher effort of the background
research.

References

Asbury Carbons Inc., URL: https://asbury.com/materials/graphite/.


Bailey, A.C., Yates, B., 1970. Anisotropic thermal expansion of pyrolytic graphite at low temperatures. Journal of Applied Physics 41, 5088–5091.
doi:10.1063/1.1658609.
Balandin, A.A., 2011. Thermal properties of graphene and nanostructured carbon materials. Nature Materials 10, 569–581. doi:10.1038/
nmat3064.
Bhattacharya, A., Hazra, A., Chatterjee, S., Sen, P., Laha, S., Basumallick, I., 2004. Expanded graphite as an electrode material for an alcohol fuel
cell. Journal of Power Sources 136, 208–210. doi:10.1016/j.jpowsour.2004.03.003.
Bonnissel, M., Luo, L., Tondeur, D., 2001. Compacted exfoliated natural graphite as heat conduction medium. Carbon 39, 2151–2161. doi:10.
1016/S0008-6223(01)00032-X.
Brooks, C.R., Bingham, R.E., 1968. The specific heat of aluminum from 330 to 890°K and contributions from the formation of vacancies and
anharmonic effects. Journal of Physics and Chemistry of Solids 29, 1553–1560. doi:10.1016/0022-3697(68)90097-8.
Butland, A.T., Maddison, R.J., 1973. The specific heat of graphite: An evaluation of measurements. Journal of Nuclear Materials 49, 45–56.
doi:10.1016/0022-3115(73)90060-3.
Celzard, A., Marêché, J.F., Furdin, G., 2005. Modelling of exfoliated graphite. volume 50. doi:10.1016/j.pmatsci.2004.01.001.
Chaudhuri, S., Gravano, L., Marian, A., 2004. Optimizing top-k selection queries over multimedia repositories. volume 16. doi:10.1109/TKDE.
2004.30.
Chen, P.H., Chung, D.D., 2012. Dynamic mechanical behavior of flexible graphite made from exfoliated graphite. Carbon 50, 283–289. URL:
http://dx.doi.org/10.1016/j.carbon.2011.08.048, doi:10.1016/j.carbon.2011.08.048.
Chen, P.H., Chung, D.D., 2013. Viscoelastic behavior of the cell wall of exfoliated graphite. Carbon 61, 305–312. URL: http://dx.doi.org/
10.1016/j.carbon.2013.05.009, doi:10.1016/j.carbon.2013.05.009.
Chen, P.H., Chung, D.D., 2014. Thermal and electrical conduction in the compaction direction of exfoliated graphite and their relation to the
structure. Carbon 77, 538–550. URL: http://dx.doi.org/10.1016/j.carbon.2014.05.059, doi:10.1016/j.carbon.2014.05.059.
Chen, P.H., Chung, D.D., 2015. Elastomeric behavior of exfoliated graphite, as shown by instrumented indentation testing. Carbon 81, 505–513.
URL: http://dx.doi.org/10.1016/j.carbon.2014.09.083, doi:10.1016/j.carbon.2014.09.083.
Chen, S., Moore, A.L., Cai, W., Suk, J.W., An, J., Mishra, C., Amos, C., Magnuson, C.W., Kang, J., Shi, L., Ruoff, R.S., 2011a. Raman measure-
ments of thermal transport in suspended monolayer graphene of variable sizes in vacuum and gaseous environments. ACS Nano 5, 321–328.
doi:10.1021/nn102915x.
Chen, S., Moore, A.L., Cai, W., Suk, J.W., An, J., Mishra, C., Amos, C., Magnuson, C.W., Kang, J., Shi, L., Ruoff, R.S., 2011b. Raman measure-
ments of thermal transport in suspended monolayer graphene of variable sizes in vacuum and gaseous environments. ACS Nano 5, 321–328.
doi:10.1021/nn102915x.
Chugh, R., Chung, D.D., 2002. Flexible graphite as a heating element. Carbon 40, 2285–2289. doi:10.1016/S0008-6223(02)00141-0.
Chung, D.D., 1987. Exfoliation of graphite. Journal of Materials Science 22, 4190–4198. doi:10.1007/BF01132008.
E. Solfiti et al. / Procedia Structural Integrity 26 (2020) 187–198 197
E. Solfiti and F. Berto / Structural Integrity Procedia 00 (2019) 000–000 11

Chung, D.D., 2012. Carbon materials for structural self-sensing, electromagnetic shielding and thermal interfacing. Carbon 50, 3342–3353. URL:
http://dx.doi.org/10.1016/j.carbon.2012.01.031, doi:10.1016/j.carbon.2012.01.031.
Chung, D.D., 2014. Interface-derived extraordinary viscous behavior of exfoliated graphite. Carbon 68, 646–652. URL: http://dx.doi.org/
10.1016/j.carbon.2013.11.045, doi:10.1016/j.carbon.2013.11.045.
Davis, J.R., et al., 2001. Copper and copper alloys. ASM international.
Dowell, M., Howard, R., 1986. Tensile and compressive properties of flexible graphite foils. Carbon 24, 311–323.
Fu, Y., Hou, M., Liang, D., Yan, X., Fu, Y., Shao, Z., Hou, Z., Ming, P., Yi, B., 2008. The electrical resistance of flexible graphite as flowfield plate
in proton exchange membrane fuel cells. Carbon 46, 19–23. doi:10.1016/j.carbon.2007.10.020.
Gandhi, J., Pathak, A.V., 2012. Performance evaluation of thermal interface material for space applications. Applied Mechanics and Materials
110-116, 135–141. doi:10.4028/www.scientific.net/AMM.110-116.135.
Grafoil®, . NeoGraf Solutions LLC. URL: www.neograf.com/products/grafoil-flexible-graphite/.
Gu, J., Leng, Y., Gao, Y., Liu, H., Kang, F., Shen, W., 2002. Fracture mechanism of flexible graphite sheets. Carbon 40, 2169–2176. doi:10.1016/
S0008-6223(02)00075-1.
Hahn, T.A., 1970. Thermal expansion of copper from 20 to 800 k - standard reference material 736. Journal of Applied Physics 41, 5096–5101.
doi:10.1063/1.1658614.
Harrison, J.W., 1977. Absolute measurements of the coefficient of thermal expansion of pyrolytic graphite from room temperature to 1200 k and a
comparison with current theory. High temperatures-High Pressures 9, 211–229.
Hashin, Z., Shtrikman, S., 1963. A variational approach to the theory of the elastic behaviour of multiphase materials. Journal of the Mechanics
and Physics of Solids 11, 127–140. doi:10.1016/0022-5096(63)90060-7.
Ho, C.Y., Powell, R.W., Liley, P.E., 1972. Thermal conductivity of the elements. Journal of Physical and Chemical Reference Data 1, 279–421.
Hoi, Y.M., Chung, D.D., 2002. Flexible graphite as a compliant thermoelectric material. Carbon 40, 1134–1136. doi:10.1016/S0008-6223(01)
00260-3.
Ionov, S.G., Avdeev, V.V., Kuvshinnikov, S.V., Pavlova, E.P., 2000. Physical and chemical properties of flexible graphite foils. Molecular
Crystals and Liquid Crystals Science and Technology Section A: Molecular Crystals and Liquid Crystals 340, 349–354. doi:10.1080/
10587250008025491.
Jenkins, G.M., 1962. Analysis of the stress-strain relationships in reactor grade graphite. British Journal of Applied Physics 13, 30–32. doi:10.
1088/0508-3443/13/1/307.
Jenkins, G.M., 1964. The thermal expansion of polycrystalline graphite. Journal of Nuclear Materials 13, 33–39. doi:10.1016/0022-3115(64)
90064-9.
Jenkins, G.M., Williamson, G.K., 1963. Deformation of graphite by thermal cycling. Journal of Applied Physics 34, 2837–2841. doi:10.1063/
1.1729818.
Kellett, E.A., Richards, B.P., 1964. The thermal expansion of graphite within the layer planes. Journal of Nuclear Materials 12, 184–192. doi:10.
1016/0022-3115(64)90139-4.
Kelly, B.T., 1981. Physics of graphite .
Khelifa, M., Fierro, V., Macutkevič, J., Celzard, A., 2018. Nanoindentation of flexible graphite: experimental versus simulation studies. Advanced
materials science , 1–11.
Kobayashi, M., Toda, H., Takeuchi, A., Uesugi, K., Suzuki, Y., 2012. Three-dimensional evaluation of the compression and recovery behavior in
a flexible graphite sheet by synchrotron radiation microtomography. Materials Characterization 69, 52–62. URL: http://dx.doi.org/10.
1016/j.matchar.2012.04.008, doi:10.1016/j.matchar.2012.04.008.
Leng, Y., Gu, J., Cao, W., Zhang, T.Y., 1998. Influences of density and flake size on the mechanical properties of flexible graphite. Carbon 36,
875–881. doi:10.1016/S0008-6223(97)00196-6.
Liu, R., Chen, J., Tan, M., Song, S., Chen, Y., Fu, D., 2013. Anisotropic high thermal conductivity of flexible graphite sheets used for advanced ther-
mal management materials. ICMREE 2013 - Proceedings: 2013 International Conference on Materials for Renewable Energy and Environment
1, 107–111. doi:10.1109/ICMREE.2013.6893625.
Luo, X., Chugh, R., Biller, B.C., Hoi, Y.M., Chung, D.D., 2002. Electronic applications of flexible graphite. Journal of Electronic Materials 31,
535–544. doi:10.1007/s11664-002-0111-x.
Luo, X., Chung, D.D., 2000. Vibration damping using flexible graphite. Carbon 38, 1510–1512. doi:10.1016/S0008-6223(00)00111-1.
Luo, X., Chung, D.D., 2001. Flexible graphite under repeated compression studied by electrical resistance measurements. Carbon 39, 985–990.
doi:10.1016/S0008-6223(00)00213-X.
Lutcov, A.I., Volga, V.I., Dymov, B.K., 1970. Thermal conductivity, electric resistivity and specific heat of dense graphites. Carbon 8, 753–760.
doi:10.1016/0008-6223(70)90100-4.
Marotta, E.E., Mazzuca, S.J., Norley, J., 2005. Thermal joint conductance for flexible graphite materials: Analytical and experimental study. IEEE
Transactions on Components and Packaging Technologies 28, 102–110. doi:10.1109/TCAPT.2004.843153.
Martin, W.H., Entwisle, M.F., 1963. Thermal expansion of graphite over different temperature ranges. Journal of Nuclear Materials 10, 1–7.
doi:10.1016/0022-3115(63)90111-9.
Mason, I., Knibbs, R., 1962. The thermal conductivity of artificial graphites and its relationship to electrical resistivity .
Morgan, W.C., 1972. Thermal expansion coefficients of graphite crystals. Carbon 10, 73–79. doi:10.1016/0008-6223(72)90011-5.
Nelson, J.B., Riley, D.P., 1945. The thermal expansion of graphite from 15°c. to 800°c.: Part I. Experimental. Proceedings of the Physical Society
57, 477–486. doi:10.1088/0959-5309/57/6/303.
Nelson, J.B., Riley, D.P., Tolpadi, S.S., Llewellyn, J.P., Smith, T., 1945. The thermal expansion of graphite: part II. Theoretical .
Nix, F.C., MacNair, D., 1941. The thermal expansion of pure metals: Copper, gold, aluminum, nickel, and iron. Physical Review 60, 597–605.
doi:10.1103/PhysRev.60.597.
198 E. Solfiti et al. / Procedia Structural Integrity 26 (2020) 187–198
12 E. Solfiti and F. Berto / Structural Integrity Procedia 00 (2019) 000–000

Olives, R., Mauran, S., 2001. A highly conductive porous medium for solid-gas reactions: Effect of the dispersed phase on the thermal tortuosity.
Transport in Porous Media 43, 377–394. doi:10.1023/A:1010780623891.
Olsen, L.C., Seeman, S.E., Scott, H.W., Douglas, M., Company, A., 1969. Expanded Pyrolytic Graphite : Structural and Transport .
Otte, H.M., Montague, W.G., Welch, D.O., 1963. X-ray diffractometer determination of the thermal expansion coefficient of aluminum near room
temperature. Journal of Applied Physics 34, 3149–3150. doi:10.1063/1.1729148.
Papyex®, . Mersen. URL: https://www.mersen.com/sites/default/files/publications-media/
6-gs-papyex-flexible-graphite-mersen.pdf.
Parker, W., Jenkins, R., Butler, C., Abbott, G., 1961. Flash method of determining thermal diffusivity, heat capacity, and thermal conductivity.
Journal of applied physics 32, 1679–1684.
Picard, S., Burns, D.T., Roger, P., 2006. Measurement of the Specific Heat Capacity of Graphite. Bureau International des Poids et Mesures , 1 –
31.
Powell, R., Touloukian, Y., 1973. Thermal conductivities of the elements. Science 181, 999–1008.
Powell, R.W., 1937. The thermal and electrical conductivities of a sample of Acheson graphite from 0°C. to 800°C. Proceedings of the Physical
Society 49, 419–426. doi:10.1088/0959-5309/49/4/312.
Py, X., Olives, R., Mauran, S., 2001. Paraffin/porous-graphite-matrix composite as a high and constant power thermal storage material. International
Journal of Heat and Mass Transfer 44, 2727–2737. doi:10.1016/S0017-9310(00)00309-4.
Reynolds, W.N., 1965. The mechanical properties of reactor graphite. Philosophical Magazine 11, 357–368. doi:10.1080/14786436508221862.
Sanchez-Coronado, J., Chung, D.D., 2003. Thermomechanical behavior of a graphite foam. Carbon 41, 1175–1180. doi:10.1016/
S0008-6223(03)00025-3.
Savchenko, D.V., Ionov, S.G., Sizov, A.I., 2010. Properties of carbon-carbon composites based on exfoliated graphite. Inorganic Materials 46,
132–138. doi:10.1134/S0020168510020081.
Savvatimskiy, A.I., 2005. Measurements of the melting point of graphite and the properties of liquid carbon (a review for 1963-2003). Carbon 43,
1115–1142. doi:10.1016/j.carbon.2004.12.027.
Seldin, E.J., 1966. Stress-strain properties of polycrystalline graphites in tension and compression at room temperature. Carbon 4, 177–191.
doi:10.1016/0008-6223(66)90079-0.
Shane, J.H., Russell, R.J., Bochman, R.A., 1968. Flexible graphite material of expanded particles compressed together. US Patent 3,404,061.
Sigraflex®, . SGL Carbon. URL: www.sglcarbon.com/en/markets-solutions/component/
sigraflex-universal-and-sigraflex-universal-pro/.
Song, Y.T., Yao, D.M., Wu, S.T., Weng, P.D., 2005. Thermal and mechanical analysis of the EAST plasma facing components. Fusion Engineering
and Design 75-79, 499–503. doi:10.1016/j.fusengdes.2005.06.187.
Sykam, N., Rao, G.M., 2018. Lightweight flexible graphite sheet for high-performance electromagnetic interference shielding. Materials Letters
233, 59–62. doi:10.1016/j.matlet.2018.08.066.
Tanaka, T., 1974. The thermal and electrical conductivities of LaB6 at high temperatures. Journal of Physics C: Solid State Physics 7. doi:10.
1088/0022-3719/7/9/001.
Taylor, R., Gilchrist, K.E., Poston, L.J., 1968. Thermal conductivity of polycrystalline graphite. Carbon 6, 537–544. doi:10.1016/
0008-6223(68)90093-6.
Toda, H., Tsubone, K., Shimizu, K., Uesugi, K., Takeuchi, A., Suzuki, Y., Nakazawa, M., Aoki, Y., Kobayashi, M., 2013. Compression and
recovery micro-mechanisms in flexible graphite. Carbon 59, 184–191. URL: http://dx.doi.org/10.1016/j.carbon.2013.03.008,
doi:10.1016/j.carbon.2013.03.008.
Tsang, D.K., Marsden, B.J., Fok, S.L., Hall, G., 2005. Graphite thermal expansion relationship for different temperature ranges. Carbon 43,
2902–2906. doi:10.1016/j.carbon.2005.06.009.
Tyler, W.W., Wilson, A.C., 1953. Thermal conductivity, electrical resistivity, and thermoelectric power of graphite. Physical Review 89, 870–875.
doi:10.1103/PhysRev.89.870.
Uher, C., Sander, L.M., 1983. Unusual temperature dependence of the resistivity of exfoliated graphites. Physical Review B 27, 1326–1332.
doi:10.1103/PhysRevB.27.1326.
Wang, F.Q., Zhang, L.Z., Li, Y., Gong, B.Y., Xu, H.Y., Xiao, G.K., Cai, R.L., 2015. Research on Compression Performance of Flexible Graphite
Packing Rings with Different Density. Procedia Engineering 130, 644–651. doi:10.1016/j.proeng.2015.12.287.
Wei, X.H., Liu, L., Zhang, J.X., Shi, J.L., Guo, Q.G., 2010. Mechanical, electrical, thermal performances and structure characteristics of flexible
graphite sheets. Journal of Materials Science 45, 2449–2455. doi:10.1007/s10853-010-4216-y.
Xi, X., Chung, D.D., 2019. Electret, piezoelectret, dielectricity and piezoresistivity discovered in exfoliated-graphite-based flexible graphite, with
applications in mechanical sensing and electric powering. Carbon 150, 531–548. URL: https://doi.org/10.1016/j.carbon.2019.05.
040, doi:10.1016/j.carbon.2019.05.040.
Xiao, L., Chung, D.D., 2016. Mechanical energy dissipation modeling of exfoliated graphite based on interfacial friction theory. Carbon 108,
291–302. doi:10.1016/j.carbon.2016.06.098.
Yoshida, A., Hishiyama, Y., Inagaki, M., 1991. Exfoliated graphite from various intercalation compounds. Carbon 29, 1227–1231. doi:10.1016/
0008-6223(91)90040-P.
Ziman, J.M., Levy, P.W., 1961. Electrons and Phonons . doi:10.1063/1.3057244.

You might also like