Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Journal Pre-proofs

Interface Mechanics in Carbon Nanomaterials-based Nanocomposites

Guorui Wang, Luqi Liu, Zhong Zhang

PII: S1359-835X(20)30449-8
DOI: https://doi.org/10.1016/j.compositesa.2020.106212
Reference: JCOMA 106212

To appear in: Composites: Part A

Received Date: 19 August 2020


Revised Date: 18 November 2020
Accepted Date: 20 November 2020

Please cite this article as: Wang, G., Liu, L., Zhang, Z., Interface Mechanics in Carbon Nanomaterials-based
Nanocomposites, Composites: Part A (2020), doi: https://doi.org/10.1016/j.compositesa.2020.106212

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.


1 / 33

Interface Mechanics in Carbon Nanomaterials-based


Nanocomposites

Guorui Wanga, b, Luqi Liua, * and Zhong Zhanga, *

a CAS Key Laboratory of Nanosystem and Hierarchical Fabrication and CAS Center for

Excellence in Nanoscience, National Center for Nanoscience and Technology, Beijing 100190,

China.
b Department of Mechanical and Industrial Engineering, University of Toronto, Toronto, ON

M5S 3G8, Canada

*Address correspondence to

Email: liulq@nanoctr.cn; zhong.zhang@nanoctr.cn


2 / 33

Abstract
Carbon nanomaterials (e.g., carbon nanotubes and graphene) have been deemed as
versatile building blocks to create a novel generation of nanocomposites desired for a variety of
commercial applications. Different from that of the conventional fiber-based composites, the
enormous interfacial areas, van der Waals forces dominated interface interactions as well as
multiple scale load transfer mechanism would greatly impact the mechanical behaviors of carbon
nanomaterials-based composites. Besides the widely concerned nanofiller-matrix interface, the
filler-filler interface has to be considered in nanocomposites. The main challenge is how to
monitor the load-transfer process and evaluate load-bearing capability of fillers in composites.
Here, we summarize the recent progress in experimental characterization of atomic-scale
interface mechanics, and clarify its importance on the reinforcement of nanocomposites. We also
highlight the significance of competed mechanism between intrinsic mechanical strength of
nanofillers and interfacial shear effect, allowing the delicate design of nanocomposites that
deliver desired properties.
3 / 33

1. Introduction
Nanostructured carbon materials (e.g., carbon nanotubes (CNTs), graphene, and fullerene)
have gained surging interests of both scientific and industrial community, owing to their vast
potential afforded by the unprecedented physical properties. Carbon nanomaterials incorporated
polymer composites featured with light weight, mechanical reinforcement, and
multifunctionality, become an efficient route to maximize these superior mechanical and
physical properties of individual building blocks at a macroscopic scale [1, 2]. Inspired by high
strength and toughness in natural composite materials (e.g., nacre, bone), many biomimetic
guided hierarchical designs were developed in carbon nanomaterials based composites, including
aligned CNT array, CNT bucky paper, continuous CNT composite fibers, nacre-like graphene
composites films etc [3-7]. Among these hierarchical structures, the key issues lie in the
understanding how structures and properties operate through interface across different length
scales within the materials. Not only the widely concerned filler-matrix interface, but also the
filler-filler interface and filler inside interface arising from multilayer structure features (e.g.,
multiwalled carbon nanotubes (MWCNTs) and graphene nanoplatelets ) need to be taken into
account [8, 9]. Furthermore, as the characteristic dimensions downscale to nanometers,
interfacial issues are anticipated to be further magnified in terms of relatively more significant
interfacial interaction that dominate the deformation behavior over huge interfacial areas [10, 11].
Compared with strong in-plane covalent bonding, van der Waals (vdW) interactions at pristine
interfaces between layers are considerably weak in resisting interlayer shear/slippage [12, 13],
potentially impairing the reinforcement effect. In this scenario, a full understanding of the
fundamental properties of these interfaces at different length scales is imperative to unlock the
ultimate potential of carbon nanomaterials at a macroscopic level.
To address these issues, there is a great need for the development of nanomechanical
testing techniques which can not only resolve the forces but also reveal deformation and failures
acting at the nanoscale. Unfortunately, the nanometer size of the reinforcing fillers poses
substantial difficulties in experimentally characterizing interfacial properties of nanocomposites.
Specifically, at the core of the challenge is monitoring the load transfer process and assessing
load-bearing capability of carbon fillers at the nanoscale. Recently, the application of
spectroscopic and probe techniques has allowed the quantitative interfacial evaluation with high
spatial resolution [14-17]. These measurements yield key parameters (e.g. interfacial shear
4 / 33

strength) to judge the efficiency of stress transfer across interfaces and, more importantly,
provide insight into failure limits and underlying mechanisms in nanostructures, linking the
interfacial properties to the macroscopic performance of nanocomposites. In addition, the vdW
nature of the interface sparks the need for the interface engineering to improve the overall
performance of composites. Though manipulating the surface chemistry of carbon nanomaterials
has proven fruitful in tailoring their interfacial properties with polymer matrix [18, 19], it calls
for a better understanding of the nanoscale structure-property relationship to evaluate the benefits
and disadvantages associated with the different functionalization degree, which facilitates the
optimal design of carbon nanocomposites.
Considerable volume of literature has been published reviewing the synthesis methods,
characterization techniques, macroscopic mechanical performance and application prospects of
carbon nanocomposites [20-23]. Instead, in this review, we survey recent advances in
experimental characterizations of carbon nanomaterials-based composites, with a particular
emphasis on interface mechanics at atomic scales. Special attention is paid to the measurement
and modification of interfacial shear interactions between both nanofiller/polymer and
nanofillers themselves. The critical role of these nanoscale interfaces in the reinforcement of
nanocomposites is unveiled, featuring a competing mechanism between in-plane stiffness and
interfacial shear effect. Having further highlighted the importance of interplay between
chemistry and physics of interfaces, this review offers a guideline for rational design of high-
performance carbon nanocomposite.

2. Interface mechanics in nanocomposites


2.1 Multiscale mechanics in nanocomposites
Graphene/CNT assembled macroscopic structures in forms of fiber, films, foams and
nanocomposites reminisce the hierarchal structural features of natural materials such as nacre,
bone, and spider silk, however, the mechanical performances are yet to approach natural
counterparts [3,5,6]. To simultaneously achieve high strength and toughness in those man-made
hierarchical structures, the major strategy is to engineer the interfacial interactions across
multiple length scales. Fundamentally, a deep understanding of the relationship between
hierarchical structures and its corresponding mechanical behaviors at multiple levels is required.
To address this issue, researchers face to develop novel experimental characterization and testing
5 / 33

methods to reveal the mechanical properties and shear interaction across multiple length scales.
Recent advances in mechanical measurements of the strength and stiffness of individual building
blocks (e.g. CNT, graphene) through the utilization of atomic force microscopy (AFM) and in
situ micro-electromechanical systems (MEMS) based testing devices, have obtained the great
achievement [24-27]. However, arising from the intrinsic larger surface energy of building
blocks, the mechanical interaction between constituents such as adjacent CNTs, or graphene
sheets, is still yet to be explored. Meanwhile, the deformation mechanisms starting from
nanoscale building blocks and their interaction to micro- and macroscale structures have to be
taken into account. Only through the comprehensive information based on the mechanical testing
and analysis at each length scale in the hierarchical structures, carbon nanomaterials can be
designed to emulate and surpass the properties of natural materials. Earlier review by Espinosa et
al. has addressed the importance of multiscale mechanics in nanocomposites [13], herein, we
mainly focus on the interfacial mechanics at nanoscale.

2.2 Interfacial shear stress transfer mechanisms


In conventional fiber-reinforced composites (e.g. Kevlar/epoxy composites, CF/epoxy
composite), the interfacial mechanics from matrix to fibers is well described by the conventional
shear-lag model, in which the perfect bonding and linear relationship between tangential relative
displacement and the interfacial shear stress are undertaken. [28, 29] The tensile stress transfers
from the matrix to the fiber by means of interfacial shear stress during the uniaxial stretching, as
shown in Fig. 1. Beyond a certain strain level, the interfacial debonding occurs, where the
friction between fiber/matrix is usually averaged out at macroscopic level [30]. Unlike the
fracture behavior of brittle carbon fibers undergoing when matrix was subjected to a large strain
level, the stress transfer between the nanofiller and matrix is still taking place through interfacial
friction at the atomic scale. Molecular dynamics (MD) simulations about vdW interaction
dominated graphene interfaces reveal that the adhesion and sliding force would spontaneously
take place even without an applied normal pressure [31]. Therefore, nanoscale frictional behavior
at interfaces should be taken into account for carbon nanomaterials-based composites. A non-
linear elastic behavior was observed in response to a uniaxial tensile loading for nanocomposites
due to the weak vdW force-dominated nanofiller/matrix interface. In detail, the linear stress
transfer occurred at earlier stage, and then followed by an interfacial sliding as the interfacial
6 / 33

shear stress reached a critical value (i.e., the interfacial shear strength). Based on the fact, instead
of widely utilized linear shear-lag model, a nonlinear shear-lag model was found to better
interpret the interfacial behavior featuring the interfacial shear sliding between the nanofiller and
matrix [32, 33]. Moreover, to improve the poor vdW dominated interfacial interactions in
nanocomposites, chemistry engineering is a promising direction to build various chemical
crosslinking (e.g. covalent, coordination, ionic, hydrogen bonds) between nanofiller and polymer
besides solving the dispersion, processing issues. Specifically, stemming from the breaking and
reforming process of hydrogen bonds (H-bonds) or vdW interactions at nanoscale interfaces, the
graphene-polymer interface could still keep its capacity of transferring loads due to the friction
forces [34]. Recent literature in graphene-based nanocomposites further highlighted the
strengthening-toughening synergy can be implemented by introducing multimodal interfacial
interactions [35].

Fig. 1. Schematic of multiscale interface shear transfer in carbon nanomaterials-based


nanocomposite.

3. Mechanical Characterization of Nanoscale Interfaces


In spite of remarkable mechanical properties of graphene and CNTs [24, 27], their
atomically smooth surface typically leads to weak out-of-plane bonding— one of the central
limitations in scaling up the intrinsic properties of carbon nanofillers to macroscopic composite
materials. Thus, a key step towards the design of high-performance carbon nanomaterials-based
composites is to quantify the interfacial properties, especially shear interaction, between
7 / 33

nanofiller and polymer matrix as well as adjacent nanofiller themselves. In light of the atomic
thickness and low-dimensionality of graphene and CNTs that invalidate conventional methods
for measuring interfacial properties, tremendous efforts have been devoted to advance
characterization techniques that are capable of capturing interface mechanics at nanoscale. In the
following sections, we will focus on the mechanical characterization of different interfaces
presented in carbon nanomaterials-based nanocomposites (Fig. 2).

Fig. 2. Schematic drawing and corresponding electron microscopy images of various nanoscale
interfaces in carbon nanocomposites, including (a) CNT/polymer interface [36] (Copyright
permission Elsevier), (b) graphene/polymer interface [18] (Copyright permission Nature
Publishing Group), CNT/CNT interface in (c) bundles [25] (Copyright permission Wiley) and (d)
double-walled CNT [37] (Copyright permission Nature Publishing Group), (e)
graphene/graphene interface [38] (Copyright permission Nature Publishing Group).

3.1 CNT-polymer interface


8 / 33

The one-dimensionality of CNTs is reminiscent of microfibers pull-out test that is widely


applied in the measurement of interfacial properties between reinforcements and polymers [39].
Wagner et al. first realized the detachment of a MWCNT from an epoxy matrix with a scanning
probe microscope (SPM) tip in TEM [40], which was later repeated by Ruoff et al. using a
homebuilt nanomanipulator in a scanning electron microscope (SEM) [41]. The CNTs protruding
from the fracture surface were found to be covered with a polymer sheathing layer,
demonstrating the substantial MWCNT-polymer interactions (Fig. 3a) [41]. To quantitatively
estimate the shear strength of the CNT-polymer interface, AFM was used to manipulate the CNT
in relation to the polymer and record their force-displacement curves during the pull-out process
with a high precision, as shown in Fig. 3b. The pull-out force (peak force) was measured to
deduce the average interfacial stress, giving a value of 47 MPa [42].
In above mentioned study, the CNT attached to the AFM tip was usually dipped into a
liquid polymer which was then cooled or cured to allow the implementation of pull-out testing.
However, this method fails to be extended to vast numbers of engineering thermoplastics such as
polypropylene, polycarbonate, polyetheretherketone etc. In order to overcome the limitation of
the matrix type, single-CNT pullout tests were carried out by a “push-pull” micro-device
mechanically actuated by a nanoindenter in SEM, as seen in Fig. 3c [43]. The bottom figure
clearly shows that the single CNT is fully pulled out, leaving a cylindrical cavity in the matrix.
Additionally, the smooth surface of the cavity is an evidence of the weak interface between CNT
and the matrix. Similarly, assuming perfect interfacial bonding and uniform shear stress
distribution, the interfacial shear strength exhibit average value of 9.62 MPa, varying with
different embedded lengths as well as the CNT diameters [44].
Obviously, various studies reported a large spread of values of the interfacial shear
strength between CNT and polymer matrix. In addition to the diameter and embedded length of
CNT in the matrix, different preparation/processing methods (e.g. twin screw extrusion, solvent
casting, etc.) may cause the discrepancy in the interfacial heterogeneity and bonding strength
[42-48]. More importantly, the CNT used in these studies are actually multiwalled samples. As
the intershell interaction is much weaker than the interfacial bonding between CNT and polymer,
one potential explanation for the scattered results is the intershell shear sliding between
nanotubes during the pullout test. Furthermore, when using the tip-based technique for force
measurement, it is technically difficult to judge whether both cores and shells are subjected to
9 / 33

loading. If only partial outer shells are attached on the tip, the obtained interfacial shear strength
would be definitely underestimated.
In analogy to the mechanical characterization of the interface between conventional
single fiber (e.g. CF, Kevlar) and polymer system developed fifty years ago [14], in situ tensile-
Raman testing was also involved to evaluate the stress transfer between the CNT and polymer
matrix in nanocomposites. Figure 3d presents the typical Raman 2D-band shift of CNT
composites under different strains [5]. The downward shifted trend indicated that the stretching
of C=C bond in CNT under the tensile deformation. After a certain applied strain level, the
Raman 2D-band reached the plateau region, indicating the occurrence of slippage at interface.
However, it is worth noting that the interfacial shear strength of CNT based nanocomposites
cannot be derived directly due to the smaller size of individual carbon nanotubes. The same
method was also utilized to evaluate the effect of interface bonding types on the load transfer
efficiency, where the relatively large shifted rate was observed for hydroxyl groups modified
single-walled CNT (SWCNT) /polyvinyl alcohol (PVA) composites in comparison with the pure
SWCNT based PVA composites [49]. To quantitatively compare the effect of varied interface
interactions on load transfer efficiency at a microscale, a strain transfer factor (STF) was
proposed, which revealed the true strain of nanotubes in response to applied macroscopic strain.
More importantly, the proposed STF favors a more accurate estimation of the effective modulus
of composites based on a modified rule of mixtures taking into the interfacial stress transfer
efficiency. It should be emphasized that the in situ tensile micro-Raman method reflect the
averaged shifted rate of many CNT fibers due to its microscale spatial resolution (~1 μm).
10 / 33

Fig. 3. (a) In situ SEM nanomanipulation experiment with an AFM tip and high resolution
observation of CNT coated with a polymer sheath protruding from the fracture surface [41]
(Copyright permission American Chemical Society). (b) Typical plot of pull-out force taken
from the AFM cantilever deflection against pull-out time. The inset is an AFM height imaging of
the pull-out area showing an exit hole in the polymer [42] (Copyright permission American
Institute of Physics). (c) MEMS-based device to perform the pull-out experiments and SEM
snapshots showing a CNT pulled out of the matrix [43] (Copyright permission Elsevier). (d)
Downshifts of the Raman 2D-band in different CNT-based systems. The inset is the schematic of
reticulate CNT structure reinforced composites [5] (Copyright permission American Chemical
Society).

3.2 Graphene-polymer interface


Different from individual nanotubes, the individual graphene flake with microscale
lateral size makes it possible to directly monitor the strain distribution through in situ tensile-
11 / 33

Raman method [50]. Young’s group first employed this method to monitor the shear stress
transfer within polymethyl methacrylate (PMMA)/graphene/SU8/PMMA model system. On the
basis of well-established linear shear-lag model, the derived maximum interfacial shear strength
was around 2.3 MPa. Once the matrix strain was increased to 0.6%, the interfacial slippage
occurred with an approximately linear distribution of axial strain from the edge to the center of
the monolayer graphene along with an apparent dip in the middle of the flake [51].
Comparatively, the derived interfacial shear strength for graphene based polymer
nanocomposites in the range of several MPa level is at least an order of magnitude lower than
CNT reinforced composites (~40 MPa) [42] as well as carbon fibers reinforced composites (~20-
40 MPa) [52, 53]. To strengthen the weak vdW interface between graphene and matrix, the
chemical modification of monolayer graphene flake becomes an effective approach. Our group
did a systematic investigation of the bonding types (vdW interface vs. H-bond interface) on the
interfacial shear strength values for the graphene/PMMA model composite system through in
situ tensile Raman measurements which was mechanically loaded by bending the cantilever
beam [34]. The various oxygen-containing groups (e.g. hydroxyl groups, carboxylic groups)
were successfully attached on the edges and plane of graphene sheets. As shown in Fig. 4a,
during the loading process, Raman 2D-band of pristine graphene displays a linear red-shifted
trend with a slope of −51 cm−1/% in the elastic stress transfer stage. The Raman shift rate is
approaching the universal value of −60 cm−1/% for the freestanding graphene [54, 55], indicating
of an efficient strain transfer at the well-bonded interface. Beyond the critical strain for the onset
of interfacial shear sliding, a plateau stage is reached ascribed to breaking and reforming nature
of vdW interactions to maintain the stress transfer [37, 56]. Hence a constant shear stress can be
expected and leads to a linear strain distribution near the edges that eventually forms a triangle-
like shape as the interfacial sliding zone propagates towards the center with increasing strain, as
presented in Fig. 4b. Based on this, experimentally, the interfacial shear strength could be
determined by the monitoring the plateau strain at the center of the nanofiller with the aid of in
situ Raman spectroscopy measurements, on the level of 0.5 MPa. Similarly, the interfacial shear
strength between graphene and polyethylene terephthalate (PET) substrate was reported in the
range of 0.46 to 0.69 MPa [32]. Comparatively, as for the H-bond dominated interface, the
interfacial shear strength values show apparent 2-3 times improvement up to 1.7 MPa, depending
on the oxygen-containing groups attached onto the graphene flakes.
12 / 33

Fig. 4. (a) Raman 2D-band shifts recorded at the center of monolayer graphene flake/PMMA
model composite during a loading/unloading cycle. (b) Strain distribution in the stretching
direction within graphene sheet at different strain levels based on Raman line mapping of 2D-
band shifts [34] (Copyright permission American Chemical Society).

3.3 Carbon nanotube-carbon nanotube interface


In view of experimental challenges related to the nano-sized diameter of CNTs, AFM
sensors or nanomanipulators are typically integrated into the in situ electron microscope to allow
precise handling and loading of CNTs, as well as the visualization of its structural evolution
during the deformation process. Earlier studies by Zettl et al. first reported the visualization of a
sword-in-sheath failure of an individual MWCNT in transmission electron microscope (TEM)
[57]. The AFM cantilever was used as the force sensor to detect the force when the core shells
were pulled out of the outer shells. The interlayer shear strength was measured to be lower than
0.05 MPa, which is comparable with that of graphene/graphite. Furthermore, such weak
intershell friction was found to be further reduced with absence of defects and large axial
curvatures in CNTs [37]. Specially, the pull-out experiment was performed on centimeter-long
double-walled carbon nanotubes (DWCNTs) in SEM equipped with nanomanipulators (Fig. 5a).
A probe was first used to pull out the inner shells which were then transferred to the force
cantilever of a silicon nanorod for the force measurement in following deformations (Fig. 5b).
The calculated shear strength between nanotube shells can reach as low as 2.6 Pa, showing an
independence of nanotube length. This is because the intershell interaction only depends on the
edge section while in the overlapped section the shear stress vanishes as vdW forces repetitively
break and reform during the pull-out process.
13 / 33

The atomically smooth carbon surface also causes the individual nanotubes to easily slip
past each other within the bundles when subject to external loading. Similar to the sword-in-
sheath failure mechanism observed in MWCNTs, inner bundles of DWCNTs can be pulled out
from the outer host using an in situ SEM testing setup with a Si cantilever-based force sensor, as
shown in Fig. 5c [58]. For a better comparison of the interactions for bundles with different
diameters, the number of intertube interactions at shear interfaces was evaluated to normalize the
pull-out force (Fig. 5d), which was translated to an average interfacial shear strength of around
7.8 MPa [15]. However, such a value only corresponded to the lower limit for the true interfacial
shear strength as the estimation assumed the interfacial area to be continuous, while the fact is
that the discrete CNT-CNT interfaces would cause highly non-uniform interfacial shear stress in
the axial direction. To this end, recent studies have demonstrated the fabrication of superstrong
CNT bundles that are centimeters long, with continuous lengths and uniform initial strains [59].
The tensile strength can be as high as 80 GPa, even higher than that of crosslinked CNT bundles
(~17 GPa) which is limited by the degradation of the crystalline structure [25].

Fig. 5. (a) Schematic of the experiment illustrating a process of pulling out the inner shell from
its outer host through a sensitive force probe [60] (Copyright permission Nature Publishing
14 / 33

Group). (b) Measured intershell friction for the ultralong MWCMTs [37] (Copyright permission
Nature Publishing Group). (c) Sequential SEM images recorded during shear testing of an
isolated bundle of DWCNTs. (d) Force-displacement curve recorded during the pull-out
experiment [58] (Copyright permission American Chemical Society).

3.4 Graphene-graphene interface


Generally, graphene and its chemical derivate graphene oxide (GO) are obtained from
chemical treatment of natural graphite under harsh experimental conditions to overcome vdW
forces between the adjacent layers. Unfortunately, it has been reported that such a top-down
approach never produces 100% monolayer graphene flakes, but a statistical distribution of
graphene stacks instead [8, 61]. Thus, the reinforcing effect in those multilayer graphene or
graphene oxide-based composites would be undermined by weak interface within neighboring
graphene layers, which haven’t received intense attention in the composite fields. Indeed, there
are competition in term of shear stress between filler-matrix interfaces and adjacent interlayers
once composites are subjected to tensile and/or compression deformation. Earlier works have
revealed that change of stacking order in few-layer graphene flakes can be identified by Raman
spectroscopy, allowing the discrimination of Raman responses between separate layers with
interlayer decoupling [62, 63]. As mentioned earlier that the interfacial shear strength value for
graphene–polymer interface through vdW force was around few MPa. Comparatively, it is hard
to directly measure the interlayer shear stress of multilayer graphene sheets due to the
ultrathinness.
In combination of in situ tensile-Raman spectroscopy and AFM, recently, we have
proposed a versatile method to measure the interlayer shear stress in bilayer graphene based on
pressurized bulging devices (Fig. 6a) [12]. The underlying mechanism lies in that the bubble
induced radial tension tends to pull the surrounding graphene on the substrate towards the hole
while the shear stress is activated to resist the relative sliding at interface. Such a competition
results in an annular shear zone with finite outer radius, as shown in Fig. 6b. By delicately tuning
the pressure difference across graphene sheets, controllable propagation of shear zone can be
activated and captured through monitoring the Raman G-band shifts, owing to the stress-
sensitive characteristics of Raman peaks of carbon nanomaterials [54, 64]. Considering the
relatively stronger interaction between graphene and supporting silicon substrate, the shear zone
around bilayer graphene bubbles is demonstrated to arise from more significant interlayer shear
15 / 33

deformation attributed to the weak interlayer shear resistances. Even the bilayer graphene was in
Bernal (AB) stacking configuration, the interlayer shear stress of 0.02-0.06 MPa was extracted
based on a membrane analysis for microscale bubbles (Fig. 6c), because of the averaged effect
induced by the limited spatial resolution of the Raman laser spot. More recently, Androulidakis
et al. have reported the presence of apparent interlayer sliding between two randomly stacked
graphene layers on polymer substrate [65]. Similar to our work [34, 50], a four-point-bending
setup under the Raman microscope (Fig. 6d) was employed for simultaneously mechanically
loading the sample and recording the evolution of Raman spectra. It was found that with
increasing strain, Raman 2D-band split into two subpeaks that exhibit distinct Raman shift rates
(cm-1/%), suggesting different stress transfer efficiency at graphene-graphene and graphene-
polymer interfaces. Interfacial sliding was clearly observed at a critical strain that occurred
earlier between graphene layers due to weaker interactions. By balancing the shear to axial stress
using the classical shear-lag theory (Fig. 6e), interlayer shear stress can be deduced based on
Raman maps of peak shift across the length of graphene flake (Fig. 6f), in the range of ~0.04–
0.13 MPa and ~0.04–0.16 MPa for exfoliated and CVD samples respectively.
In addition to high-frequency Raman measurement, the low-frequency Raman shear
mode has recently been uncovered to be a diagnostic tool to explore the interlayer shear coupling
between graphene layers [66-69]. Tan el al. reported the first measurement of the Raman
signature of the shear mode (C-band) of multilayer graphene [66]. Typically, the Raman C-band
is located at 42 cm-1 for bulk graphite, and the peak position apparently decreases as the
thickness downscales to few layers, reaching 29 cm-1 for bilayer graphene. Such a thickness
dependence can be interpreted by a linear chain model, allowing the extraction of the interlayer
shear modulus of 4.3 GPa. More recently, our group employed this method to examine the effect
of chemical doping on the interlayer coupling of multilayer graphene [70]. We unveiled that the
substitution of boron atoms in graphene lattice caused a ~20% reduction of the interlayer shear
modulus, due to the enlarged interlayer spacing. Such an attenuated interlayer coupling could
greatly weaken the shear interactions during the slippage and suppress the energy dissipation in
boron-doped multilayer graphene under external loading.
The interfacial shear stress in multilayer graphene is directly related to the vdW friction
force, which shows dependence on the relative angle between upper layer and lower layer. An
example of a representative is the concept of superlubricity presented in the incommensurate
16 / 33

interface [71]. Through the self-retraction experiments on square graphite mesas, Zheng’s group
observed microscale superlubricity under ambient conditions (Fig. 6g and h) [72]. In particular,
lithographically defined graphite mesas can be easily sheared out with a tungsten microtip and
spontaneously slide back after being released. By measuring the sheared distance, the ultralow
frictional stress was estimated to be 0.02-0.04 MPa. Interestingly, if the sheared flake was rotated
at certain angles, the frictional stress was greatly enhanced to around 0.1 GPa as the interface
locks in at commensurable state (Fig. 6i). Additionally, nanoscale friction testing based on
atomic force microscopy (AFM) provides another avenue for characterizing the superlubricity
between atomically clean surfaces [73]. In this case, graphene-coated AFM tips are required and
can be prepared by either mechanical transfer or chemical vapor deposition (CVD) growth
method (Fig. 6j). By sliding the tip relative to the graphite or graphene coated substrate in the
direction perpendicular to the cantilever long axis, the interlayer friction was achieved at the
contact area, showing a superlubric state (Fig. 6k). According to the fitting of normal load-
friction force curve with a contact mechanics model, the interfacial shear strength can be further
deduced. For example, the interlayer shear strength between adjacent GO layers was measured to
be 5.3 ± 3.2 MPa, much higher than that of pristine graphene, mainly due to the stronger
interlayer bonding with the presence of intercalated functional groups [74].
To gain a more comprehensive understanding of interfacial properties in carbon
nanocomposites, the interfacial shear strength obtained from different methods are summarized
in Table 1.
17 / 33

Fig. 6. (a) Schematic diagram of a bilayer graphene bulging device and characterizations. (b)
Raman contour maps of G-band frequency in one quadrant revealed the strain distributions in
bilayer graphene bubbles in both suspended and supported regions at different pressures. (c)
Measured shear stresses between graphene layers [12] (Copyright permission American Physical
Society). (d) Four-point-bending setup for uniaxial tension of bilayer graphene with random
incommensurate stacking. (e) Schematic demonstration of the stress transfer mechanism from
polymer substrate to the bottom layer and from the bottom layer to the top layer. (f) The
distribution profile of Raman 2D-band frequency across the length of top layer graphene of the
folded bilayer [65]. (g) Schematic illustration and (h) optical observation of the self-retraction
18 / 33

experiment. (i) Friction anisotropy indicating the 60° symmetry of lock-in orientations, between
which are regions featuring self-retraction [72] (Copyright permission American Physical
Society). (j) Tribological test of graphene-coated microsphere probe by friction force microscopy.
(k) Plot of friction force as a function of the applied normal load for different sliding materials
[73].

Table 1. Experimentally measured interfacial shear strength for different interfaces in carbon
nanocomposites.

Interfacial shear
Material interface Measuring method Ref.
stress/strength (MPa)
SWCNT PVA In-plane tension + Raman 188 [75]
PU Fragmentation in TEM 500 [46]
Polyethylene-
10–90 [42, 45]
butene AFM
Epoxy 35–376 [40]
MWCNT
PEEK Nano-manipulation in SEM 3.5-14 [48]
Epoxy 1.8-12.5 [47]
Polymer- MEMS-based tensile testing
7.6-49.3 [43, 44]
derived ceramic
Chemically
Epoxy AFM 20-138 [76]
modified MWCNT
PMMA/SU8 0.3-0.8 [51]
SU8 4-point bending + Raman 0.15-0.75 [77-80]
Exfoliated graphene PMMA 0.16-0.3 [81, 82]
PET In-plane tension + Raman 0.46-0.69 [32]
PMMA Cantilever bending + Raman 0.5-0.7 [34, 50]
PET In-plane tension + Raman 0.004-0.314 [83]
CVD graphene
PEMA nano-bubble inflation 0.48 [84]
Oxidized graphene PMMA Cantilever bending + Raman 0.8-1.7 [34]
Nano-manipulation in TEM 0.05 [57]
Intershell (MWCNT) AFM cantilever method in
0.08–0.3 [85]
SEM
Intershell (DWCNT) Nano-manipulation in SEM 2.6×10-5 [37]
Intertube (SWCNT/MWCNT) AFM 4 [86]
AFM cantilever method in
Intertube (CNT bundle) 7.8 [58]
SEM
Bulging test + Raman +
Exfoliated bilayer graphene 0.02-0.06 [12]
AFM
Exfoliated folded graphene 0.04-0.13
4-point bending + Raman [65]
Transferred bilayer CVD graphene 0.04-0.16
Graphite/graphite Frictional force microscopy 0.25-0.75 [71]
19 / 33

Graphite/graphite
0.02-0.04 [72]
(incommensurate)
Shearing/retraction test
Graphite/graphite
100-140 [72, 87]
(commensurate)
Graphene oxide/ Graphene oxide Frictional force microscopy 5.3±3.2 [74]

Note:
SWCNT: single-walled carbon nanotube; DWCNT: double-walled carbon nanotube; MWCNT: multi-
walled carbon nanotube; PVA: polyvinyl alcohol; PU: polyurethane; PEEK: poly-ether-ether-ketone;
PMMA: polymethyl methacrylate; PET: polyethylene terephthalate; PEMA: poly(ethyl methacrylate)

4. Link between Atomic-scale Interface and Bulk Mechanical Behaviors


4.1 Interlayer shear effect on intrinsic mechanical properties of vdW materials
Graphite is a well-known solid lubricant, a property that originates from the low
interlayer shear strength between individual graphene layers. The interface between graphene
layers is considerably weak, with interfacial shear strength even one magnitude lower than that
between graphene and polymer. The easy interlayer slip was demonstrated to induce
inhomogeneous strain distribution and stacking faults, having a significant impact on the
intrinsic mechanical properties of multilayer graphene. For example, a reduction in the effective
Young’s modulus and strength was reported as graphene scales up in the thickness dimension
[62, 77, 88, 89]. In addition to the in-plane properties, our group has recently conducted the
pressurized bubble experiment, for the first time investigating the interlayer shear effect on the
bending rigidity that controls the out-of-plane deformation of multilayer two-dimensional (2D)
materials (Fig. 7a) [90]. According to the classical plate theory [91], a membrane-to-plate
transition can be anticipated in terms of the deflection profile of 2D nanosheets with decreasing
bubble height or increasing thickness. Particularly, the spherical cap shaped bubble corresponds
to a membrane deformation behavior and allows the estimation of Young’s modulus (E) of 2D
materials assuming a negligible bending rigidity (Fig. 7b). In the plate regime, the bubble
becomes bump-like with a smooth edge, involving both the in-plane stretching and bending of
the 2D material, so that the bending rigidity (D) can be calculated by adopting a nonlinear plate
analysis (Fig. 7c) [92]. Interestingly, even though the bending rigidity increases with the
thickness of 2D materials as expected, the thickness dependency violates the classical plate
theory (𝐷~𝑡3, dashed line). As shown in Fig. 7c, the measured bending rigidity is much lower
than the upper bound that assumes a perfect bonding, yet still higher than the ultralubricated case
20 / 33

as the lower bound, suggesting a finite shear resistance between atomic sheets. Apparently, the
interlayer shear sliding is easily induced during the out-of-plane deformation and accounts for
the degradation of the bending rigidity. Thicker 2D layers with more interlayer interfaces would
subject to more interlayer slippage and hence more notable weakening. Likewise, increasing the
bending angle and strain also exacerbates the shear sliding and results in lower bending rigidity
[38, 93].
Extend to other reinforcing 2D materials, such as hexagonal boron nitride (hBN) is
frequently involved in recent composite due to its excellent thermal conductivity and insulating
properties [94, 95]. Unlike the weak vdW forces in multilayer graphene, the strong interlayer
coupling is revealed between hBN, which might on one hand hinder the effective exfoliation and
give rise to a poor dispersion in polymer matrix; on the other hand, it allows for a more efficient
stress transfer through the network of nanofillers. The stronger interlayer interaction, along with
the high in-plane stiffness in comparison with graphene, makes multilayer hBN potentially
outperform graphene as the mechanical reinforcement in nanocomposites under complex loading
conditions. Therefore, besides dispersion capability and cost of nanofillers, careful selection of
reinforcing fillers is an important step for the design and preparation of composites, especial for
multilayer 2D materials fillers. Second, the intrinsic parameters of 2D materials such as Young’s
modulus, bending stiffness, shear modulus and strength, are still absent, especially, the
dependence of the classic plate theory regarding D~E relationship is breakdown. As a result, we
have to carefully choose the reasonable D and/or E values to make a comparison between
experimental results and theoretical predictions.
21 / 33

Fig. 7. (a) Cross-sectional view of a pressurized microbubble formed by a multilayer nanosheet


over a microwell in the silicon substrate. Insets are illustrations comparing the microstructural
deformation upon bending of a multilayer vdW material with ultralubricated (left) and perfectly
glued (right) interfaces respectively. (b) Young’s modulus and (c) bending rigidity of three
different vdW materials as a function of thickness [90] (Copyright permission American Physical
Society).

4.2 Chemical functionalization vs. interfacial shear strength vs. in-plane strength
Surface modification is one of the important steps to improve the dispersion capability of
filler inside matrix, but also the interfacial adhesion between nanofillers and polymer matrix [96,
97]. Over the past twenty years, various chemical strategies have been proposed to meet the
demand in the wide range of applications from individual CNT- and graphene-based devices to
multifunctional nanocomposites. From the composite viewpoints, the chemical modification with
low cost, easy process, mass product, and environmental friendly routine are the major concern.
Meanwhile, it is worth noting that the defects are inevitably introduced onto nanofiller surface
during the chemical treatment process, which would compromise the intrinsic mechanical and
electrical properties of nanofillers [98, 99]. Thus, the optimal functionalized degrees of
reinforced nanofiller has to be taken into account based on the competing mechanism between
defect-related degradation of intrinsic mechanical property and stronger bonding induced higher
22 / 33

stress transfer. The effective modulus of graphene nanofiller should include the effect of defect
density, interfacial stress transfer efficiency and interlayer stress transfer efficiency for
multilayer graphene. Our earlier works reported that effect of the functionalization of nanofillers
on the interfacial adhesion and the failure modes at the microscopic level for the monolayer
graphene/PMMA nanocomposites system by means of in situ tensile micro-Raman and AFM
technique [34]. Due to the formation of H-bonds between graphene and the PMMA matrix, the
interface was strengthened with increasing functionalization degrees. The interfacial shear
strength of the functionalized graphene was found to reach up to 1.7 MPa, which approximately
tripled that of pristine graphene. However, distinct from the failure mode of interfacial sliding
observed for the graphene sheets with moderate functionalization degree, excessive oxygen-
containing groups would create defects in the graphene sheet and induce the crack initiation and
propagation under tension, as shown in Fig. 8. There is therefore a balance to be struck in the
design of graphene-based nanocomposites between the ability to achieve high interfacial
adhesion and the reduction in the elastic modulus of the functionalized graphene sheets. The
optimized functionalization degree has to be considered to achieve the maximum effective
modulus for the nanocomposites.

Fig. 8. (a) Strain distribution of monolayer graphene with excessively high functionalized degree
at different strain levels. Drops of strain correspond to the formation of cracks that result in the
failure of stress transfer. AFM images of oxidized graphene sheet (b) before the tension and (c)
after the release clearly show cracks at three marked sites, with the height profile presented in the
insets. (d) Raman image of D-band intensity of oxidized graphene sheet after tension, implying
that the oxygen-containing groups in the oxidized graphene sheet act as defect sites to induce the
initiation of cracks. (e) Schematic illustrating the fracture mechanism where crack growth
behavior of graphene is provoked from tensile stress [34] (Copyright permission American
Chemical Society).
23 / 33

4.3 Impact of topography upon the reinforcing efficiency of nanofillers


Graphene easily subjects to out-of-plane deformation due to its atomic thickness and high
flexibility. The resulting wrinkles and folds/buckles greatly affect the efficiency of graphene to
reinforce the nanocomposite [100]. Depending on whether the delamination occurs, these
topological undulations play different roles in the stress transfer process (Fig. 9a-c). For instance,
our previous study found that the interfacial sliding induced strain mismatch between graphene
and polymer substrate resulted in the compression in graphene to form buckle delamination
during the unloading. These buckles not only impair the contact and adhesion at
graphene/polymer interface, but also separate the entire flake into patches with lateral
dimensions below the critical length, which is detrimental to the overall interfacial load transfer
[50]. On the contrary, if the wrinkled graphene still conforms to the polymer substrate, the
interfacial shear strength can be maintained or even improved [80, 101]. This is explicable in
terms of the rough surface of graphene that leads to enhanced nanofiller/matrix adhesion and
interlocking at the interface as shown in Fig. 9d. In addition, the wrinkled network was found to
further improve the interlayer friction in few-layer graphene and manifest a thickness
independence of the interfacial stress transfer efficiency. The maximum shear stress can reach as
high as ∼0.75 MPa for a trilayer wrinkled graphene/polymer interface, much higher than values
obtained from flat, simply supported-graphene/polymer systems at similar strain levels [80].
Apparently, creating the wrinkle patterns of graphene provides a new avenue to enhance the load
bearing capability of graphene fillers. Moreover, it outperforms chemical surface treatments as
mechanical integrity of carbon nanofillers is maintained without the introduction of structural
defects.

Fig. 9. (a) Schematic diagram of (a) wrinkled and (b) folded graphene. (c) Schematic of strain
distributions for wrinkled and folded graphene [101] (Copyright permission IOP Publishing). (d)
24 / 33

Schematic of qualitative stress transfer of wrinkled multilayer graphene on polymer substrate [80]
(Copyright permission American Chemical Society).

4.4 Self-stiffening in nanocomposites


The vast interfacial area featured in nanocomposites is beneficial to build interface-
directed dynamic materials with stimuli-responsiveness, adaptivity, dynamically configuration,
self-stiffening under mechanical loading, which cannot be observed in conventional composites
[1]. Recent works have shown the apparent improvement in stiffness for composites in response
to a cyclic compression or stretch [102-104]. Figure 10 shows that, upon the repeated loading
condition, the initial randomly oriented polymer chains at interface region started to undergo
reorientation and realignment around the nanofillers, and eventually resulted in dynamically
increased interface stiffness. Meanwhile, the stronger stiffening effect was also assigned to the
formation of locally crystallized polymer chains because of highly ordered polymer interface.
Besides the morphology evolution of polymer chains in composites, the crumpled, curved
nanofillers started to stretch straightly, orient and align along the loading direction, leading to
discernable stiffening in bulk films. For example, compared to neat
graphene/polydimethylsiloxane (PDMS) composites, GO/PDMS composites having 2 times
PDMS interface thickness, giving rise to more pronounced stiffening effect. Furthermore, the
stiffening effect shows a dependence on the chemical bonding types at interface [103, 104].
Because of the breakable-and-reformable characteristics of H-bonds as well as vdW forces
between CNT or graphene and matrix [31,35], the stress could continuously transfer from matrix
to filler through the shear sliding at the interface to facilitate more energy dissipation, which
helps to enhance the toughness as well as mechanical strength simultaneously.
25 / 33

Fig. 10. Schematic illustration of self-stiffening behaviors of carbon nanomaterials-based


composites in response to cyclic stressing or compression loading. Inset figure adapted from Ref
[102] (Copyright permission American Chemical Society).

5. Conclusions and Outlook


Over the past decade, the pursuit of multifunctional nanocomposites has expanded
towards the deterministic, hierarchical design of predictable and optimized composite structures,
which relies heavily on the understanding of interfacial interactions across multiple length scales.
In this review, we present a full account of vdW dominated nanoscale interfaces incorporated in
carbon nanomaterials-based nanocomposites. We cover extensively various experimental
attempts to quantify the shear interactions between the constituents and reveal considerably weak
interlayer shear stresses between carbon nanofillers, which are at least one order of magnitude
lower than that of polymer/carbon nanofiller interface. It is therefore highly desirable to
incorporate the previously overlooked interlayer shear effect in the mechanical interpretation to
gain a deeper insight into reinforcement mechanism of carbon nanofillers. Regarding the
interfacial mechanics of carbon nanomaterials, experimentally, the integration of scanning probe
methods with spectroscopic techniques, such as tip-enhanced Raman spectroscopy and scanning
near-field optical microscope (SNOM), will provide unprecedented fundamental insights into
interfacial behavior at the space limit. Furthermore, relative sliding or twisting graphene layers
26 / 33

with respect to each other typically gives rise to moiré superlattice structures; these patterns act
as a magnifying glass for strain and in turn give direct information on the interlayer interactions
at the atomic scale. Beyond that, given the fact that vdW interfaces are less stable than the
surroundings, their dynamic behaviors deserve more attention. One representative example is the
graphene fatigue induced by interfacial failure through the vdW interactions under cyclic loading
[105]. In this context, in-situ TEM characterization will be instrumental to gain an atomistic
understanding of how interfaces respond to external stimuli. The advances in dynamic
transmission electron microscopy (DTEM) make it possible to visualize the dynamic response of
carbon nanostructures as well as their interfaces with atomic resolution. Theoretically, while
continuum mechanics theories have been demonstrated to be applicable to carbon nanomaterials,
more efforts are required to bridge the gap between continuum mechanics and atomic level
analysis. Besides, mechanics at the nanoscale interfaces is complicated with coupled adhesion
and shear interactions [106]. For example, edge/boundary effect has been recently reported for
multilayer assembly of graphene and CNT bundles by MD simulations, which is originated from
the interlayer/intertube adhesion energy that affects the interlayer/intertube sliding resistance
[107, 108]. However, the influence of interface adhesion on the stress transfer as well as
deformation and failure mechanisms at the interface receives limited attention in the study of
carbon nanomaterials-based nanocomposite. A generalized mixed-mode traction-separation
relationship should be hence adopted to better describe the interfacial interactions between
carbon nanofillers and polymer matrix. From a practical point of view, the effect of
environmental factors such as temperature, humidity and airborne contaminants on the interfacial
interactions should also be quantified.
Important but not-as-yet fully explored issues as to the competing mechanism between
interfacial strengthening and structural degradation upon chemical modification are also
discussed in this review. To close the gap between the high expectations and technological
achievements with carbon nanocomposites, we should establish the detailed structure-property
relationship of functionalized interfaces to optimize the chemical design strategy. Selective
chemical modification with uniform distribution of the surface functional groups tends to afford
a route to maximize the reinforcing effect, yet it remains technically challenging for arbitrary and
precise control, and calls for more experimental and theoretical efforts. Further interface
27 / 33

engineering through synergistic interactions (e.g., vdW forces, H-bonds, covalent bonds, and
coordinate bonds [109]) to enhance the interfacial binding will be the next important direction.
Inspired by biological materials such as nacre, in addition to tuning interfacial
interactions, translating carbon nanomaterials into well-defined hierarchical architectures is also
advantageous to enhance the shear transfer across interfaces. In particular, the design of
continuous network structures through chemical strategy would maximize the load bearing
capability of reinforcing fillers to overcome the intrinsic weakness of short fiber-based
composites. Alternatively, the orientational control of carbon nanomaterials in a polymer matrix
results in a high order at the macroscale and improves the stress transfer efficiency. Furthermore,
when polymer chains are restricted between aligned nanofillers, the extensive interfacial
interactions aid to suppress the molecular mobility and activate the nanoconfinement effect,
leading to the chain stiffening. In addition, the introduction of hydrogen bonding rather than
covalent bond between filler and matrix make it possible to dissipate more energy during the
interfacial sliding process, and eventually lead to the high strength and toughness. Consequently,
the nanocomposite film with optimized weight percentage of the nanofillers exhibits the highest
known Young’s modulus and tensile strength. In spite of great advances in the research of
layered nanocomposite, in-depth studies on the basic principles of some emerging assembly
techniques are necessary to offer guidance for effective control over the fabrication and further
improvement of the mechanical performance.

Acknowledgements

This work is jointly supported by National Natural Science Foundation of China (Grant Nos.
22072031, 11890682, 11832010, and 21474023). The Strategic Priority Research Program of
Chinese Academy of Sciences (CAS) under the Grant Nos. XDB36000000, XDB30201000. the
National Key Basic Research Program of China (Grant No. 2013CB934203)

Reference

[1] Kinloch IA, Suhr J, Lou J, Young RJ, Ajayan PM. Composites with carbon nanotubes and graphene:
An outlook. Science 2018;362(6414):547-53.
[2] Papageorgiou DG, Li Z, Liu M, Kinloch IA, Young RJ. Mechanisms of mechanical reinforcement by
graphene and carbon nanotubes in polymer nanocomposites. Nanoscale 2020;12(4):2228-67.
28 / 33

[3] Liu L, Ma W, Zhang Z. Macroscopic carbon nanotube assemblies: preparation, properties, and
potential applications. Small 2011;7(11):1504-20.
[4] Han JH, Zhang H, Chen MJ, Wang GR, Zhang Z. CNT buckypaper/thermoplastic polyurethane
composites with enhanced stiffness, strength and toughness. Compos Sci Technol 2014;103:63-71.
[5] Ma W, Liu L, Zhang Z, Yang R, Liu G, Zhang T, et al. High-strength composite fibers: realizing true
potential of carbon nanotubes in polymer matrix through continuous reticulate architecture and molecular
level couplings. Nano Lett 2009;9(8):2855-61.
[6] Liu L, Gao Y, Liu Q, Kuang J, Zhou D, Ju S, et al. High mechanical performance of layered graphene
oxide/poly(vinyl alcohol) nanocomposite films. Small 2013;9(14):2466-72.
[7] Dai ZH, Liu L, Qi X, Kuang J, Wei YG, Zhu HW, Zhang Z. Three-dimensional sponges with super
mechanical stability: harnessing true elasticity of individual carbon nanotubes in macroscopic
architectures. Sci Rep 2016;6:18930.
[8] Kauling AP, Seefeldt AT, Pisoni DP, Pradeep RC, Bentini R, Oliveira RVB, et al. The worldwide
graphene flake production. Adv Mater 2018;30(44):e1803784.
[9] Tehrani M, Khanbolouki P Carbon nanotubes: synthesis, characterization, and applications. Advances
in Nanomaterials: Springer; 2018. p. 3-35.
[10] Wagner HD, Vaia RA. Nanocomposites: issues at the interface. Mater Today 2004;7(11):38-42.
[11] Stankovich S, Dikin DA, Dommett GH, Kohlhaas KM, Zimney EJ, Stach EA, et al. Graphene-based
composite materials. Nature 2006;442(7100):282-6.
[12] Wang G, Dai Z, Wang Y, Tan P, Liu L, Xu Z, et al. Measuring interlayer shear stress in bilayer
graphene. Phys Rev Lett 2017;119(3):036101.
[13] Espinosa HD, Filleter T, Naraghi M. Multiscale experimental mechanics of hierarchical carbon-
based materials. Adv Mater 2012;24(21):2805-23.
[14] Gao Y, Li L, Tan P, Liu L, Zhang Z. Application of Raman spectroscopy in carbon nanotube-based
polymer composites. Chin Sci Bull 2010;55(35):3978-88.
[15] Filleter T, Espinosa H. Multi-scale mechanical improvement produced in carbon nanotube fibers by
irradiation cross-linking. Carbon 2013;56:1-11.
[16] Li X, Sun M, Shan C, Chen Q, Wei X. Mechanical properties of 2D materials studied by in situ
microscopy techniques. Adv Mater Interfaces 2018;5(5).
[17] Liechti KM. Characterizing the Interfacial Behavior of 2D Materials: a Review. Exp Mech
2019;59(3): 395-412.
[18] Ramanathan T, Abdala AA, Stankovich S, Dikin DA, Herrera-Alonso M, Piner RD, et al.
Functionalized graphene sheets for polymer nanocomposites. Nat Nanotechnol 2008;3(6):327-31.
[19] Spitalsky Z, Tasis D, Papagelis K, Galiotis C. Carbon nanotube–polymer composites: Chemistry,
processing, mechanical and electrical properties. Prog Polym Sci 2010;35(3):357-401.
[20] Coleman JN, Khan U, Blau WJ, Gun’ko YK. Small but strong: A review of the mechanical
properties of carbon nanotube–polymer composites. Carbon 2006;44(9):1624-52.
[21] Hu KS, Kulkarni DD, Choi I, Tsukruk VV. Graphene-polymer nanocomposites for structural and
functional applications. Prog Polym Sci 2014;39(11):1934-72.
[22] Mittal G, Dhand V, Rhee KY, Park S-J, Lee WR. A review on carbon nanotubes and graphene as
fillers in reinforced polymer nanocomposites. J Ind Eng Chem 2015;21:11-25.
[23] Roy S, Ryan J, Webster S, Nepal D. A review of in situ mechanical characterization of polymer
nanocomposites: prospect and challenges. Appl Mech Rev 2017;69(5): 050802.
[24] Lee C, Wei X, Kysar JW, Hone J. Measurement of the elastic properties and intrinsic strength of
monolayer graphene. Science 2008;321(5887):385-8.
[25] Filleter T, Bernal R, Li S, Espinosa HD. Ultrahigh strength and stiffness in cross-linked hierarchical
carbon nanotube bundles. Adv Mater 2011;23(25):2855-60.
[26] Cao K, Feng S, Han Y, Gao L, Hue Ly T, Xu Z, et al. Elastic straining of free-standing monolayer
graphene. Nat Commun 2020;11(1):284.
[27] Yu MF, Lourie O, Dyer MJ, Moloni K, Kelly TF, Ruoff RS. Strength and breaking mechanism of
multiwalled carbon nanotubes under tensile load. Science 2000;287(5453):637-40.
29 / 33

[28] Melanitis N, Galiotis C. Interfacial micromechanics in model composites using laser Raman-
spectroscopy. P Roy Soc Lond a Mat 1993;440(1909):379-98.
[29] Huang Y, Young RJ. Analysis of the fragmentation test for carbon-fibre/epoxy model composites by
means of Raman spectroscopy. Compos Sci Technol 1994;52(4):505-17.
[30] Piggott MR. Interfaces in composites. London: Elsevier Applied Science; 1991. P17, P60
[31] Pugno NM, Yin Q, Shi X, Capozza R. A generalization of the Coulomb’s friction law: from
graphene to macroscale. Meccanica 2013;48(8):1845-51.
[32] Jiang T, Huang R, Zhu Y. Interfacial sliding and buckling of monolayer graphene on a stretchable
substrate. Adv Funct Mater 2014;24(3):396-402.
[33] Dai Z, Wang G, Liu L, Hou Y, Wei Y, Zhang Z. Mechanical behavior and properties of hydrogen
bonded graphene/polymer nano-interfaces. Compos Sci Technol 2016;136:1-9.
[34] Wang G, Dai Z, Liu L, Hu H, Dai Q, Zhang Z. Tuning the interfacial mechanical behaviors of
monolayer graphene/PMMA nanocomposites. ACS Appl Mater Interfaces 2016;8(34):22554-62.
[35] Liu Y, Xu Z. Multimodal and self-healable interfaces enable strong and tough graphene-derived
materials. J Mech Phys Solids 2014;70:30-41.
[36] Dassios KG, Galiotis C. Polymer–nanotube interaction in MWCNT/poly(vinyl alcohol) composite
mats. Carbon 2012;50(11):4291-4.
[37] Zhang RF, Ning ZY, Zhang YY, Zheng QS, Chen Q, Xie HH, et al. Superlubricity in centimetres-
long double-walled carbon nanotubes under ambient conditions. Nat Nanotechnol 2013;8(12):912-6.
[38] Han E, Yu J, Annevelink E, Son J, Kang DA, Watanabe K, et al. Ultrasoft slip-mediated bending in
few-layer graphene. Nat Mater 2020;19(3):305-9.
[39] Chua P, Piggott M. The glass fibre polymer-interface: I—theoretical consideration for single fibre
pull-out tests. Compos Sci Technol 1985;22(1):33-42.
[40] Cooper CA, Cohen SR, Barber AH, Wagner HD. Detachment of nanotubes from a polymer matrix.
Applied Physics Letters 2002;81(20):3873.
[41] W. Ding AE, F. T. Fisher, X. Chen, D. A. Dikin, R. Andrews, L. C. Brinson, L. S. Schadler, and R. S.
Ruoff. Direct observation of polymer sheathing in carbon nanotube-polycarbonate composites. Nano Lett
2003;3(11):1593-7.
[42] Barber AH, Cohen SR, Wagner HD. Measurement of carbon nanotube–polymer interfacial strength.
Appl Phys Lett 2003;82(23):4140-2.
[43] Yang Y, Liang X, Chen W, Cao L, Li M, Sheldon BW, et al. Quantification and promotion of
interfacial interactions between carbon nanotubes and polymer derived ceramics. Carbon 2015;95:964-71.
[44] Kessman AJ, Zhang J, Vasudevan S, Lou J, Sheldon BW. Carbon nanotube pullout, interfacial
properties, and toughening in ceramic nanocomposites: mechanistic insights from single fiber pullout
analysis. Adv Mater Interfaces 2015;2(2):1400110.
[45] Barber AH, Cohen SR, Kenig S, Wagner HD. Interfacial fracture energy measurements for multi-
walled carbon nanotubes pulled from a polymer matrix. Compos Sci Technol 2004;64(15):2283-9.
[46] Wagner HD, Lourie O, Feldman Y, Tenne R. Stress-induced fragmentation of multiwall carbon
nanotubes in a polymer matrix. Appl Phys Lett 1998;72(2):188-90.
[47] Ganesan Y, Peng C, Lu Y, Loya PE, Moloney P, Barrera E, et al. Interface toughness of carbon
nanotube reinforced epoxy composites. ACS Appl Mater Interfaces 2011;3(2):129-34.
[48] Tsuda T, Ogasawara T, Deng F, Takeda N. Direct measurements of interfacial shear strength of
multi-walled carbon nanotube/PEEK composite using a nano-pullout method. Compos Sci Technol
2011;71(10):1295-300.
[49] Liu L, Barber AH, Nuriel S, Wagner HD. Mechanical Properties of Functionalized Single-Walled
Carbon-Nanotube/Poly(vinyl alcohol) Nanocomposites. Adv Funct Mater 2005;15(6):975-80.
[50] Wang G, Gao E, Dai Z, Liu L, Xu Z, Zhang Z. Degradation and recovery of graphene/polymer
interfaces under cyclic mechanical loading. Compos Sci Technol 2017;149(8):220-7.
[51] Gong L, Kinloch IA, Young RJ, Riaz I, Jalil R, Novoselov KS. Interfacial stress transfer in a
graphene monolayer nanocomposite. Adv Mater 2010;22(24):2694-7.
30 / 33

[52] Jin SY, Young RJ, Eichhorn SJ. Hybrid carbon fibre–carbon nanotube composite interfaces. Compos
Sci Technol 2014;95:114-20.
[53] Liu L, Li L, Gao Y, Tang L, Zhang Z. Single carbon fiber fracture embedded in an epoxy matrix
modified by nanoparticles. Compos Sci Technol 2013;77:101-9.
[54] Mohiuddin T, Lombardo A, Nair R, Bonetti A, Savini G, Jalil R, et al. Uniaxial strain in graphene by
Raman spectroscopy: G peak splitting, Grüneisen parameters, and sample orientation. Phys Rev B
2009;79(20):205433.
[55] Androulidakis C, Koukaras EN, Frank O, Tsoukleri G, Sfyris D, Parthenios J, et al. Failure processes
in embedded monolayer graphene under axial compression. Sci Rep 2014;4:5271.
[56] Wang W, Dai S, Li X, Yang J, Srolovitz DJ, Zheng Q. Measurement of the cleavage energy of
graphite. Nat Commun 2015;6:7853-9.
[57] Kis A, Jensen K, Aloni S, Mickelson W, Zettl A. Interlayer forces and ultralow sliding friction in
multiwalled carbon nanotubes. Phys Rev Lett 2006;97(2):025501.
[58] Filleter T, Yockel S, Naraghi M, Paci JT, Compton OC, Mayes ML, et al. Experimental-
computational study of shear interactions within double-walled carbon nanotube bundles. Nano Lett
2012;12(2):732-42.
[59] Bai Y, Zhang R, Ye X, Zhu Z, Xie H, Shen B, et al. Carbon nanotube bundles with tensile strength
over 80 GPa. Nat Nanotechnol 2018;13(7):589-95.
[60] Urbakh M. Towards macroscale superlubricity. Nat Nanotechnol 2013;8(12):893-4.
[61] Boggild P. The war on fake graphene. Nature 2018;562(7728):502-3.
[62] Gong L, Young RJ, Kinloch IA, Haigh SJ, Warner JH, Hinks JA, et al. Reversible loss of bernal
stacking during the deformation of few-layer graphene in nanocomposites. ACS Nano 2013;7(8):7287-94.
[63] Frank O, Bousa M, Riaz I, Jalil R, Novoselov KS, Tsoukleri G, et al. Phonon and Structural Changes
in Deformed Bernal Stacked Bilayer Graphene. Nano Lett 2012;12(2):687-93.
[64] Ma W, Liu L, Yang R, Zhang T, Zhang Z, Song L, et al. Monitoring a micromechanical process in
macroscale carbon nanotube films and fibers. Adv Mater 2009;21(5):603-8.
[65] Androulidakis C, Koukaras EN, Paterakis G, Trakakis G, Galiotis C. Tunable macroscale structural
superlubricity in two-layer graphene via strain engineering. Nat Commun 2020;11(1).
[66] Tan PH, Han WP, Zhao WJ, Wu ZH, Chang K, Wang H, et al. The shear mode of multilayer
graphene. Nat Mater 2012;11(4):294-300.
[67] Wu JB, Lin ML, Cong X, Liu HN, Tan PH. Raman spectroscopy of graphene-based materials and its
applications in related devices. Chem Soc Rev 2018;47(5):1822-73.
[68] Wang H, Feng M, Zhang X, Tan P-H, Wang Y. In-phase family and self-similarity of interlayer
vibrational frequencies in van der Waals layered materials. J Physl Chem C 2015;119(12):6906-11.
[69] Zhang X, Han WP, Wu JB, Milana S, Lu Y, Li QQ, et al. Raman spectroscopy of shear and layer
breathing modes in multilayer MoS2. Phys Rev B 2013;87(11):115413.
[70] Wang G, Li X, Wang Y, Zheng Z, Dai Z, Qi X, et al. Interlayer coupling behaviors of boron doped
multilayer graphene. J Physl Chem C 2017;121(46):26034–43.
[71] Dienwiebel M, Verhoeven GS, Pradeep N, Frenken JW, Heimberg JA, Zandbergen HW.
Superlubricity of graphite. Phys Rev Lett 2004;92(12):126101.
[72] Liu Z, Yang J, Grey F, Liu JZ, Liu Y, Wang Y, et al. Observation of microscale superlubricity in
graphite. Phys Rev Lett 2012;108(20):205503.
[73] Liu SW, Wang HP, Xu Q, Ma TB, Yu G, Zhang C, et al. Robust microscale superlubricity under
high contact pressure enabled by graphene-coated microsphere. Nat Commun 2017;8:14029.
[74] Daly M, Cao C, Sun H, Sun Y, Filleter T, Singh CV. Interfacial shear strength of multilayer
graphene oxide films. ACS Nano 2016;10(2):1939-47.
[75] Roy D, Bhattacharyya S, Rachamim A, Plati A, Saboungi M-L. Measurement of interfacial shear
strength in single wall carbon nanotubes reinforced composite using Raman spectroscopy. J Appl Phys
2010;107(4):043501.
[76] Barber AH, Cohen SR, Eitan A, Schadler LS, Wagner HD. Fracture transitions at a carbon-
nanotube/polymer interface. Adv Mater 2006;18(1):83-7.
31 / 33

[77] Gong L, Young RJ, Kinloch IA, Riaz I, Jalil R, Novoselov KS. Optimizing the reinforcement of
polymer-based nanocomposites by graphene. ACS Nano 2012;6(3):2086-95.
[78] Young RJ, Gong L, Kinloch IA, Riaz I, Jalil R, Novoselov KS. Strain mapping in a graphene
monolayer nanocomposite. ACS Nano 2011;5(4):3079-84.
[79] Anagnostopoulos G, Androulidakis C, Koukaras EN, Tsoukleri G, Polyzos I, Parthenios J, et al.
Stress transfer mechanisms at the submicron level for graphene/polymer systems. ACS Appl Mater
Interfaces 2015;7(7):4216-23.
[80] Androulidakis C, Koukaras EN, Rahova J, Sampathkumar K, Parthenios J, Papagelis K, et al.
Wrinkled Few-Layer Graphene as Highly Efficient Load Bearer. ACS Appl Mater Interfaces
2017;9(31):26593-601.
[81] Zhao X, Papageorgiou DG, Zhu L, Ding F, Young RJ. The strength of mechanically-exfoliated
monolayer graphene deformed on a rigid polymer substrate. Nanoscale 2019;11(30):14339-53.
[82] Manikas AC, Pastore Carbone MG, Woods CR, Wang Y, Souli I, Anagnostopoulos G, et al. Stress
transfer at the nanoscale on graphene ribbons of regular geometry. Nanoscale 2019;11(30):14354-61.
[83] Xu C, Xue T, Qiu W, Kang Y. Size effect of the interfacial mechanical behavior of graphene on a
stretchable substrate. ACS Appl Mater Interfaces 2016;8(40):27099-106.
[84] Li X, Warzywoda J, McKenna GB. Mechanical responses of a polymer graphene-sheet nano-
sandwich. Polymer 2014;55(19):4976-82.
[85] Min-Feng Yu BIY, Rodney S. Ruoff. Controlled sliding and pullout of nested shells in individual
multiwalled carbon nanotubes. J Phys Chem B 2000;104:8764-7.
[86] Bhushan B, Ling X, Jungen A, Hierold C. Adhesion and friction of a multiwalled carbon nanotube
sliding against single-walled carbon nanotube. Phys Rev B 2008;77(16).
[87] Liu Z, Zhang SM, Yang JR, Liu JZ, Yang YL, Zheng QS. Interlayer shear strength of single
crystalline graphite. Acta Mech Sin 2012;28(4):978-82.
[88] Wei X, Meng Z, Ruiz L, Xia W, Lee C, Kysar JW, et al. Recoverable slippage mechanism in
multilayer graphene leads to repeatable energy dissipation. ACS Nano 2016;10(2):1820-8.
[89] Han J, Ryu S, Kim D-K, Woo W, Sohn D. Effect of interlayer sliding on the estimation of elastic
modulus of multilayer graphene in nanoindentation simulation. EPL (Europhysics Letters)
2016;114(6):68001.
[90] Wang G, Dai Z, Xiao J, Feng S, Weng C, Liu L, et al. Bending of multilayer van der Waals materials.
Phys Rev Lett 2019;123(11):116101.
[91] Mansfield EH. The bending and stretching of plates: Cambridge university press; 2005.
[92] Peng W, Gao W, Cao Z, Liechti KM, Huang R. Numerical analysis of circular graphene bubbles. J
Appl Mech 2013;(80):040905.
[93] Pan F, Wang G, Liu L, Chen Y, Zhang Z, Shi X. Bending induced interlayer shearing, rippling and
kink buckling of multilayered graphene sheets. J Mech Phys Solids 2019;122:340-63.
[94] Owais M, Zhao J, Imani A, Wang G, Zhang H, Zhang Z. Synergetic effect of hybrid fillers of boron
nitride, graphene nanoplatelets, and short carbon fibers for enhanced thermal conductivity and electrical
resistivity of epoxy nanocomposites. Compos Part A Appl Sci Manuf 2019;117:11-22.
[95] Hamidinejad M, Zandieh A, Lee JH, Papillon J, Zhao B, Moghimian N, et al. Insight into the
directional thermal transport of hexagonal boron nitride composites. ACS Appl Mater Interfaces
2019;11(44):41726-35.
[96] Zhang Y, Liu L, Sun B, Wang G, Zhang Z. Preparation of lipophilic graphene oxide derivates via a
concise route and its mechanical reinforcement in thermoplastic polyurethane. Compos Sci Technol
2016;134:36-42.
[97] Najafi F, Wang G, Mukherjee S, Cui T, Filleter T, Singh CV. Toughening of graphene-based
polymer nanocomposites via tuning chemical functionalization. Compos Sci Technol 2020;194:108140.
[98] Dai Z, Wang G, Zheng Z, Wang Y, Zhang S, Qi X, et al. Mechanical responses of boron-doped
monolayer graphene. Carbon 2019;147:594-601.
[99] Gao E, Cao Y, Liu Y, Xu Z. Optimizing interfacial cross-linking in graphene-derived materials,
which balances intralayer and interlayer load transfer. ACS Appl Mater Interfaces 2017;9(29):24830-9.
32 / 33

[100] Shang J, Chen Y, Zhou Y, Liu L, Wang G, Li X, et al. Effect of folded and crumpled morphologies
of graphene oxide platelets on the mechanical performances of polymer nanocomposites. Polymer
2015;68:131-9.
[101] Li Z, Young RJ, Papageorgiou DG, Kinloch IA, Zhao X, Yang C, et al. Interfacial stress transfer in
strain engineered wrinkled and folded graphene. 2D Mater 2019;6: 045026.
[102] Li Y, Zhou P, An F, Liu Y, Lu C. Dynamic self-stiffening and structural evolutions of
polyacrylonitrile/carbon nanotube nanocomposites. ACS Appl Mater Interfaces 2017;9(6):5653-9.
[103] Dai ZH, Wang YL, Liu LQ, Liu XL, Tan PH, Xu ZP, et al. Hierarchical graphene-based films with
dynamic self-stiffening for biomimetic artificial muscle. Adv Funct Mater 2016;26(38):7003-10.
[104] Chen Y, Dai Z, Weng C, Wang G, Liu X, Cong X, et al. Engineering the interface in mechanically
responsive graphene-based films. RSC Adv 2018;8(63):36257-63.
[105] Cui T, Yip K, Hassan A, Wang G, Liu X, Sun Y, Filleter T. Graphene fatigue through van der
Waals interactions. Sci Adv 2020;6:eabb1335.
[106] Dai Z, Lu N, Liechti KM, Huang R. Mechanics at the interfaces of 2D materials: Challenges and
opportunities. Curr Opin Solid State Mater Sci 2020;24(4): 100837.
[107] He Z, Zhu Y, Wu H. Edge effect on interlayer shear in multilayer two-dimensional material
assemblies. Int J Solids Struct 2020;204:128-37.
[108] Liu M, Ye X, Bai Y, Zhang R, Wei F, Li X. Multi-scale analysis of the interaction in ultra-long
carbon nanotubes and bundles. J Mech Phys Solids 2020;142:104032.
[109] Zhang Y, Gong S, Zhang Q, Ming P, Wan S, Peng J, et al. Graphene-based artificial nacre
nanocomposites. Chem Soc Rev 2016;45(9):2378-95.
33 / 33

Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

Guorui Wang: Writing, Visualization, Conceptualization. Luqi Liu: Conceptualization, Writing, Supervision,
Project administration. Zhong Zhang: Project administration.

You might also like